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

Rational Solutions of First Order Algebraic Ordinary Differential Equations

Ruyong Feng and Shuang Feng KLMM,Academy of Mathematics and Systems Science, Chinese Academy of Sciences and School of Mathematics, University of Chinese Academy of Sciences, 100190, Beijing
ryfeng@amss.ac.cn, fengshuang15@mails.ucas.ac.cn
Abstract

Let f(t,y,y)=i=0dai(t,y)yi=0f(t,y,y^{\prime})=\sum_{i=0}^{d}a_{i}(t,y)y^{\prime i}=0 be a first order ordinary differential equation with polynomial coefficients. Eremenko in 1999 proved that there exists a constant CC such that every rational solution of f(t,y,y)=0f(t,y,y^{\prime})=0 is of degree not greater than CC. Examples show that this degree bound CC depends not only on the degrees of ff in t,y,yt,y,y^{\prime} but also on the coefficients of ff viewed as polynomial in t,y,yt,y,y^{\prime}. In this paper, we show that if

maxi=0d{deg(ai,y)2(di)}>0\max_{i=0}^{d}\{\deg(a_{i},y)-2(d-i)\}>0

then the degree bound CC only depends on the degrees of ff, and furthermore we present an explicit expression for CC in terms of the degrees of ff.

keywords:
first order AODE, rational solution, degree bound, height
MSC:
[2010] 34A05, 68W30
journal: Journal of  Templatesmytitlenotemytitlenotefootnotetext: This work was supported by NSFC under Grants No.11771433 and No.11688101, and by Beijing Natural Science Foundation (Z190004).

1 Introduction

The study of first order algebraic ordinary differential equations (AODEs in short) has a long history, which can be at least tracked back to the time of Fuchs and Poincaré. Fuchs presented a sufficient and necessary condition so called Fuchs’ criterion for a first order AODE having no movable singularities. Roughly speaking, an AODE is said to have movable singularities if it has a solution (with arbitrary constants) whose branch points depend on arbitrary constants. For instance the solution y=t+cy=\sqrt{t+c} of 2yy1=02yy^{\prime}-1=0 has branch points t=ct=-c, where cc is an arbitrary constant, so 2yy1=02yy^{\prime}-1=0 has movable singularities. Based on differential algebra developed by Ritt [15] and the theory of algebraic function field of one variable, Matsuda [14] reproduced many classic results of first order AODEs. In particular, he presented an algebraic definition of movable singularities. In 1999, combining Matsuda’s results and height estimates of points on plane algebraic curves, Eremenko showed that rational solutions of first order AODEs have bounded degrees. In [7], we proved that if a first order AODE has movable singularities then it has only finitely many rational solutions. As for algebraic solutions of first order AODEs, Freitag and Moosa [8] showed that they are of bounded heights.

On the other hand, the algorithmic aspects of computing closed form solutions of AODEs have been extensively studied in the past decades. Several algorithms have been developed for computing closed form solutions (e.g. liouvillian solutions) of linear homogeneous differential equations (see [2, 12, 19, 17, 18] etc). Yet, the situation is different in the nonlinear case. Existing algorithms were only valid for AODEs of special types. Based on parametrization of algebraic curves, Aroca et al [1, 6] gave two complete methods for finding rational and algebraic solutions of first order autonomous AODEs. Their methods were generalized by Winkler and his collegues to the class of first order non-autonomous AODEs whose rational general solutions involve arbitrary constants rationally as well as some other certain classes of AODEs (see [21, 20, 3, 22] etc). Particularly, in [21], the authors introduced a class of first order AODEs called maximally comparable AODEs and presented an algorithm to compute a degree bound for rational solutions of this kind of equations as well as first order quasi-linear AODEs. Readers are referred to [22] for a survey of recent developments in this direction. Theoretically, it suffices to compute a degree bound for all rational solutions of a first order AODE to find all its rational solutions. The following example implies that the degrees of rational solutions may depend not only on the degrees of the original equation but also on its constant coefficients.

Example 1.1.

Let nn be an integer. Then y=tny=t^{n} is a rational solution of tyny=0ty^{\prime}-ny=0. The degree of tnt^{n} depends on the constant coefficient nn of tynyty^{\prime}-ny.

Let f=i=0dai(t,y)yi=0f=\sum_{i=0}^{d}a_{i}(t,y)y^{\prime i}=0 be an irreducible first order AODE. Set

m.s.index(f)=maxi=0d{deg(ai,y)2(di)}.{\rm m.s.index}(f)=\max_{i=0}^{d}\{\deg(a_{i},y)-2(d-i)\}.

Fuchs’ criterion (see Remark on page 14 of [14]) implies that f=0f=0 has movable singularities if m.s.index(f)>0{\rm m.s.index}(f)>0. On the other hand, it was proved in [5] that if f=0f=0 has movable singularities then it can be transferred into an AODE gg with positive m.s.index{\rm m.s.index}. This motivates us to focus on first order AODEs with positive m.s.index{\rm m.s.index}. We prove that for an irreducible first order AODE f=0f=0 with m.s.index(f)>0{\rm m.s.index}(f)>0 the degrees of rational solutions of f=0f=0 are independent of the constant coefficients of ff and furthermore we present an explicit degree bound in terms of the degrees of ff. The key step to obtain this degree bound is to estimate the heights of points on plane algebraic curves. This height estimate is a special case of the result about heights on complete nonsingular varieties (see for instance Proposition 3 on page 89 of [13]). Eremenko in [5] provided a simple proof for this special case based on the Riemann-Roch Theorem. We follow Eremenko’s proof but present explicit bounds for each step.

The paper is organized as follows. In Section 2, we introduce some basic materials used in the later sections. In Sections 3, we estimate the degrees and heights for elements in a Riemann-Roth space. In Section 4, we present an explicit bound for the heights of points on a plane algebraic curve. Finally, in Section 5, we apply the results in Section 4 to first order AODEs.

Throughout this paper, \mathbb{Z} stands for the ring of integers, k,Kk,K for algebraically closed fields of characteristic zero, and RR and {\mathcal{R}} for algebraic function fields over kk and KK respectively. m()\mathbb{P}^{m}(\cdot) denotes the projective space of dimension mm over a field and 𝕍()\mathbb{V}(\cdot) denotes the variety in a projective space defined by a set of homogeneous polynomials.

2 Basic materials

In this section, we will introduce some basic materials used in this paper, including differential rings, algebraic function fields of one variable and heights. Readers are referred to [15, 14, 4, 13] for details.

2.1 Differential fields associated to AODEs

In this subsection, we introduce some basic notations of differential algebra.

Definition 2.1.

A derivation on a ring {\mathcal{R}} is a map δ:\delta:{\mathcal{R}}\rightarrow{\mathcal{R}} satisfying that for all a,ba,b\in{\mathcal{R}},

δ(a+b)=δ(a)+δ(b),δ(ab)=δ(a)b+aδ(b).\delta(a+b)=\delta(a)+\delta(b),\,\,\delta(ab)=\delta(a)b+a\delta(b).

A ring (resp. field) equipped with a derivation is called a differential ring (resp. differential field). An ideal II\subset{\mathcal{R}} is called a differential ideal if δ(I)I\delta(I)\subset I.

The field k(t)k(t) of rational functions in tt can be endowed with a structure of differential field whose derivation δ\delta is the usual derivation with respect to tt, i.e. δ=ddt\delta=\frac{\rm d}{{\rm d}t}. Set y0=yy_{0}=y and denote

k(t){y}=k(t)[y0,y1,]k(t)\{y\}=k(t)[y_{0},y_{1},\dots]

where y0,y1,y_{0},y_{1},\dots are indeterminates. One can extend the derivation δ\delta on k(t)k(t) to a derivation δ\delta^{\prime} on k(t){y}k(t)\{y\} by assigning yi=δi(y0)y_{i}=\delta^{\prime i}(y_{0}) so that k(t){y}k(t)\{y\} becomes a differential ring. For the sake of notations, we use δ\delta in place of δ\delta^{\prime}. Elements in k(t){y}k(t)\{y\} are called differential polynomials over k(t)k(t). Let ff be a differential polynomial not in k(t)k(t). Then there is a unique integer dd such that fk(t)[y0,,yd]k(t)[y0,,yd1]f\in k(t)[y_{0},\dots,y_{d}]\setminus k(t)[y_{0},\dots,y_{d-1}]. This integer is called the order of ff. We shall use [][\cdot] (resp. \langle\cdot\rangle) to stand for the differential (resp. algebraic) ideal generated by a set of differential polynomials (resp. polynomials) respectively. Suppose that ff is irreducible viewed as an algebraic polynomial. Set

Σf={Ak(t){y}|m>0s.t.SmAm[f]}\Sigma_{f}=\left\{A\in k(t)\{y\}|\,\exists\,m>0\,\,\mbox{s.t.}\,S^{m}A^{m}\in[f]\right\}

where S=f/ydS=\partial f/\partial y_{d} and dd is the order of ff. It was proved on page 30 of [15] that Σf\Sigma_{f} is a prime differential ideal and so k(t){y}/Σfk(t)\{y\}/\Sigma_{f} is a differential domain. Lemma 2.2 of [7] implies that the field of fractions of k(t){y}/Σfk(t)\{y\}/\Sigma_{f} is isomorphic to that of k(t)[y0,,yd]/fk(t)[y_{0},\dots,y_{d}]/\langle f\rangle. Under this isomorphism, the field of fractions of k(t)[y0,,yd]/fk(t)[y_{0},\dots,y_{d}]/\langle f\rangle can be endowed with a structure of differential field. We shall still use δ\delta, or in short, to denote the induced derivation on the field of fractions of k(t)[y0,,yd]/fk(t)[y_{0},\dots,y_{d}]/\langle f\rangle.

In this paper, the first order AODEs under consideration are differential equations of the following form

f(y,y)=0f(y,y^{\prime})=0 (1)

where f(y,y)k(t)[y,y]k(t)f(y,y^{\prime})\in k(t)[y,y^{\prime}]\setminus k(t).

Definition 2.2.

An element r(t)k(t)r(t)\in k(t) satisfying f(r(t),r(t))=0f(r(t),r^{\prime}(t))=0 is called a rational solution of f(y,y)=0f(y,y^{\prime})=0.

Remark that the derivation δ\delta in k(t)k(t) can be uniquely extended to a derivation in k(t)¯{\overline{k(t)}} which we shall still denote by δ\delta. Assume that viewed as a polynomial in k(t)¯[y,y]{\overline{k(t)}}[y,y^{\prime}], ff is irreducible over k(t)¯\overline{k(t)}. Then the field of fractions of k(t)¯[y,y]/f(y,y){\overline{k(t)}}[y,y^{\prime}]/\langle f(y,y^{\prime})\rangle is not only an algebraic function field over k(t)¯{\overline{k(t)}} but also a differential field.

2.2 Algebraic function fields of one variable

Let KK be an algebraically closed field of characteristic zero and {\mathcal{R}} an extension field of KK. We say {\mathcal{R}} is an algebraic function field of one variable over KK if {\mathcal{R}} satisfies the following conditions: there is an element aa of {\mathcal{R}} which is transcendental over KK, and {\mathcal{R}} is algebraic of finite degree over K(a)K(a). Assume {\mathcal{R}} is an algebraic function field of one variable over KK. A valuation ring of {\mathcal{R}} over KK is a subring VV satisfying that

  1. 1.

    KVK\subset V\neq{\mathcal{R}}; and

  2. 2.

    if aVa\in{\mathcal{R}}\setminus V, then a1Va^{-1}\in V.

All non-invertible elements of VV form a maximal ideal 𝔓{\mathfrak{P}} which is called a place of {\mathcal{R}}, and VV is called the corresponding valuation ring of 𝔓{\mathfrak{P}}. Let VV be a valuation ring with 𝔓{\mathfrak{P}} as place. There is an element uVu\in V, called a local uniformizer of 𝔓{\mathfrak{P}} or VV, such that 𝔓=uV{\mathfrak{P}}=uV and n=1unV={0}\bigcap_{n=1}^{\infty}u^{n}V=\{0\}. The factor ring V/𝔓V/{\mathfrak{P}} is equal to KK since KK is algebraically closed. For every valuation ring VV with place 𝔓{\mathfrak{P}}, we define a map

π𝔓:K{}\pi_{{\mathfrak{P}}}:{\mathcal{R}}\longrightarrow K\cup\{\infty\}

satisfying if aVa\in V then π𝔓(a)=a+𝔓V/𝔓=K\pi_{{\mathfrak{P}}}(a)=a+{\mathfrak{P}}\in V/{\mathfrak{P}}=K, otherwise π𝔓(a)=\pi_{{\mathfrak{P}}}(a)=\infty. It is well-known that {\mathcal{R}} admits infinitely many places, and there is one-to-one correspondence between places and valuation rings.

Let 𝔓{\mathfrak{P}} be a place of {\mathcal{R}} and VV the corresponding valuation ring of 𝔓{\mathfrak{P}}. Let uu be a local uniformizer of 𝔓{\mathfrak{P}}. Then for every non-zero element aa of {\mathcal{R}}, there is a unique integer nn such that

a=unva=u^{n}v

for some invertible element vVv\in V. It is easy to see that the integer nn is independent of the choice of local uniformizers. Such nn is called the order of aa at 𝔓{\mathfrak{P}} and denoted by ν𝔓(a)\nu_{{\mathfrak{P}}}(a). We make the convention to write ν𝔓(0)=\nu_{{\mathfrak{P}}}(0)=\infty. Then the place 𝔓{\mathfrak{P}} induces a map ν𝔓\nu_{{\mathfrak{P}}} from {\mathcal{R}} to \mathbb{Z} sending aa to ν𝔓(a)\nu_{{\mathfrak{P}}}(a). This map ν𝔓\nu_{{\mathfrak{P}}} is called the order function at 𝔓{\mathfrak{P}}. For a,ba,b\in{\mathcal{R}}, we have

ν𝔓(ab)=ν𝔓(a)+ν𝔓(b),ν𝔓(a+b)min{ν𝔓(a),ν𝔓(b)}\nu_{{\mathfrak{P}}}(ab)=\nu_{{\mathfrak{P}}}(a)+\nu_{{\mathfrak{P}}}(b),\,\,\nu_{{\mathfrak{P}}}(a+b)\geq\min\{\nu_{{\mathfrak{P}}}(a),\nu_{{\mathfrak{P}}}(b)\}

where the equality in the later formula holds if ν𝔓(a)ν𝔓(b)\nu_{{\mathfrak{P}}}(a)\neq\nu_{{\mathfrak{P}}}(b). Let aa\in{\mathcal{R}} and 𝔓{\mathfrak{P}} be a place. We say 𝔓{\mathfrak{P}} is a zero of aa if ν𝔓(a)>0\nu_{{\mathfrak{P}}}(a)>0, and a pole of aa if ν𝔓(a)<0\nu_{{\mathfrak{P}}}(a)<0. Every non-zero element of {\mathcal{R}} admits only finitely many zeros and poles.

A divisor in {\mathcal{R}} is a formal sum

D=𝔓n𝔓𝔓D=\sum_{{\mathfrak{P}}}n_{{\mathfrak{P}}}{\mathfrak{P}}

for all the places of {\mathcal{R}}, where n𝔓n_{{\mathfrak{P}}}\in\mathbb{Z} and n𝔓=0n_{{\mathfrak{P}}}=0 for all but finitely many 𝔓{\mathfrak{P}}. It is easy to see that the set of divisors in {\mathcal{R}} forms an abelian group. DD is effective if n𝔓0n_{{\mathfrak{P}}}\geq 0 for all 𝔓{\mathfrak{P}}. The degree of DD, denoted by deg(D)\deg(D), is defined to be n𝔓\sum n_{{\mathfrak{P}}} and the support of DD, denoted by supp(D){\rm supp}(D), is defined to be {𝔓|n𝔓0}\{{\mathfrak{P}}\,|\,n_{{\mathfrak{P}}}\neq 0\}. For brief, we denote

D+=n𝔓>0n𝔓𝔓,D=n𝔓<0n𝔓𝔓.D^{+}=\sum_{n_{\mathfrak{P}}>0}n_{{\mathfrak{P}}}{\mathfrak{P}},\quad D^{-}=\sum_{n_{{\mathfrak{P}}}<0}-n_{{\mathfrak{P}}}{\mathfrak{P}}.

Let D1=𝔓n𝔓𝔓D_{1}=\sum_{{\mathfrak{P}}}n_{{\mathfrak{P}}}{\mathfrak{P}} and D2=𝔓m𝔓𝔓D_{2}=\sum_{{\mathfrak{P}}}m_{{\mathfrak{P}}}{\mathfrak{P}} be two divisors in {\mathcal{R}}, we write D1D2D_{1}\geq D_{2} provided D1D2D_{1}-D_{2} is effective. For every non-zero element aa of {\mathcal{R}}, we denote

div(a)=𝔓ν𝔓(a)𝔓{\rm div}(a)=\sum_{{\mathfrak{P}}}\nu_{{\mathfrak{P}}}(a){\mathfrak{P}}

where 𝔓{\mathfrak{P}} ranges over all places of {\mathcal{R}}. Then div(a){\rm div}(a) is a divisor of degree 0. For a divisor DD, we denote

𝔏(D)={adiv(a)+D0}{0},{\mathfrak{L}}(D)=\{a\in{\mathcal{R}}\mid{\rm div}(a)+D\geq 0\}\cup\{0\},

which is called the Riemann-Roch space of DD. It is well-known that each Riemann-Roch space is a KK-vector space of finite dimension. The Riemann-Roch Theorem implies that if DD is a divisor whose degree is not less than the genus of {\mathcal{R}} then 𝔏(D){\mathfrak{L}}(D) is of positive dimension.

Let fK[x0,x1]Kf\in K[x_{0},x_{1}]\setminus K be irreducible. One sees that the field of fractions of K[x0,x1]/fK[x_{0},x_{1}]/\langle f\rangle is an algebraical function field of one variable over KK which is called the algebraic function field of ff. For an irreducible homogeneous polynomial FF in K[x0,x1,x2]K[x_{0},x_{1},x_{2}], the corresponding algebraic function field is defined to be the algebraic function field of F(x0,x1,1)F(x_{0},x_{1},1). Remark that the algebraic function fields of F(1,x1,x2),F(x0,1,x2)F(1,x_{1},x_{2}),F(x_{0},1,x_{2}) and F(x0,x1,1)F(x_{0},x_{1},1) are all isomorphic.

2.3 Models of algebraic function fields of one variable

Let {\mathcal{R}} be an algebraic function field of one variable over KK. The set of all places of {\mathcal{R}} can be viewed as a nonsingular model of {\mathcal{R}}. On the other hand, let FF be an irreducible homogeneous polynomial FK[x0,x1,x2]F\in K[x_{0},x_{1},x_{2}] whose algebraic function field is {\mathcal{R}}. Then the projective curve F=0F=0 is another model of {\mathcal{R}}. There is a surjective map from a nonsingular model of {\mathcal{R}} to the curve F=0F=0. To describe this map precisely, let ξ0,ξ1,ξ2\xi_{0},\xi_{1},\xi_{2} be three nonzero elements of {\mathcal{R}} satisfying that

=K(ξ0/ξ2,ξ1/ξ2)andF(ξ0,ξ1,ξ2)=0.{\mathcal{R}}=K(\xi_{0}/\xi_{2},\xi_{1}/\xi_{2})\,\,\mbox{and}\,\,F(\xi_{0},\xi_{1},\xi_{2})=0.

Set 𝝃=(ξ0,ξ1,ξ2){\bm{\xi}}=(\xi_{0},\xi_{1},\xi_{2}). Let 𝔓{\mathfrak{P}} be a place of {\mathcal{R}} with uu as local uniformizer. Denote by =mini{ν𝔓(ξi)}\ell=\min_{i}\{\nu_{{\mathfrak{P}}}(\xi_{i})\}. One sees that ν𝔓(uξi)0\nu_{{\mathfrak{P}}}(u^{-\ell}\xi_{i})\geq 0 and moreover not all π𝔓(uξi)\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{i}) are zero. Therefore (π𝔓(uξ0),π𝔓(uξ1),π𝔓(uξ2))(\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{0}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{1}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{2})) defines a point of 2(K)\mathbb{P}^{2}(K). Remark that this point does not depend on the choice of uu.

Definition 2.3.

We call (π𝔓(uξ0),π𝔓(uξ1),π𝔓(uξ2))(\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{0}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{1}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{2})) the center of 𝔓{\mathfrak{P}} with respect to 𝛏{\bm{\xi}}. Denote by 𝒞(𝛏){\mathcal{C}}({\bm{\xi}}) the set of centers with respect to 𝛏{\bm{\xi}}.

We claim that 𝒞(𝝃){\mathcal{C}}({\bm{\xi}}) is the plane projective curve in 2(K)\mathbb{P}^{2}(K) defined by FF and the map sending 𝔓{\mathfrak{P}} to the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}} is the required map. It is easy to verify that F(𝐜)=0F({\bf c})=0 for all 𝐜𝒞(𝝃){\bf c}\in{\mathcal{C}}({\bm{\xi}}). Conversely, let (c0,c1,c2)(c_{0},c_{1},c_{2}) be a point of F=0F=0. Without loss of generality, we may assume that c00c_{0}\neq 0. Then F(1,c1/c0,c2/c0)=0F(1,c_{1}/c_{0},c_{2}/c_{0})=0. Remark tha =K(ξ1/ξ0,ξ2/ξ0){\mathcal{R}}=K(\xi_{1}/\xi_{0},\xi_{2}/\xi_{0}). As F(1,ξ1/ξ0,ξ2/ξ0)=0F(1,\xi_{1}/\xi_{0},\xi_{2}/\xi_{0})=0, due to Corollary 2 on page 8 of [4], there is a place 𝔓{\mathfrak{P}} containing ξ1/ξ0c1/c0\xi_{1}/\xi_{0}-c_{1}/c_{0} and ξ2/ξ0c2/c0\xi_{2}/\xi_{0}-c_{2}/c_{0}. For this place, one has that ν𝔓(ξ1)ν𝔓(ξ0),ν𝔓(ξ2)ν𝔓(ξ0)\nu_{{\mathfrak{P}}}(\xi_{1})\geq\nu_{{\mathfrak{P}}}(\xi_{0}),\nu_{{\mathfrak{P}}}(\xi_{2})\geq\nu_{{\mathfrak{P}}}(\xi_{0}) and furthermore π𝔓(ξi/ξ0)=ci/c0\pi_{{\mathfrak{P}}}(\xi_{i}/\xi_{0})=c_{i}/c_{0}. Write =ν𝔓(ξ0)\ell=\nu_{{\mathfrak{P}}}(\xi_{0}). Then the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}} is

(π𝔓(uξ0),π𝔓(uξ1),π𝔓(uξ2))=π𝔓(uξ0)(1,c1/c0,c2/c0).\displaystyle(\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{0}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{1}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{2}))=\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{0})(1,c_{1}/c_{0},c_{2}/c_{0}).

This implies that (c0,c1,c2)𝒞(𝝃)(c_{0},c_{1},c_{2})\in{\mathcal{C}}({\bm{\xi}}).

Definition 2.4.

We call 𝒞(𝛏){\mathcal{C}}({\bm{\xi}}) or F=0F=0 a plane projective model of {\mathcal{R}}.

The plane projective models of {\mathcal{R}} usually have singularities. Let 𝒞{\mathcal{C}} be an irreducible projective curve in 2(K)\mathbb{P}^{2}(K) defined by a homogeneous polynomial FF. A point 𝐜{\bf c} of 𝒞{\mathcal{C}} is said to be of multiplicity rr, if all derivatives of FF up to and including the (r1)(r-1)-th vanish at 𝐜{\bf c} but not all the rr-th derivatives vanish at 𝐜{\bf c}. Suppose that 𝐜{\bf c} is a point of 𝒞{\mathcal{C}} with multiplicity rr. If r=1r=1, then 𝐜{\bf c} is called a simple point of 𝒞{\mathcal{C}}, otherwise a singular point of 𝒞{\mathcal{C}}. A point of multiplicity rr is ordinary if the rr tangents to 𝒞{\mathcal{C}} at this point are distinct, otherwise it is non-ordinary. Due to Propositon on page of [9], {\mathcal{R}} has always a plane projective model with only ordinary singularities.

Let Φ=(ϕ0,ϕ1,ϕ2)\Phi=(\phi_{0},\phi_{1},\phi_{2}) be an invertible transformation, where ϕ0,ϕ1,ϕ2\phi_{0},\phi_{1},\phi_{2} are homogeneous polynomials in K[x0,x1,x2]K[x_{0},x_{1},x_{2}] of the same degree and they have no common factors. We further assume that ϕi(𝝃)0\phi_{i}({\bm{\xi}})\neq 0 for all i=0,1,2i=0,1,2. Then

=K(ϕ0(𝝃)ϕ2(𝝃),ϕ1(𝝃)ϕ2(𝝃)).{\mathcal{R}}=K\left(\frac{\phi_{0}({\bm{\xi}})}{\phi_{2}({\bm{\xi}})},\frac{\phi_{1}({\bm{\xi}})}{\phi_{2}({\bm{\xi}})}\right).
Proposition 2.5.

Let Φ,𝛏\Phi,{\bm{\xi}} be as above and 𝔓{\mathfrak{P}} a place of {\mathcal{R}}. Assume that 𝐜{\bf c} is the center of 𝔓{\mathfrak{P}} with respect to 𝛏{\bm{\xi}}. If Φ(𝐜)(0,0,0)\Phi({\bf c})\neq(0,0,0), then Φ(𝐜)\Phi({\bf c}) is the center of 𝔓{\mathfrak{P}} with respect to Φ(𝛏)\Phi({\bm{\xi}}).

Proof.

Let uu be a local uniformizer of 𝔓{\mathfrak{P}} and =mini=02{ν𝔓(ξi)}\ell=\min_{i=0}^{2}\{\nu_{{\mathfrak{P}}}(\xi_{i})\}. One has that

(π𝔓(uξ0),π𝔓(uξ1),π𝔓(uξ2))=λ𝐜(\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{0}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{1}),\pi_{{\mathfrak{P}}}(u^{-\ell}\xi_{2}))=\lambda{\bf c}

for some nonzero λK\lambda\in K. Denote m=tdeg(ϕi)m={\rm tdeg}(\phi_{i}). Then

π𝔓(Φ(u𝝃))=Φ(π𝔓(u𝝃))=Φ(λ𝐜)=λmΦ(𝐜)(0,0,0).\pi_{{\mathfrak{P}}}(\Phi(u^{-\ell}{\bm{\xi}}))=\Phi(\pi_{\mathfrak{P}}(u^{-\ell}{\bm{\xi}}))=\Phi(\lambda{\bf c})=\lambda^{m}\Phi({\bf c})\neq(0,0,0).

This implies that mini=02{ν𝔓(ϕi(u𝝃))}=0.\min_{i=0}^{2}\{\nu_{{\mathfrak{P}}}(\phi_{i}(u^{-\ell}{\bm{\xi}}))\}=0. In other words,

mini=02{ν𝔓(ϕi(𝝃))}=m.\min_{i=0}^{2}\{\nu_{{\mathfrak{P}}}(\phi_{i}({\bm{\xi}}))\}=m\ell.

Then the center of 𝔓{\mathfrak{P}} with respect to Φ(𝝃)\Phi({\bm{\xi}}) is

π𝔓(umΦ(𝝃))=π𝔓(Φ(u𝝃))=Φ(λ𝐜)=λmΦ(𝐜).\pi_{{\mathfrak{P}}}(u^{-m\ell}\Phi({\bm{\xi}}))=\pi_{{\mathfrak{P}}}(\Phi(u^{-\ell}{\bm{\xi}}))=\Phi(\lambda{\bf c})=\lambda^{m}\Phi({\bf c}).

Hence Φ(𝐜)\Phi({\bf c}) is the center of 𝔓{\mathfrak{P}} with respect to Φ(𝝃)\Phi({\bm{\xi}}). ∎

2.4 Heights

All algebraic function fields under consideration in this subsection are finite extensions of k(t)k(t). They are algebraic function fields of one variable over kk and the places and order functions in them are defined as the same as in the previous subsection.

Definition 2.6.

All points are considered as points in some suitable projective spaces over k(t)¯{\overline{k(t)}}.

  1. 1.

    Given 𝐚=(a0,,am)m(k(t)¯){\bf a}=(a_{0},\dots,a_{m})\in\mathbb{P}^{m}({\overline{k(t)}}), let RR be a finite extension of k(t)k(t) containing all aia_{i}. We define the height of 𝐚{\bf a}, denoted by T(𝐚)T({\bf a}), to be

    𝔭maxi=0m{ν𝔭(ai)}[R:k(t)]\frac{\sum_{{\mathfrak{p}}}\max_{i=0}^{m}\{-\nu_{{\mathfrak{p}}}(a_{i})\}}{[R:k(t)]}

    where 𝔭{\mathfrak{p}} ranges over all places of RR.

  2. 2.

    For A=(ai,j)GL3(k(t)¯)A=(a_{i,j})\in{\rm GL}_{3}({\overline{k(t)}}), we define

    T(A)=T((a1,1,a1,2,a1,3,,a3,3)).T(A)=T((a_{1,1},a_{1,2},a_{1,3},\dots,a_{3,3})).
  3. 3.

    For ak(t)¯a\in{\overline{k(t)}}, we define the height of aa to be T((1,a))T((1,a)), denoted by T(a)T(a).

  4. 4.

    Let FF be a polynomial in k(t)¯[x0,,xm]{\overline{k(t)}}[x_{0},\dots,x_{m}]. Suppose that FF contains at least two terms. We define the height of FF, denoted by T(F)T(F), to be T(𝐜)T({\bf c}) where 𝐜{\bf c} is the point in a suitable projective space formed by the coefficients of FF. For convention, when FF only contains one term, we defined T(F)T(F) to be zero.

  5. 5.

    Let VV be a hypersurface in m(k(t)¯)\mathbb{P}^{m}({\overline{k(t)}}) defined by Fk(t)¯[x0,,xm]F\in{\overline{k(t)}}[x_{0},\dots,x_{m}]. We define the height of VV, denoted by T(V)T(V), to be T(F)T(F).

Remark 2.7.

Assume that 𝐚=(a0,,am),𝐛=(b0,,bm)m(k(t)¯){\bf a}=(a_{0},\dots,a_{m}),{\bf b}=(b_{0},\dots,b_{m})\in\mathbb{P}^{m}({\overline{k(t)}}).

  1. 1.

    One sees that T(𝐚)T({\bf a}) is independent of the choice of homogeneous coordinates and the choice of RR. Without loss of generality, we suppose a0=1a_{0}=1, then

    T(𝐚)=𝔭max{0,ν𝔭(a1),,ν𝔭(am)}[R:k(t)]0.T({\bf a})=\frac{\sum_{{\mathfrak{p}}}\max\{0,-\nu_{\mathfrak{p}}(a_{1}),\dots,\nu_{\mathfrak{p}}(a_{m})\}}{[R:k(t)]}\geq 0.
  2. 2.

    Assume RR is a finite extension of k(t)k(t) containing all aia_{i} and bib_{i}. Then one sees that if maxi{ν𝔭(ai)}maxi{ν𝔭(bi)}\max_{i}\{-\nu_{\mathfrak{p}}(a_{i})\}\geq\max_{i}\{-\nu_{{\mathfrak{p}}}(b_{i})\} for all places 𝔭{\mathfrak{p}} of RR then T(𝐚)T(𝐛)T({\bf a})\geq T({\bf b}).

  3. 3.

    Suppose that a0,a1,,amk[t]a_{0},a_{1},\dots,a_{m}\in k[t] and gcd(a0,,am)=1\gcd(a_{0},\dots,a_{m})=1. Then

    T(𝐚)=max{deg(a0),,deg(am)}.T({\bf a})=\max\{\deg(a_{0}),\dots,\deg(a_{m})\}.

    To see this, let R=k(t)R=k(t). Then

    T(𝐚)=𝔭max{ν𝔭(a0),,ν𝔭(am)}T({\bf a})=\sum_{{\mathfrak{p}}}\max\{-\nu_{{\mathfrak{p}}}(a_{0}),\dots,-\nu_{\mathfrak{p}}(a_{m})\}

    where 𝔭{\mathfrak{p}} ranges over all the places of RR. Note that every place of RR has a local uniformizer of the form 1/t1/t or tct-c for some ckc\in k. Suppose the place 𝔭{\mathfrak{p}} has tct-c as a local uniformizer. Then ν𝔭(ai)>0\nu_{\mathfrak{p}}(a_{i})>0 if and only if (tc)|ai(t-c)|a_{i}. Since gcd(a0,,am)=1\gcd(a_{0},\dots,a_{m})=1, there is some i0i_{0} such that ν𝔭(ai0)=0\nu_{\mathfrak{p}}(a_{i_{0}})=0. This implies that for places 𝔭{\mathfrak{p}} with tc,ckt-c,c\in k as local uniformizers,

    max{ν𝔭(a0),,ν𝔭(am)}=0.\max\{-\nu_{\mathfrak{p}}(a_{0}),\dots,-\nu_{\mathfrak{p}}(a_{m})\}=0.

    For the place with 1/t1/t as local uniformizer, one has that ν𝔭(ai)=deg(ai)\nu_{\mathfrak{p}}(a_{i})=-\deg(a_{i}). So for this place,

    max{ν𝔭(a0),,ν𝔭(am)}=max{deg(a0),,deg(am)}.\max\{-\nu_{\mathfrak{p}}(a_{0}),\dots,-\nu_{\mathfrak{p}}(a_{m})\}=\max\{\deg(a_{0}),\dots,\deg(a_{m})\}.

    Consequently, T(𝐚)=max{deg(a0),,deg(am)}T({\bf a})=\max\{\deg(a_{0}),\dots,\deg(a_{m})\}.

  4. 4.

    Let ak(t)¯a\in{\overline{k(t)}} and R=k(t,a)R=k(t,a). Let g(t,x)g(t,x) be a nonzero irreducible polynomial over kk such that g(t,a)=0g(t,a)=0. It is clear that T(a)=0T(a)=0 if aka\in k. Assume that aka\notin k and 𝔭1,,𝔭s{\mathfrak{p}}_{1},\dots,{\mathfrak{p}}_{s} are all distinct poles of aa in RR, then

    T(a)=i=1sν𝔭i(a)[R:k(t)]=[R:k(a)][R:k(t)]=deg(g,t)deg(g,x).T(a)=\frac{-\sum_{i=1}^{s}\nu_{{\mathfrak{p}}_{i}}(a)}{[R:k(t)]}=\frac{[R:k(a)]}{[R:k(t)]}=\frac{\deg(g,t)}{\deg(g,x)}.

    In particular, if ak(t)a\in k(t) then T(a)=deg(a)T(a)=\deg(a) which is defined to be the maximun of the degrees of the denominator and numerator of aa.

From the above remark, it is easy to see that for ak(t)¯{0}a\in{\overline{k(t)}}\setminus\{0\} and ii\in\mathbb{Z}

T(ai)=|i|T(a).T(a^{i})=|i|T(a).
Proposition 2.8.

Let a,bk(t)¯a,b\in{\overline{k(t)}}, c1,,c4kc_{1},\dots,c_{4}\in k with c1c4c2c30c_{1}c_{4}-c_{2}c_{3}\neq 0. Then

  1. 1.

    T(c1a+c2c3a+c4)=T(a)T\left(\frac{c_{1}a+c_{2}}{c_{3}a+c_{4}}\right)=T(a) if c3a+c40c_{3}a+c_{4}\neq 0;

  2. 2.

    T(ab)T(a)+T(b)T(ab)\leq T(a)+T(b);

  3. 3.

    T(a+λb)T(a)+T(b)T(a+\lambda b)\leq T(a)+T(b) for all λk\lambda\in k.

Proof.

1.1. If aka\in k then the assertion is obvious. Suppose that aka\notin k. Let R=k(t,a)R=k(t,a). Then R=k(t,(c1a+c2)/(c3a+c4))R=k(t,(c_{1}a+c_{2})/(c_{3}a+c_{4})). The assertion follows from Remark 2.7 and the fact that

[R:k(c1a+c2c3a+c4)]=[R:k(a)].\left[R:k\left(\frac{c_{1}a+c_{2}}{c_{3}a+c_{4}}\right)\right]=[R:k(a)].

2.2. Let R=k(t,a,b)R=k(t,a,b). For each place 𝔭{\mathfrak{p}} of RR, ν𝔭(ab)=ν𝔭(a)ν𝔭(b)-\nu_{{\mathfrak{p}}}(ab)=-\nu_{{\mathfrak{p}}}(a)-\nu_{{\mathfrak{p}}}(b) and thus

max{0,ν𝔭(ab)}max{0,ν𝔭(a)}+max{0,ν𝔭(b)}.\max\{0,-\nu_{{\mathfrak{p}}}(ab)\}\leq\max\{0,-\nu_{{\mathfrak{p}}}(a)\}+\max\{0,-\nu_{{\mathfrak{p}}}(b)\}.

The assertion then follows from Remark 2.7.

3.3. Use an argument similar to that in 2. and the fact that

ν𝔭(a+λb)max{ν𝔭(a),ν𝔭(b)}.-\nu_{{\mathfrak{p}}}(a+\lambda b)\leq\max\{-\nu_{{\mathfrak{p}}}(a),-\nu_{{\mathfrak{p}}}(b)\}.

The following examples show that the equalities may hold in 22 and 33 of Proposition 2.8.

Example 2.9.

Let a=t2,b=t3+1a=t^{2},b=t^{3}+1 and λk{0}\lambda\in k\setminus\{0\}. Then

T(ab)=5=T(a)+T(b),T(1/a+λ/b)=5=T(1/a)+T(1/b.)T(ab)=5=T(a)+T(b),\,\,T(1/a+\lambda/b)=5=T(1/a)+T(1/b.)

Moreover both of them are greater than the maximun of T(a),T(b)T(a),T(b).

In the following, 𝐲{\bf y} stands for the vector with indeterminates y1,,ysy_{1},\dots,y_{s} and 𝐲𝐝{\bf y}^{\bf d} denotes i=1syidi\prod_{i=1}^{s}y_{i}^{d_{i}} for 𝐝=(d1,,ds)s{\bf d}=(d_{1},\dots,d_{s})\in\mathbb{Z}^{s}.

Proposition 2.10.

Let ff and gg be polynomials in k(t)¯[𝐲]{\overline{k(t)}}[{\bf y}], then

  1. 1.

    T(fg)T(f)+T(g);T(fg)\leq T(f)+T(g);

  2. 2.

    If both ff and gg have 1 as a coefficient, then T(f+g)T(f)+T(g).T(f+g)\leq T(f)+T(g).

Proof.

Write f=𝐝a𝐝𝐲𝐝f=\sum_{{\bf d}}a_{{\bf d}}{\bf y}^{\bf d} and g=𝐝b𝐝𝐲𝐝g=\sum_{{\bf d}}b_{\bf d}{\bf y}^{\bf d} with a𝐝,b𝐝k(t)¯a_{\bf d},b_{\bf d}\in{\overline{k(t)}}. Let RR be a finite extension of k(t)k(t) containing all a𝐝,b𝐝a_{\bf d},b_{\bf d}, and 𝔭{\mathfrak{p}} a place of RR.

1. Each coefficient of fgfg is of the form i=1sa𝐝ib𝐝i\sum_{i=1}^{s}a_{{\bf d}_{i}}b_{{\bf d}_{i}}, where s1s\geq 1. Since

νp(i=1sa𝐝ib𝐝i)maxi=1s{νp(a𝐝i)νp(b𝐝i)},-\nu_{p}\left(\sum_{i=1}^{s}a_{{\bf d}_{i}}b_{{\bf d}_{i}}\right)\leq\max_{i=1}^{s}\{-\nu_{p}(a_{{\bf d}_{i}})-\nu_{p}(b_{{\bf d}_{i}})\},

we have

maxcC{ν𝔭(c)}max𝐝{ν𝔭(a𝐝)}+max𝐝{ν𝔭(b𝐝)}\max_{\mbox{$c\in C$}}\{-\nu_{{\mathfrak{p}}}(c)\}\leq\max_{{\bf d}}\{-\nu_{{\mathfrak{p}}}(a_{\bf d})\}+\max_{{\bf d}}\{-\nu_{{\mathfrak{p}}}(b_{\bf d})\}

where CC is the set of all coefficients of fgfg. It follows that T(fg)T(f)+T(g).T(fg)\leq T(f)+T(g).

2. The assertion follows from the fact that

ν𝔭(a𝐝+b𝐝)\displaystyle-\nu_{\mathfrak{p}}(a_{\bf d}+b_{\bf d}) max𝐝{ν𝔭(a𝐝),ν𝔭(b𝐝)}max𝐝{0,ν𝔭(a𝐝),ν𝔭(b𝐝)}\displaystyle\leq\max_{\bf d}\{-\nu_{\mathfrak{p}}(a_{\bf d}),-\nu_{\mathfrak{p}}(b_{\bf d})\}\leq\max_{{\bf d}}\{0,-\nu_{\mathfrak{p}}(a_{\bf d}),-\nu_{\mathfrak{p}}(b_{\bf d})\}
max𝐝{0,ν𝔭(a𝐝)}+max𝐝{0,ν𝔭(b𝐝)}.\displaystyle\leq\max_{{\bf d}}\{0,-\nu_{\mathfrak{p}}(a_{\bf d})\}+\max_{{\bf d}}\{0,-\nu_{\mathfrak{p}}(b_{\bf d})\}.

Proposition 2.11.

Let f=xn+an1xn1++a0f=x^{n}+a_{n-1}x^{n-1}+\dots+a_{0} where n>0n>0 and aik(t)¯a_{i}\in{\overline{k(t)}}. Suppose that α\alpha is a zero of ff in k(t)¯{\overline{k(t)}} and RR is a finite extension of k(t)k(t) containing α\alpha and all aia_{i}. Then for each place 𝔭{\mathfrak{p}} of RR,

max{0,ν𝔭(α)}max{0,ν𝔭(a0),,ν𝔭(an1)}.\max\{0,-\nu_{{\mathfrak{p}}}(\alpha)\}\leq\max\{0,-\nu_{\mathfrak{p}}(a_{0}),\dots,-\nu_{\mathfrak{p}}(a_{n-1})\}.
Proof.

The assertion is clear if αk\alpha\in k or 𝔭{\mathfrak{p}} is not a pole of α\alpha. Assume αk(t)¯k\alpha\in{\overline{k(t)}}\setminus k and 𝔭{\mathfrak{p}} is a pole of α\alpha. Then

ν𝔭(αn)=νp(i=0n1aiαi)mini=0n1{iν𝔭(α)+ν𝔭(ai)}=i0ν𝔭(α)+ν𝔭(ai0)\nu_{{\mathfrak{p}}}(\alpha^{n})=\nu_{p}\left(\sum_{i=0}^{n-1}a_{i}\alpha^{i}\right)\geq\min_{i=0}^{n-1}\left\{i\nu_{{\mathfrak{p}}}(\alpha)+\nu_{{\mathfrak{p}}}(a_{i})\right\}=i_{0}\nu_{\mathfrak{p}}(\alpha)+\nu_{\mathfrak{p}}(a_{i_{0}})

for some i0i_{0} with 0i0n10\leq i_{0}\leq n-1. This together with the fact that ν𝔭(α)<0\nu_{\mathfrak{p}}(\alpha)<0 implies that

ν𝔭(α)ν𝔭(ai0)ni0ν𝔭(ai0)mini=0n1{ν𝔭(ai)},\nu_{{\mathfrak{p}}}(\alpha)\geq\frac{\nu_{{\mathfrak{p}}}(a_{i_{0}})}{n-i_{0}}\geq\nu_{{\mathfrak{p}}}(a_{i_{0}})\geq\min_{i=0}^{n-1}\left\{\nu_{\mathfrak{p}}(a_{i})\right\},

i.e.

ν𝔭(α)maxi=0n1{ν𝔭(ai)}.-\nu_{{\mathfrak{p}}}(\alpha)\leq\max_{i=0}^{n-1}\left\{-\nu_{{\mathfrak{p}}}(a_{i})\right\}.

Consequently, one has that

max{0,ν𝔭(α)}max{0,ν𝔭(a0),,ν𝔭(an1)}.\max\{0,-\nu_{{\mathfrak{p}}}(\alpha)\}\leq\max\{0,-\nu_{{\mathfrak{p}}}(a_{0}),\dots,-\nu_{{\mathfrak{p}}}(a_{n-1})\}.

Corollary 2.12.

Let ff be a polynomial in k(t)¯[x]{\overline{k(t)}}[x] and α\alpha a zero of ff in k(t)¯{\overline{k(t)}}. Then T(α)T(f)T(\alpha)\leq T(f).

The equality in Corollary 2.12 may hold as shown in the following example.

Example 2.13.

Let

f=x2(t3+1)x+t3=(xt3)(x1),f=x^{2}-(t^{3}+1)x+t^{3}=(x-t^{3})(x-1),

then T(t3)=T(f)=3T(t^{3})=T(f)=3.

Lemma 2.14.

Assume f,gk(t)¯[x]f,g\in{\overline{k(t)}}[x] and gg is a factor of ff. Then T(g)T(f).T(g)\leq T(f).

Proof.

Without loss of generality, we may assume both ff and gg are monic. Write

f=xn+i=0n1aixi,aik(t)¯.f=x^{n}+\sum_{i=0}^{n-1}a_{i}x^{i},\,\,a_{i}\in{\overline{k(t)}}.

We first show that if f=ghf=gh and degg=1\deg g=1 then T(h)T(f)T(h)\leq T(f). Suppose that

h=xn1+j=0n2bjxj,g=x+αh=x^{n-1}+\sum_{j=0}^{n-2}b_{j}x^{j},\,\,g=x+\alpha

where α,bjk(t)¯\alpha,b_{j}\in{\overline{k(t)}}. Let R=k(t,α,b0,,bn2)R=k(t,\alpha,b_{0},\dots,b_{n-2}). For each place 𝔭{\mathfrak{p}} of RR, denote

N𝔭=max{0,ν𝔭(a0),,νp(an1)}.N_{\mathfrak{p}}=\max\{0,-\nu_{\mathfrak{p}}(a_{0}),\dots,-\nu_{p}(a_{n-1})\}.

Then max{0,ν𝔭(ai)}N𝔭\max\{0,-\nu_{\mathfrak{p}}(a_{i})\}\leq N_{\mathfrak{p}} for all i=0,,n1i=0,\dots,n-1 and all places 𝔭{\mathfrak{p}} of RR. From f=ghf=gh, one has that

bn2+α=an1,b0α=a0,biα+bi1=ai,i=1,,n2.b_{n-2}+\alpha=a_{n-1},\,\,b_{0}\alpha=a_{0},\,\,b_{i}\alpha+b_{i-1}=a_{i},\,\,i=1,\dots,n-2.

We claim that max{0,ν𝔭(bj)}N𝔭\max\{0,-\nu_{\mathfrak{p}}(b_{j})\}\leq N_{\mathfrak{p}} for all j=0,,n2j=0,\dots,n-2 and all places 𝔭{\mathfrak{p}} of RR. For j=n2j=n-2, one has that

ν𝔭(bn2)=ν𝔭(an1α)min{ν𝔭(an1),ν𝔭(α)}.\nu_{\mathfrak{p}}(b_{n-2})=\nu_{\mathfrak{p}}(a_{n-1}-\alpha)\geq\min\{\nu_{\mathfrak{p}}(a_{n-1}),\nu_{\mathfrak{p}}(\alpha)\}.

This implies that max{0,ν𝔭(bn2)}max{0,ν𝔭(an1),ν𝔭(α)}\max\{0,-\nu_{\mathfrak{p}}(b_{n-2})\}\leq\max\{0,-\nu_{\mathfrak{p}}(a_{n-1}),-\nu_{\mathfrak{p}}(\alpha)\}. By Proposition 2.11, we have that max{0,ν𝔭(α)}N𝔭\max\{0,-\nu_{\mathfrak{p}}(\alpha)\}\leq N_{\mathfrak{p}}. Hence max{0,ν𝔭(bn2)}N𝔭\max\{0,-\nu_{\mathfrak{p}}(b_{n-2})\}\leq N_{\mathfrak{p}} for all places 𝔭{\mathfrak{p}} of RR. Now assume that there is a place 𝔮{\mathfrak{q}} of RR and j0j_{0} with 0j0<n20\leq j_{0}<n-2 such that max{0,ν𝔮(bj0)}>N𝔮\max\{0,-\nu_{\mathfrak{q}}(b_{j_{0}})\}>N_{\mathfrak{q}} but max{0,ν𝔮(bj0+1)}N𝔮\max\{0,-\nu_{\mathfrak{q}}(b_{j_{0}+1})\}\leq N_{\mathfrak{q}}. Then one has that ν𝔮(bj0)<0\nu_{\mathfrak{q}}(b_{j_{0}})<0 and ν𝔮(bj0)<ν𝔮(ai)\nu_{\mathfrak{q}}(b_{j_{0}})<\nu_{\mathfrak{q}}(a_{i}) for all i=0,,n1i=0,\dots,n-1. On the other hand, since max{0,ν𝔮(bj0+1)}N𝔮\max\{0,-\nu_{\mathfrak{q}}(b_{j_{0}+1})\}\leq N_{\mathfrak{q}}, ν𝔮(bj0)>N𝔮ν𝔮(bj0+1)-\nu_{\mathfrak{q}}(b_{j_{0}})>N_{\mathfrak{q}}\geq-\nu_{\mathfrak{q}}(b_{j_{0}+1}) i.e. ν𝔮(bj0+1)<ν𝔮(bj0)<0\nu_{\mathfrak{q}}(b_{j_{0}+1})<\nu_{{\mathfrak{q}}}(b_{j_{0}})<0. From αbj0+1=aj0+1bj0\alpha b_{j_{0}+1}=a_{j_{0}+1}-b_{j_{0}}, one has that

ν𝔮(αbj0+1)=ν𝔮(α)+ν𝔮(bj0+1)=min{ν𝔮(aj0+1),ν𝔮(bj0)}=ν𝔮(bj0).\nu_{\mathfrak{q}}(\alpha b_{j_{0}+1})=\nu_{\mathfrak{q}}(\alpha)+\nu_{\mathfrak{q}}(b_{j_{0}+1})=\min\{\nu_{\mathfrak{q}}(a_{j_{0}+1}),\nu_{\mathfrak{q}}(b_{j_{0}})\}=\nu_{{\mathfrak{q}}}(b_{j_{0}}).

The last equality holds because ν𝔮(bj0)<ν𝔮(ai)\nu_{{\mathfrak{q}}}(b_{j_{0}})<\nu_{\mathfrak{q}}(a_{i}) for all ii. Thus ν𝔮(α)<0\nu_{\mathfrak{q}}(\alpha)<0 which implies that

ν𝔮(αbj0)=ν𝔮(α)+ν𝔮(bj0)<ν𝔮(ai)\nu_{\mathfrak{q}}(\alpha b_{j_{0}})=\nu_{\mathfrak{q}}(\alpha)+\nu_{\mathfrak{q}}(b_{j_{0}})<\nu_{\mathfrak{q}}(a_{i})

for all i=0,,n1i=0,\dots,n-1. As bj01=aj0αbj0b_{j_{0}-1}=a_{j_{0}}-\alpha b_{j_{0}}, one has that νq(bj01)=νq(α)+νq(bj0)<0\nu_{q}(b_{j_{0}-1})=\nu_{q}(\alpha)+\nu_{q}(b_{j_{0}})<0 and furthermore νq(bj01)<νq(ai)\nu_{q}(b_{j_{0}-1})<\nu_{q}(a_{i}) for all i=0,,n1i=0,\dots,n-1. In other words, max{0,νq(bj01)}>Nq\max\{0,-\nu_{q}(b_{j_{0}-1})\}>N_{q}. Applying a similar argument to the equalities bj=aj+1αbj+1b_{j}=a_{j+1}-\alpha b_{j+1} for j=j02,,0j=j_{0}-2,\dots,0 successively yields that ν𝔮(bj)<0\nu_{\mathfrak{q}}(b_{j})<0 and max{0,ν𝔮(bj)}>N𝔮\max\{0,-\nu_{\mathfrak{q}}(b_{j})\}>N_{\mathfrak{q}} for all j=j02,,0j=j_{0}-2,\dots,0. However, one has that αb0=a0\alpha b_{0}=a_{0}. This implies that ν𝔮(a0)<ν𝔮(b0)<0\nu_{\mathfrak{q}}(a_{0})<\nu_{\mathfrak{q}}(b_{0})<0 and thus N𝔮ν𝔮(a0)>ν𝔮(b0)>0N_{\mathfrak{q}}\geq-\nu_{\mathfrak{q}}(a_{0})>-\nu_{\mathfrak{q}}(b_{0})>0. That is to say, N𝔮max{0,ν𝔮(b0)}N_{\mathfrak{q}}\geq\max\{0,-\nu_{\mathfrak{q}}(b_{0})\}, a contradiction. This proves the claim. This claim and Remark 2.7 imply that T(h)T(f)T(h)\leq T(f).

Now we prove the assertion by induction on deg(f)\deg(f). The base case deg(f)=1\deg(f)=1 is obvious. Suppose that the assertion holds for deg(f)n\deg(f)\leq n. Consider the case deg(f)=n+1\deg(f)=n+1. If f=gf=g then there is nothing to prove. Suppose that fgf\neq g. Then there is βk(t)¯\beta\in{\overline{k(t)}} such that (x+β)g(x+\beta)g divides ff. Let f=(x+β)f~f=(x+\beta)\tilde{f} for some f~k(t)¯[x]\tilde{f}\in{\overline{k(t)}}[x]. Then T(f~)T(f)T(\tilde{f})\leq T(f) and gg divides f~\tilde{f}. By induction hypothesis, one has that T(g)T(f~)T(f)T(g)\leq T(\tilde{f})\leq T(f). ∎

Corollary 2.15.

Assume f(x,y),g(x,y)k(t)¯[x,y]f(x,y),g(x,y)\in{\overline{k(t)}}[x,y] and g(x,y)g(x,y) is a factor of f(x,y)f(x,y). Then T(g(x,y))T(f(x,y)).T(g(x,y))\leq T(f(x,y)).

Proof.

Suppose that f(x,y)=i,jci,jxiyjf(x,y)=\sum_{i,j}c_{i,j}x^{i}y^{j} where ci,jk(t)¯c_{i,j}\in{\overline{k(t)}}. Let dd be an integer greater than tdeg(f(x,y)){\rm tdeg}(f(x,y)). One has that

f(x,xd)=i,jci,jxi+jd.f(x,x^{d})=\sum_{i,j}c_{i,j}x^{i+jd}.

Note that for 0i,j,l,m<d0\leq i,j,l,m<d, i+jd=l+mdi+jd=l+md if and only if (i,j)=(l,m)(i,j)=(l,m). This implies that the set of the coefficients of f(x,y)f(x,y) coincides with that of f(x,xd)f(x,x^{d}). Hence T(f(x,y))=T(f(x,xd))T(f(x,y))=T(f(x,x^{d})). Similarly, T(g(x,y))=T(g(x,xd))T(g(x,y))=T(g(x,x^{d})). It is clear that g(x,xd)g(x,x^{d}) is still a factor of f(x,xd)f(x,x^{d}). By Lemma 2.14, T(g(x,xd))T(f(x,xd))T(g(x,x^{d}))\leq T(f(x,x^{d})). Thus T(g(x,y))T(f(x,y))T(g(x,y))\leq T(f(x,y)). ∎

Proposition 2.16.

Assume f,gk(t)¯[𝐲,z]f,g\in{\overline{k(t)}}[{\bf y},z], then

T(resz(f,g))deg(g,z)T(f)+deg(f,z)T(g)T({\rm res}_{z}(f,g))\leq\deg(g,z)T(f)+\deg(f,z)T(g)

where resz(f,g){\rm res}_{z}(f,g) is the resultant of ff and gg with respect to zz.

Proof.

The assertion is clear that if resz(f,g)=0{\rm res}_{z}(f,g)=0. Consider the case resz(f,g)0{\rm res}_{z}(f,g)\neq 0. Assume deg(f,z)=n,deg(g,z)=m\deg(f,z)=n,\deg(g,z)=m. Write

f=i=0nai(𝐲)zi,g=i=0mbi(𝐲)zif=\sum_{i=0}^{n}a_{i}({\bf y})z^{i},\quad g=\sum_{i=0}^{m}b_{i}({\bf y})z^{i}

where ai(𝐲),bi(𝐲)k(t)¯[𝐲]a_{i}({\bf y}),b_{i}({\bf y})\in{\overline{k(t)}}[{\bf y}]. Denote further by C1,C2C_{1},C_{2} the sets of the coefficients in 𝐲,z{\bf y},z of f,gf,g respectively. Then

resz(f,g)=|anan1a0anan1a0bmbm1b0bmbm1b0|.{\rm res}_{z}(f,g)=\begin{vmatrix}a_{n}&a_{n-1}&\cdots&a_{0}\\ &\ddots&\ddots&&\ddots\\ &&a_{n}&a_{n-1}&\cdots&a_{0}\\ b_{m}&b_{m-1}&\cdots&b_{0}\\ &\ddots&\ddots&&\ddots\\ &&b_{m}&b_{m-1}&\cdots&b_{0}\\ \end{vmatrix}.

By the definitions of determinant, we can write

resz(f,g)=𝐝(j=1𝐝β𝐝,j𝐦𝐝,j𝐧𝐝,j)𝐲𝐝{\rm res}_{z}(f,g)=\sum_{{\bf d}}\left(\sum_{j=1}^{\ell_{\bf d}}\beta_{{\bf d},j}{\bf m}_{{\bf d},j}{\bf n}_{{\bf d},j}\right){\bf y}^{{\bf d}}

where β𝐝,j,𝐝,𝐝0\beta_{{\bf d},j},\ell_{\bf d}\in\mathbb{Z},\ell_{\bf d}\geq 0, 𝐦𝐝,j{\bf m}_{{\bf d},j} is a monomial in C1C_{1} with total degree mm and 𝐧𝐝,j{\bf n}_{{\bf d},j} is a monomial in C2C_{2} with total degree nn. Let R=k(t,C1,C2)R=k(t,C_{1},C_{2}). For each place 𝔭{\mathfrak{p}} of RR, we have

ν𝔭(jβ𝐝,j𝐦𝐝,j𝐧𝐝,j)\displaystyle-\nu_{{\mathfrak{p}}}\left(\sum_{j}\beta_{{\bf d},j}{\bf m}_{{\bf d},j}{\bf n}_{{\bf d},j}\right) maxj{ν𝔭(𝐦𝐝,j𝐧𝐝,j)}\displaystyle\leq\max_{j}\{-\nu_{{\mathfrak{p}}}({\bf m}_{{\bf d},j}{\bf n}_{{\bf d},j})\}
mmaxcC1{ν𝔭(c)}+nmaxcC2{ν𝔭(c)}.\displaystyle\leq m\max_{c\in C_{1}}\{-\nu_{{\mathfrak{p}}}(c)\}+n\max_{c\in C_{2}}\{-\nu_{{\mathfrak{p}}}(c)\}.

Therefore by Remark 2.7,

T(resz(f,g))mT(f)+nT(g).T({\rm res}_{z}(f,g))\leq mT(f)+nT(g).

3 Degrees and Heights on Riemann-Roth spaces

Throughout this section, {\mathcal{R}} denotes an algebraic function field of one variable over k(t)¯{\overline{k(t)}}. Let 𝝃=(ξ0,ξ1,ξ2)3{\bm{\xi}}=(\xi_{0},\xi_{1},\xi_{2})\in{\mathcal{R}}^{3} be such that 𝒞(𝝃){\mathcal{C}}({\bm{\xi}}) a plane projective model of {\mathcal{R}}, i.e. =k(t)¯(ξ0/ξ2,ξ1/ξ2){\mathcal{R}}={\overline{k(t)}}(\xi_{0}/\xi_{2},\xi_{1}/\xi_{2}). Each hh\in{\mathcal{R}} can be presented by G(𝝃)/H(𝝃)G({\bm{\xi}})/H({\bm{\xi}}) where G,HG,H are two homogeneous polynomials in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] of the same degree and having no common factors, and H(𝝃)0H({\bm{\xi}})\neq 0. We call (G,H)(G,H) a representation of hh. This section shall focus on determining the degrees and heights of representations of elements in Riemann-Roch spaces. There are several algorithms for computing the bases of Riemann-Roth spaces (see for example [10, 11]). However no existing algorithm provided explicit bounds for the degrees and heights of GG and HH where (G,H)(G,H) represents an element in these bases. These bounds play an essential role in estimating the heights of points on a plane algebraic curve. In this section, we shall follow the algorithm developed in [11] to obtain these bounds. For this purpose, we need to resolve singularities of a given plane algebraic curve to obtain the one with only ordinary singularities. This can be done by a sequence of quadratic transformations.

In this section, unless otherwise stated, FF always stands for an irreducible homogeneous polynomial in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] of degree n>0n>0 which defines a plane projective model of {\mathcal{R}}, i.e. there is 𝝃=(ξ0,ξ1,ξ2)3{\bm{\xi}}=(\xi_{0},\xi_{1},\xi_{2})\in{\mathcal{R}}^{3} such that =k(t)¯(ξ0/ξ2,ξ1/ξ2){\mathcal{R}}={\overline{k(t)}}(\xi_{0}/\xi_{2},\xi_{1}/\xi_{2}) and F(𝝃)=0F({\bm{\xi}})=0.

3.1 Quadratic transformations

Let DD be a divisor in {\mathcal{R}}. Due to Proposition on page of [9], there is a birational transformation {\mathcal{B}} such that the transformation of F=0F=0 under {\mathcal{B}} is a plane projective curve with only ordinary singularties, moreover {\mathcal{B}} can be chosen to be the composition of a sequence of suitable quadratic transformations. In this subsection, we shall investigate the degree and height of the transformation of F=0F=0 under a quadratic transformation.

Definition 3.1.
  1. 1.

    {\mathcal{L}} stands for a projective change of coordinates on 2(k(t)¯)\mathbb{P}^{2}({\overline{k(t)}}) that is defined as (𝐜)=𝐜M{\mathcal{L}}({\bf c})={\bf c}M_{\mathcal{L}} where MGL3(k(t)¯)M_{\mathcal{L}}\in{\rm GL}_{3}({\overline{k(t)}}), and 𝒬{\mathcal{Q}} denotes the standard quadratic transformation that is defined as

    𝒬(𝐜)=(c1c2,c0c2,c0c1){\mathcal{Q}}({\bf c})=(c_{1}c_{2},c_{0}c_{2},c_{0}c_{1})

    where 𝐜=(c0,c1,c2){\bf c}=(c_{0},c_{1},c_{2}).

  2. 2.

    The height of {\mathcal{L}}, denoted by T()T({\mathcal{L}}), is defined as T(M)T(M_{\mathcal{L}}).

Notation 3.2.
  1. 1.

    FF^{\mathcal{L}} stands for F((x0,x1,x2)M)F((x_{0},x_{1},x_{2})M_{\mathcal{L}}).

  2. 2.

    F𝒬F^{\mathcal{Q}} stands for the irreducible polynomial F~\tilde{F} satisfying

    F(x1x2,x0x2,x0x1)=x0d0x1d1x2d2F~F(x_{1}x_{2},x_{0}x_{2},x_{0}x_{1})=x_{0}^{d_{0}}x_{1}^{d_{1}}x_{2}^{d_{2}}\tilde{F}

    where di0d_{i}\geq 0.

One sees that 𝕍(F)\mathbb{V}(F^{\mathcal{L}}) (resp. 𝕍(F𝒬)\mathbb{V}(F^{\mathcal{Q}})) is the variaty 1(𝕍(F)){\mathcal{L}}^{-1}(\mathbb{V}(F)) (resp. the projective closure of 𝒬1(𝕍(F)𝕍(x0x1x2){\mathcal{Q}}^{-1}(\mathbb{V}(F)\setminus\mathbb{V}(x_{0}x_{1}x_{2})).

Remark 3.3.

𝒬{\mathcal{Q}} is bijective on 2(k(t)¯)𝕍(x0x1x2)\mathbb{P}^{2}({\overline{k(t)}})\setminus\mathbb{V}(x_{0}x_{1}x_{2}) and 𝒬1=𝒬{\mathcal{Q}}^{-1}={\mathcal{Q}}.

Let us first bound the heights of the common points of two algebraic curves in 2(k(t)¯)\mathbb{P}^{2}({\overline{k(t)}}).

Proposition 3.4.

Let F,GF,G be two homogenenous polynomials in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] of degree n,mn,m respectively. Suppose that FF and GG have no common factor, and 𝐜2(k(t)¯){\bf c}\in\mathbb{P}^{2}({\overline{k(t)}}) is a common point of F=0F=0 and G=0G=0. Then

T(𝐚)2(mT(F)+nT(G)).T({\bf a})\leq 2(mT(F)+nT(G)).

Furthermore, if G=c0x1+c1x1+c2x2G=c_{0}x_{1}+c_{1}x_{1}+c_{2}x_{2} with cikc_{i}\in k then T(𝐚)T(F)T({\bf a})\leq T(F).

Proof.

Let Hi(xj,xl)=resxi(F,G)H_{i}(x_{j},x_{l})={\rm res}_{x_{i}}(F,G) where {i,j,l}={0,1,2}\{i,j,l\}=\{0,1,2\}. Proposition 2.16 implies that T(Hi)mT(F)+nT(G).T(H_{i})\leq mT(F)+nT(G). Without loss of generality, suppose 𝐚=(1,a1,a2){\bf a}=(1,a_{1},a_{2}). Since 𝐚{\bf a} is a common point of F=0F=0 and G=0G=0,

H2(1,a1)=H1(1,a2)=0.H_{2}(1,a_{1})=H_{1}(1,a_{2})=0.

It follows from Corollary 2.12 that T(ai)mT(F)+nT(G)T(a_{i})\leq mT(F)+nT(G) for all i=1,2i=1,2. Let R=k(t,a1,a2)R=k(t,a_{1},a_{2}) and 𝔭{\mathfrak{p}} a place of RR. Then

max{0,ν𝔭(a1),ν𝔭(a2)}max{0,ν𝔭(a1)}+max{0,ν𝔭(a2)}.\max\{0,-\nu_{{\mathfrak{p}}}(a_{1}),-\nu_{{\mathfrak{p}}}(a_{2})\}\leq\max\{0,-\nu_{{\mathfrak{p}}}(a_{1})\}+\max\{0,-\nu_{{\mathfrak{p}}}(a_{2})\}.

Whence

T(𝐚)T(a1)+T(a2)2(mT(F)+nT(G)).T({\bf a})\leq T(a_{1})+T(a_{2})\leq 2(mT(F)+nT(G)).

It remains to pove the second assertion. Since a0=10a_{0}=1\neq 0, not all c1,c2c_{1},c_{2} are zero. Without loss of generality, assume that c10c_{1}\neq 0. Substituting a1=(c0+a2c2)/c1a_{1}=-(c_{0}+a_{2}c_{2})/c_{1} into F=0F=0 yields that

F(1,(c0+a2c2)/c1,a2)=0.F(1,-(c_{0}+a_{2}c_{2})/c_{1},a_{2})=0.

This implies that T(a2)T(F)T(a_{2})\leq T(F). On the other hand, one sees that

ν𝔭(a1)=ν𝔭((c0+a2c2)/c1)min{ν𝔭(c0),ν𝔭(a2)}.\nu_{\mathfrak{p}}(a_{1})=\nu_{\mathfrak{p}}(-(c_{0}+a_{2}c_{2})/c_{1})\geq\min\{\nu_{\mathfrak{p}}(c_{0}),\nu_{\mathfrak{p}}(a_{2})\}.

So

max{0,ν𝔭(a1),ν𝔭(a2)}max{0,ν𝔭(a2)}.\max\{0,-\nu_{\mathfrak{p}}(a_{1}),-\nu_{\mathfrak{p}}(a_{2})\}\leq\max\{0,-\nu_{\mathfrak{p}}(a_{2})\}.

which results in T(𝐚)T(a2)T(F)T({\bf a})\leq T(a_{2})\leq T(F). ∎

Corollary 3.5.

If 𝐚2(k(t)¯){\bf a}\in\mathbb{P}^{2}({\overline{k(t)}}) is a singular point of F=0F=0, then T(𝐚)2(2n1)T(F).T({\bf a})\leq 2(2n-1)T(F).

Lemma 3.6.

Suppose that {\mathcal{L}} is a projective change of coordinates. Then

  1. 1.

    T(F)T(F)+deg(F)T()T(F^{\mathcal{L}})\leq T(F)+\deg(F)T({\mathcal{L}});

  2. 2.

    for each 𝐜2(k(t)¯){\bf c}\in\mathbb{P}^{2}({\overline{k(t)}}), T((𝐜))T(𝐜)+T()T({\mathcal{L}}({\bf c}))\leq T({\bf c})+T({\mathcal{L}}).

Proof.

Suppose that M=(ai,j)M_{\mathcal{L}}=(a_{i,j}) and

F=i=0nj=0nici,jx0ix1jx2nijF=\sum_{i=0}^{n}\sum_{j=0}^{n-i}c_{i,j}x_{0}^{i}x_{1}^{j}x_{2}^{n-i-j}

where n=deg(F)n=\deg(F) and ci,jk(t)¯c_{i,j}\in{\overline{k(t)}}.

1. One has that

F=i=0nj=0nici,j(l=13al,1xl1)i(l=13al,2xl1)j(l=13al,3xl1)nij.F^{\mathcal{L}}=\sum_{i=0}^{n}\sum_{j=0}^{n-i}c_{i,j}\left(\sum_{l=1}^{3}a_{l,1}x_{l-1}\right)^{i}\left(\sum_{l=1}^{3}a_{l,2}x_{l-1}\right)^{j}\left(\sum_{l=1}^{3}a_{l,3}x_{l-1}\right)^{n-i-j}.

Let ρ\rho be a coefficient of FF^{\mathcal{L}} viewed as polynomial in x0,x1,x2x_{0},x_{1},x_{2}. Then ρ\rho is a kk-linear combination of monomials ci,ja1,1e1,1a3,3e3,3c_{i,j}a_{1,1}^{e_{1,1}}\dots a_{3,3}^{e_{3,3}} with ei,j=n\sum e_{i,j}=n. Let RR be a finite extension of k(t)k(t) containing all ci,jc_{i,j} and ai,ja_{i,j}. Suppose that 𝔭{\mathfrak{p}} is a place of RR. Then one has that

ν𝔭(ρ)\displaystyle\nu_{{\mathfrak{p}}}(\rho) mini,j,i,j{ν𝔭(ci,j)+i,jei,jν𝔭(ai,j)}\displaystyle\geq\min_{i,j,i^{\prime},j^{\prime}}\left\{\nu_{{\mathfrak{p}}}(c_{i,j})+\sum_{i^{\prime},j^{\prime}}e_{i^{\prime},j^{\prime}}\nu_{\mathfrak{p}}(a_{i^{\prime},j^{\prime}})\right\}
mini,j{ν𝔭(ci,j)}+mini,j{i,jei,jν𝔭(ai,j)},\displaystyle\geq\min_{i,j}\{\nu_{{\mathfrak{p}}}(c_{i,j})\}+\min_{i,j}\left\{\sum_{i,j}e_{i,j}\nu_{\mathfrak{p}}(a_{i,j})\right\},

i.e.

ν𝔭(ρ)\displaystyle-\nu_{{\mathfrak{p}}}(\rho) maxi,j{ν𝔭(ci,j)}+maxi,j{i,jei,jν𝔭(ai,j)}\displaystyle\leq\max_{i,j}\{-\nu_{{\mathfrak{p}}}(c_{i,j})\}+\max_{i,j}\left\{-\sum_{i,j}e_{i,j}\nu_{\mathfrak{p}}(a_{i,j})\right\}
maxi,j{ν𝔭(ci,j)}+nmaxi,j{ν𝔭(ai,j)}.\displaystyle\leq\max_{i,j}\{-\nu_{\mathfrak{p}}(c_{i,j})\}+n\max_{i,j}\{-\nu_{\mathfrak{p}}(a_{i,j})\}.

Therefore T(F)T(F)+nT()T(F^{\mathcal{L}})\leq T(F)+nT({\mathcal{L}}) due to Remark 2.7.

2. Suppose that 𝐜=(c0,c1,c2){\bf c}=(c_{0},c_{1},c_{2}) and (𝐜)=(b0,b1,b2){\mathcal{L}}({\bf c})=(b_{0},b_{1},b_{2}). Then bi=j=13aj,icj1b_{i}=\sum_{j=1}^{3}a_{j,i}c_{j-1}. Let RR be a finite extension of k(t)k(t) containing all cic_{i} and ai,ja_{i,j}, and 𝔭{\mathfrak{p}} a place of RR. Then

ν𝔭(bi)=ν𝔭(j=13aj,icj1)minj{ν𝔭(aj,i)+ν𝔭(cj1)}\nu_{\mathfrak{p}}(b_{i})=\nu_{\mathfrak{p}}\left(\sum_{j=1}^{3}a_{j,i}c_{j-1}\right)\geq\min_{j}\{\nu_{\mathfrak{p}}(a_{j,i})+\nu_{\mathfrak{p}}(c_{j-1})\}

i.e.

ν𝔭(bi)maxj{ν𝔭(aj,i)ν𝔭(cj1)}maxj{ν𝔭(aj,i)}+maxj{ν𝔭(cj1)}.-\nu_{\mathfrak{p}}(b_{i})\leq\max_{j}\{-\nu_{\mathfrak{p}}(a_{j,i})-\nu_{\mathfrak{p}}(c_{j-1})\}\leq\max_{j}\{-\nu_{\mathfrak{p}}(a_{j,i})\}+\max_{j}\{-\nu_{\mathfrak{p}}(c_{j-1})\}.

So T((𝐜))T(𝐜)+T()T({\mathcal{L}}({\bf c}))\leq T({\bf c})+T({\mathcal{L}}). ∎

Corollary 3.7.

Suppose that 𝐜=(c0,c1,1)2(k(t)¯){\bf c}=(c_{0},c_{1},1)\in\mathbb{P}^{2}({\overline{k(t)}}). Let {\mathcal{L}} be a projective change of coordinates with

M=(a1a2a3a4a5a6c0c11)M_{\mathcal{L}}=\begin{pmatrix}a_{1}&a_{2}&a_{3}\\ a_{4}&a_{5}&a_{6}\\ c_{0}&c_{1}&1\end{pmatrix} (2)

where aika_{i}\in k. Then

  1. 1.

    ((0,0,1))=𝐜{\mathcal{L}}((0,0,1))={\bf c};

  2. 2.

    T(F),T(F1)T(F)+deg(F)T(𝐜)T(F^{{\mathcal{L}}}),T(F^{{\mathcal{L}}^{-1}})\leq T(F)+\deg(F)T({\bf c});

  3. 3.

    for each 𝐛2(k(t)¯){\bf b}\in\mathbb{P}^{2}({\overline{k(t)}}), T((𝐛)),T(1(𝐛))T(𝐛)+T(𝐜)T({\mathcal{L}}({\bf b})),T({\mathcal{L}}^{-1}({\bf b}))\leq T({\bf b})+T({\bf c}).

Proof.

The first assertion is obvious. The second and third assertions follows from Lemma 3.6 and the fact that T(M)=T(𝐜)T(M_{\mathcal{L}})=T({\bf c}) and T(M1)T(𝐜)T(M_{{\mathcal{L}}^{-1}})\leq T({\bf c}). ∎

Definition 3.8.
  1. 1.

    The points (1,0,0),(0,1,0),(0,0,1)2(k(t)¯)(1,0,0),(0,1,0),(0,0,1)\in\mathbb{P}^{2}({\overline{k(t)}}) are called fundamental points.

  2. 2.

    Assume (0,0,1)(0,0,1) is a singular point of F=0F=0 with multiplicity rr. F=0F=0 is said to be in excellent position if it satisfies the following two conditions:

    1. (a)

      the line x2=0x_{2}=0 intersects F=0F=0 in nn distinct non-fundamental points;

    2. (b)

      the lines x0=0,x1=0x_{0}=0,x_{1}=0 intersect F=0F=0 in nrn-r distinct points other than fundamental points.

Lemma 3.9.

Suppose that 𝐜=(c0,c1,c2){\bf c}=(c_{0},c_{1},c_{2}) is a singular point of F=0F=0 of multiplicity rr. There is a projective change of coordinates {\mathcal{L}} with MM_{\mathcal{L}} having the form (2) such that ((0,0,1))=𝐜{\mathcal{L}}((0,0,1))={\bf c} and F=0F^{\mathcal{L}}=0 is in excellent position.

Proof.

Denote 𝐲=(y1,,y6){\bf y}=(y_{1},\dots,y_{6}). Let 𝐲{\mathcal{L}}^{\prime}_{\bf y} be the projective change of coordinates with M𝐲M_{{\mathcal{L}}^{\prime}_{\bf y}} of the form

(y1y2y3y4y5y6c0c1c2).\begin{pmatrix}y_{1}&y_{2}&y_{3}\\ y_{4}&y_{5}&y_{6}\\ c_{0}&c_{1}&c_{2}\end{pmatrix}.

One sees that there are polynomials f1,,fsk(t)¯[y1,,y6]f_{1},\dots,f_{s}\in{\overline{k(t)}}[y_{1},\dots,y_{6}] such that 𝐛{\mathcal{L}}^{\prime}_{\bf b} with 𝐛k(t)¯6{\bf b}\in{\overline{k(t)}}^{6} satisfies the required conditions if and only if

𝐛S={𝐛k(t)¯6i=1,,s,fi(𝐛)0}.{\bf b}\in S=\left\{{\bf b}\in{\overline{k(t)}}^{6}\mid\forall\,i=1,\dots,s,\,f_{i}({\bf b})\neq 0\right\}.

Note that if {\mathcal{L}} is a projective change of coordinates such that (𝐜)=(0,0,1){\mathcal{L}}({\bf c})=(0,0,1) then =𝐛{\mathcal{L}}={\mathcal{L}}^{\prime}_{\bf b} for some 𝐛k(t)¯6{\bf b}\in{\overline{k(t)}}^{6}. Due to Lemma 1 on page of [9], there are projective changes of coordinates satisfying the above conditions. In other words, SS\neq\emptyset. Therefore Sk6S\cap k^{6}\neq\emptyset. For every 𝐛Sk6{\bf b}\in S\cap k^{6}, 𝐛{\mathcal{L}}^{\prime}_{\bf b} is as required. ∎

Lemma 3.10.
  1. 1.

    T(F𝒬)=T(F)T(F^{{\mathcal{Q}}})=T(F);

  2. 2.

    For each 𝐚=(a0,a1,a2)2(k(t)¯){\bf a}=(a_{0},a_{1},a_{2})\in\mathbb{P}^{2}({\overline{k(t)}}), T(𝒬(𝐚))2T(𝐚)T({\mathcal{Q}}({\bf a}))\leq 2T({\bf a}).

Proof.

1.1. Assume that F=i=0nj=0nici,jx0ix1jx2nijF=\sum_{i=0}^{n}\sum_{j=0}^{n-i}c_{i,j}x_{0}^{i}x_{1}^{j}x_{2}^{n-i-j} where ci,jk(t)¯c_{i,j}\in{\overline{k(t)}}. Then

F(x1x2,x0x2,x0x1)\displaystyle F(x_{1}x_{2},x_{0}x_{2},x_{0}x_{1}) =i=0nj=0nici,j(x1x2)i(x0x2)j(x0x1)nij\displaystyle=\sum_{i=0}^{n}\sum_{j=0}^{n-i}c_{i,j}(x_{1}x_{2})^{i}(x_{0}x_{2})^{j}(x_{0}x_{1})^{n-i-j}
=i=0nj=0nici,jx0nix1njx2i+j.\displaystyle=\sum_{i=0}^{n}\sum_{j=0}^{n-i}c_{i,j}x_{0}^{n-i}x_{1}^{n-j}x_{2}^{i+j}.

From this, one sees that the set of coefficients of FF is equal to that of F𝒬F^{{\mathcal{Q}}}. Hence T(F)=T(F𝒬)T(F)=T(F^{{\mathcal{Q}}}).

2.2. One has that 𝒬(𝐚)=(a1a2,a0a2,a0a1){\mathcal{Q}}({\bf a})=(a_{1}a_{2},a_{0}a_{2},a_{0}a_{1}). Let RR be a finite exntesion of k(t)k(t) containing all aia_{i} and 𝔭{\mathfrak{p}} a place of RR. Note that

ν𝔭(aiaj)=ν𝔭(ai)+ν𝔭(aj)2min{ν𝔭(a0),ν𝔭(a1),ν𝔭(a2)},\nu_{{\mathfrak{p}}}(a_{i}a_{j})=\nu_{{\mathfrak{p}}}(a_{i})+\nu_{{\mathfrak{p}}}(a_{j})\geq 2\min\{\nu_{{\mathfrak{p}}}(a_{0}),\nu_{{\mathfrak{p}}}(a_{1}),\nu_{{\mathfrak{p}}}(a_{2})\},

i.e. ν𝔭(aiaj)2max{ν𝔭(a0),ν𝔭(a1),ν𝔭(a2)}-\nu_{{\mathfrak{p}}}(a_{i}a_{j})\leq 2\max\{-\nu_{{\mathfrak{p}}}(a_{0}),-\nu_{{\mathfrak{p}}}(a_{1}),-\nu_{{\mathfrak{p}}}(a_{2})\}. So T(𝒬(𝐚))2T(𝐚)T({\mathcal{Q}}({\bf a}))\leq 2T({\bf a}). ∎

Definition 3.11.
  1. 1.

    We call a projective change of coordinates in Lemma 3.9 a projective change of coordinates centered at 𝐜{\bf c}.

  2. 2.

    Let 𝐜{\mathcal{L}}_{\bf c} be a projective change of coordinates centered at 𝐜{\bf c} and 𝒬{\mathcal{Q}} the stand quadratic transformation. We call 𝒬𝐜=𝐜𝒬{\mathcal{Q}}_{\bf c}={\mathcal{L}}_{\bf c}\circ{\mathcal{Q}}, the composition of 𝒬{\mathcal{Q}} and 𝐜{\mathcal{L}}_{\bf c}, a quadratic transformation centered at 𝐜{\bf c}.

Notation 3.12.

Let 𝒬𝐜{\mathcal{Q}}_{\bf c} be a quadratic transformation centered at 𝐜{\bf c}. We shall denote F𝒬𝐜=(F𝐜)𝒬F^{{\mathcal{Q}}_{\bf c}}=(F^{{\mathcal{L}}_{\bf c}})^{{\mathcal{Q}}}.

Corollary 3.13.

Let 𝐜{\bf c} be a singular point of F=0F=0 and 𝒬𝐜{\mathcal{Q}}_{\bf c} a quadratic transformation centered at 𝐜{\bf c}. Then

  1. 1.

    T(F𝒬𝐜)T(F)+deg(F)T(𝐜)T(F^{{\mathcal{Q}}_{\bf c}})\leq T(F)+\deg(F)T({\bf c});

  2. 2.

    for 𝐚2(k(t)¯){\bf a}\in\mathbb{P}^{2}({\overline{k(t)}}), T(𝒬𝐜1(𝐚))2(T(𝐜)+T(𝐚))T({\mathcal{Q}}_{\bf c}^{-1}({\bf a}))\leq 2(T({\bf c})+T({\bf a}));

Proof.

11 and 22 follow from the fact that 𝒬1=𝒬{\mathcal{Q}}^{-1}={\mathcal{Q}} and Lemmas 3.6 and 3.10. ∎

Proposition 3.14.

Let 𝒞(𝛏){\mathcal{C}}({\bm{\xi}}) be a plane projective model of {\mathcal{R}} defined by FF and 𝔓{\mathfrak{P}} a place of {\mathcal{R}}. Let 𝐚{\bf a} be the center of 𝔓{\mathfrak{P}} with respect to 𝛏{\bm{\xi}}. Assume that 𝒬𝐜{\mathcal{Q}}_{\bf c} is a quadratic transformation centered at 𝐜{\bf c} for some singular point 𝐜{\bf c} of F=0F=0 and 𝐚{\bf a}^{\prime} is the center of 𝔓{\mathfrak{P}} with respect to 𝒬𝐜1(𝛏){\mathcal{Q}}_{\bf c}^{-1}({\bm{\xi}}). Then

T(𝐚)max{2(T(𝐜)+T(𝐚)),T(F)+deg(F)T(𝐜)}.T({\bf a}^{\prime})\leq\max\{2(T({\bf c})+T({\bf a})),T(F)+\deg(F)T({\bf c})\}.
Proof.

We first claim that if 𝐚𝐜{\bf a}\neq{\bf c} then 𝒬𝐜1(𝐚)(0,0,0){\mathcal{Q}}_{\bf c}^{-1}({\bf a})\neq(0,0,0). Otherwise assume that 𝒬𝐜1(𝐚)=𝒬1𝐜1(𝐚)=(0,0,0){\mathcal{Q}}_{\bf c}^{-1}({\bf a})={\mathcal{Q}}^{-1}{\mathcal{L}}_{\bf c}^{-1}({\bf a})=(0,0,0). Then 𝐜1(𝐚){\mathcal{L}}_{\bf c}^{-1}({\bf a}) is a fundamental point of F𝐜=0F^{{\mathcal{L}}_{\bf c}}=0. Since neither (1,0,0)(1,0,0) nor (0,1,0)(0,1,0) is a point of F𝐜=0F^{{\mathcal{L}}_{\bf c}}=0. One has that 𝐜1(𝐚)=(0,0,1){\mathcal{L}}_{\bf c}^{-1}({\bf a})=(0,0,1). Hence 𝐚=𝐜{\bf a}={\bf c}. This proves our claim.

Suppose that 𝐚𝐜{\bf a}\neq{\bf c}. Then 𝒬𝐜1(𝐚)(0,0,0){\mathcal{Q}}_{\bf c}^{-1}({\bf a})\neq(0,0,0) and thus 𝐚=𝒬𝐜1(𝐚){\bf a}^{\prime}={\mathcal{Q}}_{\bf c}^{-1}({\bf a}) by Proposition 2.5. Corollary 3.13 then implies that T(𝐚)2(T(𝐚)+T(𝐜)).T({\bf a}^{\prime})\leq 2(T({\bf a})+T({\bf c})). Now suppose that 𝐚=𝐜{\bf a}={\bf c}. Denote 𝝃=𝐜1(𝝃)=(ξ0,ξ1,ξ2){\bm{\xi}}^{\prime}={\mathcal{L}}_{\bf c}^{-1}({\bm{\xi}})=(\xi_{0}^{\prime},\xi_{1}^{\prime},\xi_{2}^{\prime}). By Proposition 2.5 again, (0,0,1)(0,0,1) is the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}}^{\prime}. Suppose that uu is a local uniformizer of 𝔓{\mathfrak{P}} and i=ν𝔓(ξi)\ell_{i}=\nu_{{\mathfrak{P}}}(\xi_{i}^{\prime}). From the definition of center, one sees that i>2\ell_{i}>\ell_{2} for all i=0,1i=0,1. Write ξi=ui(ci+uηi)\xi_{i}^{\prime}=u^{\ell_{i}}(c_{i}+u\eta_{i}) where cik(t)¯{0}c_{i}\in{\overline{k(t)}}\setminus\{0\}, ηi\eta_{i}\in{\mathcal{R}} with ν𝔓(ηi)0\nu_{{\mathfrak{P}}}(\eta_{i})\geq 0. One then has that

𝒬1(𝝃)\displaystyle{\mathcal{Q}}^{-1}({\bm{\xi}}^{\prime}) =(ξ1ξ2,ξ0ξ2,ξ0ξ1)\displaystyle=(\xi_{1}^{\prime}\xi_{2}^{\prime},\xi_{0}^{\prime}\xi_{2}^{\prime},\xi_{0}^{\prime}\xi_{1}^{\prime})
=(u1+2(c1c2+uη0~),u0+2(c0c2+uη1~),u1+0(c0c1+uη2~))\displaystyle=\left(u^{\ell_{1}+\ell_{2}}(c_{1}c_{2}+u\tilde{\eta_{0}}),u^{\ell_{0}+\ell_{2}}(c_{0}c_{2}+u\tilde{\eta_{1}}),u^{\ell_{1}+\ell_{0}}(c_{0}c_{1}+u\tilde{\eta_{2}})\right)

where ηi~\tilde{\eta_{i}}\in{\mathcal{R}} with ν𝔓(ηi~)0\nu_{{\mathfrak{P}}}(\tilde{\eta_{i}})\geq 0. Set

μ=min{ν𝔓(ξ1ξ2),ν𝔓(ξ0ξ2),ν𝔓(ξ1ξ0)}.\mu=\min\{\nu_{{\mathfrak{P}}}(\xi_{1}^{\prime}\xi_{2}^{\prime}),\nu_{{\mathfrak{P}}}(\xi_{0}^{\prime}\xi_{2}^{\prime}),\nu_{{\mathfrak{P}}}(\xi_{1}^{\prime}\xi_{0}^{\prime})\}.

Since both 0\ell_{0} and 1\ell_{1} are greater than 2\ell_{2}, μ=0+2\mu=\ell_{0}+\ell_{2} or 1+2\ell_{1}+\ell_{2}. In the case that μ=0+2=1+2\mu=\ell_{0}+\ell_{2}=\ell_{1}+\ell_{2}, one has that 𝐚=(c1c2,c0c2,0)=c2(c1,c0,0){\bf a}^{\prime}=(c_{1}c_{2},c_{0}c_{2},0)=c_{2}(c_{1},c_{0},0). So 𝐚𝕍(F𝒬𝐜)𝕍(x2){\bf a}^{\prime}\in\mathbb{V}(F^{{\mathcal{Q}}_{\bf c}})\cap\mathbb{V}(x_{2}). By Propositon 3.4, T(𝐚)T(F𝒬𝐜)T(F)+deg(F)T(𝐜)T({\bf a}^{\prime})\leq T(F^{{\mathcal{Q}}_{\bf c}})\leq T(F)+\deg(F)T({\bf c}). In other cases, one sees that 𝐚{\bf a}^{\prime} is a fundamental point and so T(𝐚)=0T({\bf a}^{\prime})=0. Hence, in each case, one has that

T(𝐚)max{2(T(𝐜)+T(𝐚)),T(F)+deg(F)T(𝐜)}.T({\bf a}^{\prime})\leq\max\{2(T({\bf c})+T({\bf a})),T(F)+\deg(F)T({\bf c})\}.

3.2 Degrees and Heights for Riemann-Roch Spaces

Let D=i=1mni𝔓iD=\sum_{i=1}^{m}n_{i}{\mathfrak{P}}_{i} be a disivor in {\mathcal{R}} where ni0n_{i}\neq 0. Let 𝒞(𝝃){\mathcal{C}}({\bm{\xi}}) defined by FF be a plane projective model of {\mathcal{R}}. Suppose that h𝔏(D)h\in{\mathfrak{L}}(D) and h=G(𝝃)/H(𝝃)h=G({\bm{\xi}})/H({\bm{\xi}}) where G,HG,H are two homogeneous polynomials of the same degree in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}]. In this subsection, we shall estimate deg(G)\deg(G) and T(G),T(H)T(G),T(H) in terms of deg(F)\deg(F) and T(F)T(F). For this, we introduce the following notations and definitions.

Notation 3.15.

Let 𝒞(𝛏){\mathcal{C}}({\bm{\xi}}) be a plane projective model of {\mathcal{R}} and DD a divisor.

  1. 1.

    𝒮𝝃(D):={the centers of places in supp(D) with respect to 𝝃}{\mathcal{S}}_{\bm{\xi}}(D):=\{\mbox{the centers of places in ${\rm supp}(D)$ with respect to ${\bm{\xi}}$}\}.

  2. 2.

    T𝝃(D):=max{T(𝐜)𝐜𝒮𝝃(D)}T_{\bm{\xi}}(D):=\max\left\{T({\bf c})\mid{\bf c}\in{\mathcal{S}}_{\bm{\xi}}(D)\right\}.

Definition 3.16.

Let G,HG,H be two homogeneous polynomials in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] of the same degree. Write 𝛏=(ξ0,ξ1,ξ2){\bm{\xi}}=(\xi_{0},\xi_{1},\xi_{2}).

  • (1).(1).

    Define

    ord𝔓(G(𝝃))=ν𝔓(G(𝝃))deg(G)mini=02{ν𝔓(ξi)}.{\rm ord}_{{\mathfrak{P}}}(G({\bm{\xi}}))=\nu_{{\mathfrak{P}}}\left(G({\bm{\xi}})\right)-\deg(G)\min_{i=0}^{2}\{\nu_{{\mathfrak{P}}}(\xi_{i})\}.
  • (2).(2).

    Define

    div𝝃(G)=ord𝔓(G(𝝃))𝔓{\rm div}_{\bm{\xi}}(G)=\sum{\rm ord}_{{\mathfrak{P}}}(G({\bm{\xi}})){\mathfrak{P}}

    where the sum ranges over all places of {\mathcal{R}}. Furthermore, define

    div𝝃(G/H)=div𝝃(G)div𝝃(H).{\rm div}_{\bm{\xi}}(G/H)={\rm div}_{\bm{\xi}}(G)-{\rm div}_{\bm{\xi}}(H).

It is easy to see that ord𝔓(G(𝝃))0{\rm ord}_{{\mathfrak{P}}}(G({\bm{\xi}}))\geq 0 and ord𝔓(G(𝝃))>0{\rm ord}_{\mathfrak{P}}(G({\bm{\xi}}))>0 if and only if G(𝐜)=0G({\bf c})=0 where 𝐜{\bf c} is the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}}. Futhermore ord𝔓(G(λ𝝃))=ord𝔓(G(𝝃)){\rm ord}_{{\mathfrak{P}}}(G(\lambda{\bm{\xi}}))={\rm ord}_{{\mathfrak{P}}}(G({\bm{\xi}})) for all nonzero λ\lambda\in{\mathcal{R}}.

Remark 3.17.

On page 182 of [9], ord𝔓(G){\rm ord}_{{\mathfrak{P}}}(G) is defined to be the order at 𝔓{\mathfrak{P}} of the image of GG in the valutaion ring of 𝔓{\mathfrak{P}}. Remark that

ord𝔓(G(𝝃))=ν𝔓(G(𝝃)/ξi0d){\rm ord}_{\mathfrak{P}}(G({\bm{\xi}}))=\nu_{{\mathfrak{P}}}(G({\bm{\xi}})/\xi_{i_{0}}^{d})

where ξi0\xi_{i_{0}} satisfies that ν𝔓(ξi0)=mini=02{ν𝔓(ξi)}\nu_{{\mathfrak{P}}}(\xi_{i_{0}})=\min_{i=0}^{2}\{\nu_{\mathfrak{P}}(\xi_{i})\}. Under the map sending xjx_{j} to ξj/ξi0\xi_{j}/\xi_{i_{0}} for all j=0,1,2j=0,1,2, GG is sent to G(𝛏)/ξi0dG({\bm{\xi}})/\xi_{i_{0}}^{d} which lies in the valuation ring of 𝔓{\mathfrak{P}}. Therefore ord𝔓{\rm ord}_{\mathfrak{P}} given in Definition 3.16 concides with the one given in [9] and div𝛏(G){\rm div}_{\bm{\xi}}(G) is nothing else but the intersection cycle of GG and HH (see page 119 of [9]).

The lemma below follows easily from the definition.

Lemma 3.18.

Suppose that G,Hk(t)¯[x0,x1,x2]G,H\in{\overline{k(t)}}[x_{0},x_{1},x_{2}] are two homogeneous polynomials of the same degree. Then

  1. 1.

    div𝝃(GH)=div𝝃(G)div𝝃(H)=div(G(𝝃)H(𝝃)){\rm div}_{\bm{\xi}}\left(\frac{G}{H}\right)={\rm div}_{\bm{\xi}}(G)-{\rm div}_{\bm{\xi}}(H)={\rm div}\left(\frac{G({\bm{\xi}})}{H({\bm{\xi}})}\right); and

  2. 2.

    deg(div𝝃(G))=deg(G)deg(F)\deg({\rm div}_{\bm{\xi}}(G))=\deg(G)\deg(F).

Lemma 3.19.

Suppose that 𝐜=(c0,c1,1){\bf c}=(c_{0},c_{1},1) is an ordinary singular point of F=0F=0 of multiplicity rr, n=deg(F)n=\deg(F) and SS is a finite set of points of F=0F=0.

  1. 1.

    Let Lλ=x0c0x2λ(x1c1x2)L_{\lambda}=x_{0}-c_{0}x_{2}-\lambda(x_{1}-c_{1}x_{2}). Then for all but a finite number of λ\lambda, Lλ=0L_{\lambda}=0 intersects F=0F=0 in nrn-r distinct points other than the points in {𝐜}S\{{\bf c}\}\cup S.

  2. 2.

    Write

    F=Fr(x0c0x2,x1c1x2)x2nr++Fn(x0c0x2,x1c1x2)F=F_{r}(x_{0}-c_{0}x_{2},x_{1}-c_{1}x_{2})x_{2}^{n-r}+\dots+F_{n}(x_{0}-c_{0}x_{2},x_{1}-c_{1}x_{2})

    where Fi(y0,y1)F_{i}(y_{0},y_{1}) is a homogeneous polynomial in y0,y1y_{0},y_{1} of degree ii. Assume that Fr(1,0)Fr(0,1)Fn(1,0)Fn(0,1)0F_{r}(1,0)F_{r}(0,1)F_{n}(1,0)F_{n}(0,1)\neq 0. Let

    Gλ=α(x1c1x2)x2(x0c0x2)x2λ(x0c0x2)(x1c1x2)G_{\lambda}=\alpha(x_{1}-c_{1}x_{2})x_{2}-(x_{0}-c_{0}x_{2})x_{2}-\lambda(x_{0}-c_{0}x_{2})(x_{1}-c_{1}x_{2})

    where αk(t)¯{0}\alpha\in{\overline{k(t)}}\setminus\{0\} satisfies that Fr(α,1)=0F_{r}(\alpha,1)=0. Then for all but a finite number of λ\lambda, Gλ=0G_{\lambda}=0 intersects F=0F=0 in 2nr12n-r-1 distinct points other than the points in {𝐜}S\{{\bf c}\}\cup S.

Proof.

1.1. Under the projective change of coordinates {\mathcal{L}} with (x0)=x0+c0x2,(x1)=x1+c1x2{\mathcal{L}}(x_{0})=x_{0}+c_{0}x_{2},{\mathcal{L}}(x_{1})=x_{1}+c_{1}x_{2} and (x2)=x2{\mathcal{L}}(x_{2})=x_{2}, we may assume that 𝐜=(0,0,1){\bf c}=(0,0,1). Set Lλ=x0λx1L_{\lambda}=x_{0}-\lambda x_{1} where λk(t)¯\lambda\in{\overline{k(t)}}. Substituting x0=λx1x_{0}=\lambda x_{1} into FF yields that

x1r(Fr(λ,1)x2nr+Fr+1(λ,1)x1x2nr1++Fn(λ,1)x1nr).x_{1}^{r}\left(F_{r}(\lambda,1)x_{2}^{n-r}+F_{r+1}(\lambda,1)x_{1}x_{2}^{n-r-1}+\dots+F_{n}(\lambda,1)x_{1}^{n-r}\right).

Set Hλ(t)=Fλ(λ,1)tnr++Fn(λ,1)H_{\lambda}(t)=F_{\lambda}(\lambda,1)t^{n-r}+\dots+F_{n}(\lambda,1). For every root γ\gamma of Hλ(t)=0H_{\lambda}(t)=0, one sees that (λ,1,γ)(\lambda,1,\gamma) is a common point of Lλ=0L_{\lambda}=0 and F=0F=0 other than 𝐜{\bf c}. Moreover if Hλ(t)=0H_{\lambda}(t)=0 has nrn-r distinct roots then Lλ=0L_{\lambda}=0 intersects F=0F=0 in nrn-r distinct points other than 𝐜{\bf c}. So it suffices to prove that for all but a finite number of λ\lambda, Hλ(t)=0H_{\lambda}(t)=0 has nrn-r distinct roots. Note tht substituting x0=λx1x_{0}=\lambda x_{1} into F/x2\partial F/\partial x_{2} yields that

x1ri=rn(ni)Fi(λ,1)x2ni1x1ir.x_{1}^{r}\sum_{i=r}^{n}(n-i)F_{i}(\lambda,1)x_{2}^{n-i-1}x_{1}^{i-r}.

From this, one sees that if γ\gamma is a common root of Hλ(t)=0H_{\lambda}(t)=0 and Hλ(t)/t=0\partial H_{\lambda}(t)/\partial t=0 then (λ,1,γ)(\lambda,1,\gamma) is a common point of F=0F=0 and F/x2=0\partial F/\partial x_{2}=0. Since F=0F=0 and F/x2=0\partial F/\partial x_{2}=0 have only finitely many common points, there are only finitely many λ\lambda such that Hλ(t)=0H_{\lambda}(t)=0 has multiple roots. In other words, there is only a finite number of λ\lambda such that Lλ=0L_{\lambda}=0 intersects F=0F=0 in less than nrn-r distinct points other than 𝐜{\bf c}. It remains to show that there are only finitely many λ\lambda such that (S{𝐜})𝕍(Lλ)(S\setminus\{{\bf c}\})\cap\mathbb{V}(L_{\lambda})\neq\emptyset. Assume that (a0,a1,a2)S{𝐜}(a_{0},a_{1},a_{2})\in S\setminus\{{\bf c}\} which lies in the line Lλ=0L_{\lambda}=0. If a1c1a20a_{1}-c_{1}a_{2}\neq 0 then

λ=(a0c2a2)/(a1c1a2).\lambda=(a_{0}-c_{2}a_{2})/(a_{1}-c_{1}a_{2}).

If a1c1a2=0a_{1}-c_{1}a_{2}=0 then a0c0a2=0a_{0}-c_{0}a_{2}=0 and thus (a0,a1,a2)=a2(c0,c1,1)(a_{0},a_{1},a_{2})=a_{2}(c_{0},c_{1},1). In other words, (a0,a1,a2)=(0,0,0)(a_{0},a_{1},a_{2})=(0,0,0) or 𝐜{\bf c}, which is impossible.

2.2. Similarly, we may assume that 𝐜=(0,0,1){\bf c}=(0,0,1). Then

Gλ\displaystyle G_{\lambda} =αx1x2x0x2λx0x1,\displaystyle=\alpha x_{1}x_{2}-x_{0}x_{2}-\lambda x_{0}x_{1},
F\displaystyle F =Fr(x0,x1)x2nr++Fn(x0,x1).\displaystyle=F_{r}(x_{0},x_{1})x_{2}^{n-r}+\dots+F_{n}(x_{0},x_{1}).

Applying the standard quadratic transformation 𝒬{\mathcal{Q}} to GλG_{\lambda} and FF, one obtains that

Gλ𝒬\displaystyle G_{\lambda}^{{\mathcal{Q}}} =αx0x1λx2,\displaystyle=\alpha x_{0}-x_{1}-\lambda x_{2},
F(x1x2,x0x2,x0x1)\displaystyle F(x_{1}x_{2},x_{0}x_{2},x_{0}x_{1}) =x2r(Fr(x1,x0)(x0x1)nr++Fn(x1,x0)x2nr).\displaystyle=x_{2}^{r}\left(F_{r}(x_{1},x_{0})(x_{0}x_{1})^{n-r}+\dots+F_{n}(x_{1},x_{0})x_{2}^{n-r}\right).

Since 𝐜{\bf c} is an ordinary singular point and Fr(1,0)Fr(0,1)0F_{r}(1,0)F_{r}(0,1)\neq 0, by on page of [9] (1,α,0)(1,\alpha,0) is a simple point of F𝒬=0F^{{\mathcal{Q}}}=0. Moveover, as Fn(1,0)Fn(0,1)0F_{n}(1,0)F_{n}(0,1)\neq 0, neither (1,0,0)(1,0,0) nor (0,1,0)(0,1,0) is a point of F𝒬=0F^{{\mathcal{Q}}}=0 and so deg(F𝒬)=2nr\deg(F^{{\mathcal{Q}}})=2n-r. Thus

F𝒬=Fr(x1,x0)(x0x1)nr++Fn(x1,x0)x2nr.F^{\mathcal{Q}}=F_{r}(x_{1},x_{0})(x_{0}x_{1})^{n-r}+\dots+F_{n}(x_{1},x_{0})x_{2}^{n-r}.

For every common point (γ0,γ1,γ2)(\gamma_{0},\gamma_{1},\gamma_{2}) of Gλ𝒬=0G_{\lambda}^{{\mathcal{Q}}}=0 and F𝒬=0F^{{\mathcal{Q}}}=0 with λγ20\lambda\gamma_{2}\neq 0, (γ1γ2,γ0γ2,γ0γ1)(0,0,0)(\gamma_{1}\gamma_{2},\gamma_{0}\gamma_{2},\gamma_{0}\gamma_{1})\neq(0,0,0) and then it is a common point of Gλ=0G_{\lambda}=0 and F=0F=0 other than 𝐜{\bf c}. Therefore it suffices to show that for all but a finite number of λ\lambda, Gλ𝒬=0G_{\lambda}^{{\mathcal{Q}}}=0 and F𝒬=0F^{{\mathcal{Q}}}=0 have 2nr12n-r-1 distinct common points (γ0,γ1,γ2)(\gamma_{0},\gamma_{1},\gamma_{2}) with γ20\gamma_{2}\neq 0. Let {\mathcal{L}} be the projective change of coordinates such that (x0)=(x0+x1)/α,(x1)=x1,(x2)=x2{\mathcal{L}}(x_{0})=(x_{0}+x_{1})/\alpha,{\mathcal{L}}(x_{1})=x_{1},{\mathcal{L}}(x_{2})=x_{2}. Then (Gλ𝒬)=x0λx2(G_{\lambda}^{{\mathcal{Q}}})^{\mathcal{L}}=x_{0}-\lambda x_{2}. Note that 1((1,α,0))=(0,α,0){\mathcal{L}}^{-1}((1,\alpha,0))=(0,\alpha,0) which is a simple point of (F𝒬)=0(F^{{\mathcal{Q}}})^{\mathcal{L}}=0. Thus

(F𝒬)=F~1(x0,x2)x12nr1+F~2(x0,x2)x12nr2++F~2nr(x0,x2).(F^{{\mathcal{Q}}})^{\mathcal{L}}=\tilde{F}_{1}(x_{0},x_{2})x_{1}^{2n-r-1}+\tilde{F}_{2}(x_{0},x_{2})x_{1}^{2n-r-2}+\dots+\tilde{F}_{2n-r}(x_{0},x_{2}).

By (1)(1), for all but a finite number of λ\lambda, (Gλ𝒬)=0(G_{\lambda}^{{\mathcal{Q}}})^{\mathcal{L}}=0 intersects (F𝒬)=0(F^{{\mathcal{Q}}})^{\mathcal{L}}=0 in 2nr12n-r-1 distinct points (γ0,γ1,γ2)(\gamma_{0}^{\prime},\gamma_{1}^{\prime},\gamma_{2}^{\prime}) with γ20\gamma_{2}^{\prime}\neq 0. Remark that if (γ0,γ1,γ2)(\gamma^{\prime}_{0},\gamma_{1}^{\prime},\gamma_{2}^{\prime}) is a common point of (Gλ𝒬)=0(G_{\lambda}^{{\mathcal{Q}}})^{\mathcal{L}}=0 and (F𝒬)=0(F^{{\mathcal{Q}}})^{\mathcal{L}}=0 with γ20\gamma_{2}^{\prime}\neq 0 then (γ0+β/αγ1,γ1,γ2)(\gamma_{0}^{\prime}+\beta/\alpha\gamma_{1}^{\prime},\gamma_{1}^{\prime},\gamma_{2}^{\prime}) is a common point of Gλ𝒬=0G_{\lambda}^{{\mathcal{Q}}}=0 and F𝒬=0F^{{\mathcal{Q}}}=0 with γ20\gamma_{2}^{\prime}\neq 0. These imply that for all but a finite number of λ\lambda, Gλ𝒬=0G_{\lambda}^{{\mathcal{Q}}}=0 intersects F𝒬=0F^{{\mathcal{Q}}}=0 in 2nr12n-r-1 distinct points (γ0,γ1,γ2)(\gamma_{0},\gamma_{1},\gamma_{2}) with γ20\gamma_{2}\neq 0.

Finally, we need to prove that there are only finitely many λ\lambda such that S{𝐜}𝕍(Gλ)S\setminus\{{\bf c}\}\cap\mathbb{V}(G_{\lambda})\neq\emptyset. Assume that 𝐚=(a0,a1,a2)S{𝐜}{\bf a}=(a_{0},a_{1},a_{2})\in S\setminus\{{\bf c}\} which lies in Gλ=0G_{\lambda}=0. We claim that (a0c0a2)(a1c1a2)0(a_{0}-c_{0}a_{2})(a_{1}-c_{1}a_{2})\neq 0. Suppose on the contrary that (a0c0a2)(a1c1a2)=0(a_{0}-c_{0}a_{2})(a_{1}-c_{1}a_{2})=0. Then by Gλ(𝐚)=0G_{\lambda}({\bf a})=0, one sees that either a2=0a_{2}=0 or both a0c0a2a_{0}-c_{0}a_{2} and a1c1a2a_{1}-c_{1}a_{2} are zero. This implies that 𝐚{\bf a} must be one of three points (1,0,0),(0,1,0),a2(c0,c1,1)(1,0,0),(0,1,0),a_{2}(c_{0},c_{1},1). This is impossible and then our claim holds. It follows from the claim that λ\lambda is uniquely determined by 𝐚{\bf a}. ∎

Proposition 3.20.

Suppose that F=0F=0 has only ordinary singularities and DD is an effective divisor in {\mathcal{R}}. Let DD^{\prime} be a divisor in {\mathcal{R}}.

  1. 1.

    Assume further that D=i=1r𝔓iD=\sum_{i=1}^{r}{\mathfrak{P}}_{i} where all 𝔓i{\mathfrak{P}}_{i} have the same center which is a point of F=0F=0 with multiplicity rr. Then there is a linear homogeneous polynomial GG in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] such that

    div𝝃(G)=D+A{\rm div}_{\bm{\xi}}(G)=D+A

    where AA is a very simple and effective divisor of degree nrn-r, supp(A)(supp(D){𝔓1,,𝔓r})={\rm supp}(A)\cap({\rm supp}(D^{\prime})\cup\{{\mathfrak{P}}_{1},\dots,{\mathfrak{P}}_{r}\})=\emptyset, and

    T(G)T𝝃(D),T𝝃(A)2(T(F)+nT𝝃(D)).T(G)\leq T_{\bm{\xi}}(D),\,\,T_{\bm{\xi}}(A)\leq 2(T(F)+nT_{\bm{\xi}}(D)).
  2. 2.

    Assume that D=𝔓D={\mathfrak{P}} where the center of 𝔓{\mathfrak{P}} is a singular point of F=0F=0. Then there are two homogeneous polynomials G,Hk(t)¯[x0,x1,x2]G,H\in{\overline{k(t)}}[x_{0},x_{1},x_{2}] of degree two such that

    div𝝃(G/H)=D+A{\rm div}_{\bm{\xi}}(G/H)=D+A

    where AA is a very simple divisor, supp(A)(supp(D){𝔓})={\rm supp}(A)\cap({\rm supp}(D^{\prime})\cup\{{\mathfrak{P}}\})=\emptyset and

    T(G),T(H)T(F)+nT𝝃(D),T𝝃(A)(2n+4)T(F)+2n2T𝝃(D).T(G),T(H)\leq T(F)+nT_{\bm{\xi}}(D),\,\,T_{\bm{\xi}}(A)\leq(2n+4)T(F)+2n^{2}T_{\bm{\xi}}(D).
Proof.

1.1. Suppose 𝐜=(c0,c1,c2){\bf c}=(c_{0},c_{1},c_{2}) is the center of 𝔓i{\mathfrak{P}}_{i} with respect to 𝝃{\bm{\xi}}. Without loss of generality, we assume that c20c_{2}\neq 0 and 𝐜=(c0,c1,1){\bf c}=(c_{0},c_{1},1). Set Lλ=x0c0x2λ(x1c1x2)L_{\lambda}=x_{0}-c_{0}x_{2}-\lambda(x_{1}-c_{1}x_{2}). Due to Lemma 3.19, for all but a finite number of λ\lambda, Lλ=0L_{\lambda}=0 intersects F=0F=0 in nrn-r distincet points other than the points in {𝐜}𝒮𝝃(D)\{{\bf c}\}\cup{\mathcal{S}}_{\bm{\xi}}(D^{\prime}). Let λk\lambda^{\prime}\in k be such that Lλ=0L_{\lambda^{\prime}}=0 satisfies the above condition. Then

div𝝃(Lλ)=i=1r𝔓i+A{\rm div}_{{\bm{\xi}}}(L_{\lambda^{\prime}})=\sum_{i=1}^{r}{\mathfrak{P}}_{i}+A

where AA is an effective divisor of degree nrn-r and supp(A)(supp(D){𝔓1,,𝔓r})={\rm supp}(A)\cap({\rm supp}(D^{\prime})\cup\{{\mathfrak{P}}_{1},\dots,{\mathfrak{P}}_{r}\})=\emptyset. It is clear that AA is very simple since Lλ=0L_{\lambda^{\prime}}=0 intersects FF in nrn-r distinct points other than 𝐜{\bf c}. Finally, one easily sees that T(Lλ)T(𝐜)=T𝝃(D)T(L_{\lambda^{\prime}})\leq T({\bf c})=T_{\bm{\xi}}(D). As the points in T𝝃(A)T_{\bm{\xi}}(A) are the intersection points of F=0F=0 and Lλ=0L_{\lambda^{\prime}}=0, T𝝃(A)2(T(F)+nT𝝃(D))T_{\bm{\xi}}(A)\leq 2(T(F)+nT_{\bm{\xi}}(D)) by Proposition 3.4.

2.2. Suppose that 𝐜=(c0,c1,c2){\bf c}=(c_{0},c_{1},c_{2}) is the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}}, and 𝐜{\bf c} is of multiplicity r>0r>0. Since 𝐜{\bf c} is an ordinary singular point, there are exactly rr places of {\mathcal{R}} with 𝐜{\bf c} as the center with respect to 𝝃{\bm{\xi}}. Denote these rr places by 𝔓1=𝔓,,𝔓r{\mathfrak{P}}_{1}={\mathfrak{P}},\dots,{\mathfrak{P}}_{r}. Without loss of generality, we may assume that c20c_{2}\neq 0 and 𝐜=(c0,c1,1){\bf c}=(c_{0},c_{1},1). Write

F=Fr(x0c0x2,x1c1x2)x2nr++Fn(x0c0x2,x1c1x2)F=F_{r}(x_{0}-c_{0}x_{2},x_{1}-c_{1}x_{2})x_{2}^{n-r}+\dots+F_{n}(x_{0}-c_{0}x_{2},x_{1}-c_{1}x_{2})

where Fi(y0,y1)F_{i}(y_{0},y_{1}) is a homogeneous polynomial of degree ii. Choose a projective change of coordinates {\mathcal{L}} with M=diag(B,1),BGL2(k)M_{\mathcal{L}}={\rm diag}(B,1),B\in{\rm GL}_{2}(k) such that

F=F~r(x0c~0x2,x1c~1x2)x2nr++F~n(x0c~0x2,x1c~1x2)F^{{\mathcal{L}}}=\tilde{F}_{r}(x_{0}-\tilde{c}_{0}x_{2},x_{1}-\tilde{c}_{1}x_{2})x_{2}^{n-r}+\dots+\tilde{F}_{n}(x_{0}-\tilde{c}_{0}x_{2},x_{1}-\tilde{c}_{1}x_{2})

satisfies that F~r(1,0)F~r(0,1)F~n(1,0)F~n(0,1)0\tilde{F}_{r}(1,0)\tilde{F}_{r}(0,1)\tilde{F}_{n}(1,0)\tilde{F}_{n}(0,1)\neq 0, where F~i=Fi((y0,y1)B)\tilde{F}_{i}=F_{i}((y_{0},y_{1})B) and (c~0,c~1)=(c0,c1)B1(\tilde{c}_{0},\tilde{c}_{1})=(c_{0},c_{1})B^{-1}. By Lemma 3.6, T(F)T(F)T(F^{\mathcal{L}})\leq T(F). Denote

𝝃~=(ξ~0,ξ~1,ξ~2)=𝝃M1.\tilde{{\bm{\xi}}}=(\tilde{\xi}_{0},\tilde{\xi}_{1},\tilde{\xi}_{2})={\bm{\xi}}M_{\mathcal{L}}^{-1}.

Then 𝐜~=(c~0,c~1,1)=𝐜M1\tilde{{\bf c}}=(\tilde{c}_{0},\tilde{c}_{1},1)={\bf c}M_{\mathcal{L}}^{-1} is the center of 𝔓1{\mathfrak{P}}_{1} with respect to 𝝃~\tilde{{\bm{\xi}}}. For i=0,1i=0,1, write

ξ~i/ξ~2=c~i+αiud+ud+1ηi\tilde{\xi}_{i}/\tilde{\xi}_{2}=\tilde{c}_{i}+\alpha_{i}u^{d}+u^{d+1}\eta_{i}

where uu is a local uniformizer of 𝔓1{\mathfrak{P}}_{1}, d1d\geq 1, αik(t)¯\alpha_{i}\in{\overline{k(t)}} not all zero, and ν𝔓(ηi)0\nu_{{\mathfrak{P}}}(\eta_{i})\geq 0. Furthermore,

0=F(ξ~0/ξ~2,ξ~1/ξ~2,1)=udrF~r(α0,α1)+udr+1β0=F^{\mathcal{L}}(\tilde{\xi}_{0}/\tilde{\xi}_{2},\tilde{\xi}_{1}/\tilde{\xi}_{2},1)=u^{dr}\tilde{F}_{r}(\alpha_{0},\alpha_{1})+u^{dr+1}\beta

where ν𝔓1(β)0\nu_{{\mathfrak{P}}_{1}}(\beta)\geq 0. This implies that F~r(α0,α1)=0\tilde{F}_{r}(\alpha_{0},\alpha_{1})=0. Since F~r(0,1)F~r(1,0)0\tilde{F}_{r}(0,1)\tilde{F}_{r}(1,0)\neq 0, α0α10\alpha_{0}\alpha_{1}\neq 0. Set α¯=α0/α1\bar{\alpha}=\alpha_{0}/\alpha_{1} and

G~λ=α¯(x1c~1x2)x2(x0c~0x2)x2λ(x0c~0x2)(x1c~1x2).\tilde{G}_{\lambda}=\bar{\alpha}(x_{1}-\tilde{c}_{1}x_{2})x_{2}-(x_{0}-\tilde{c}_{0}x_{2})x_{2}-\lambda(x_{0}-\tilde{c}_{0}x_{2})(x_{1}-\tilde{c}_{1}x_{2}).

Due to Lemma 3.19, for all but a finite number of λ\lambda, G~λ\tilde{G}_{\lambda} intersects F=0F^{\mathcal{L}}=0 in 2nr12n-r-1 distinct points other than the points {𝐜~}𝒮𝝃~(D)\{\tilde{{\bf c}}\}\cup{\mathcal{S}}_{\tilde{{\bm{\xi}}}}(D^{\prime}). Let AλA_{\lambda} be the very simple divisor consisting of the 2nr12n-r-1 places whose centers with respect to 𝝃~\tilde{{\bm{\xi}}} are the intersection points of G~λ=0\tilde{G}_{\lambda}=0 and F=0F^{\mathcal{L}}=0 other than 𝐜~\tilde{{\bf c}} respectively. Then supp(Aλ)(supp(D){𝔓1,,𝔓r})={\rm supp}(A_{\lambda})\cap({\rm supp}(D^{\prime})\cup\{{\mathfrak{P}}_{1},\dots,{\mathfrak{P}}_{r}\})=\emptyset. We claim that for the above G~λ\tilde{G}_{\lambda},

div𝝃~(G~λ)=𝔓1+i=1r𝔓i+Aλ{\rm div}_{\tilde{{\bm{\xi}}}}(\tilde{G}_{\lambda})={\mathfrak{P}}_{1}+\sum_{i=1}^{r}{\mathfrak{P}}_{i}+A_{\lambda}

Note that

G~λ(𝝃~)ξ~22\displaystyle\frac{\tilde{G}_{\lambda}(\tilde{{\bm{\xi}}})}{\tilde{\xi}_{2}^{2}} =α¯(α1ud+ud+1η1)(α0ud+ud+1η0)\displaystyle=\bar{\alpha}\left(\alpha_{1}u^{d}+u^{d+1}\eta_{1}\right)-(\alpha_{0}u^{d}+u^{d+1}\eta_{0})
λ(α0ud+ud+1η0)(α1ud+ud+1η1)=ud+1γ\displaystyle-\lambda(\alpha_{0}u^{d}+u^{d+1}\eta_{0})(\alpha_{1}u^{d}+u^{d+1}\eta_{1})=u^{d+1}\gamma

where ν𝔓1(γ)0\nu_{{\mathfrak{P}}_{1}}(\gamma)\geq 0. This implies that ord𝔓1(G~λ(𝝃~))d+12{\rm ord}_{{\mathfrak{P}}_{1}}(\tilde{G}_{\lambda}(\tilde{{\bm{\xi}}}))\geq d+1\geq 2. Hence

div𝝃~(G~λ)𝔓1+i=1r𝔓i+Aλ.{\rm div}_{\tilde{{\bm{\xi}}}}(\tilde{G}_{\lambda})\geq{\mathfrak{P}}_{1}+\sum_{i=1}^{r}{\mathfrak{P}}_{i}+A_{\lambda}.

On the other hand, since deg(div𝝃~(G~λ))=2n\deg({\rm div}_{\tilde{{\bm{\xi}}}}(\tilde{G}_{\lambda}))=2n, one has that

div𝝃(G~λ)=𝔓1+i=1r𝔓i+Aλ.{\rm div}_{{\bm{\xi}}^{\prime}}(\tilde{G}_{\lambda})={\mathfrak{P}}_{1}+\sum_{i=1}^{r}{\mathfrak{P}}_{i}+A_{\lambda}.

This proves our claim. Now set Gλ=G~λ1G_{\lambda}=\tilde{G}_{\lambda}^{{\mathcal{L}}^{-1}}. As ξ~2=ξ2\tilde{\xi}_{2}=\xi_{2}, one sees that

minj{ν𝔓i(ξ~j)}=ν𝔓i(ξ~2)=ν𝔓i(ξ2)=minj{ν𝔓i(ξj)}.\min_{j}\{\nu_{{\mathfrak{P}}_{i}}(\tilde{\xi}_{j})\}=\nu_{{\mathfrak{P}}_{i}}(\tilde{\xi}_{2})=\nu_{{\mathfrak{P}}_{i}}(\xi_{2})=\min_{j}\{\nu_{{\mathfrak{P}}_{i}}(\xi_{j})\}.

This implies that

ord𝔓i(Gλ(𝝃))\displaystyle{\rm ord}_{{\mathfrak{P}}_{i}}(G_{\lambda}({\bm{\xi}})) =ν𝔓i(Gλ(𝝃))2mini{ν𝔓i(ξi)}\displaystyle=\nu_{{\mathfrak{P}}_{i}}\left(G_{\lambda}({\bm{\xi}})\right)-2\min_{i}\{\nu_{{\mathfrak{P}}_{i}}(\xi_{i})\}
=ν𝔓i(G~λ1(𝝃))2mini{ν𝔓i(ξ~i)}\displaystyle=\nu_{{\mathfrak{P}}_{i}}\left(\tilde{G}_{\lambda}^{{\mathcal{L}}^{-1}}({\bm{\xi}})\right)-2\min_{i}\{\nu_{{\mathfrak{P}}_{i}}(\tilde{\xi}_{i})\}
=ν𝔓i(G~λ(𝝃~))2mini{ν𝔓i(ξ~i)}=ord𝔓i(G~λ(𝝃~)).\displaystyle=\nu_{{\mathfrak{P}}_{i}}\left(\tilde{G}_{\lambda}(\tilde{{\bm{\xi}}})\right)-2\min_{i}\{\nu_{{\mathfrak{P}}_{i}}(\tilde{\xi}_{i})\}={\rm ord}_{{\mathfrak{P}}_{i}}(\tilde{G}_{\lambda}(\tilde{{\bm{\xi}}})).

Therefore div𝝃(Gλ)=𝔓1+i=1r𝔓i+A1{\rm div}_{\bm{\xi}}(G_{\lambda})={\mathfrak{P}}_{1}+\sum_{i=1}^{r}{\mathfrak{P}}_{i}+A_{1}. Note that we can choose λk\lambda\in k. For such λ\lambda, one has that

T(Gλ)T(G~λ)2T(𝐜)+T((α0,α1,0))2T(𝐜)+T(F~r).T(G_{\lambda})\leq T(\tilde{G}_{\lambda})\leq 2T({\bf c})+T((\alpha_{0},\alpha_{1},0))\leq 2T({\bf c})+T(\tilde{F}_{r}).

Since deg(F~r)=r2\deg(\tilde{F}_{r})=r\geq 2, one sees that T(F~r)T(F)+(n2)T(𝐜)T(\tilde{F}_{r})\leq T(F)+(n-2)T({\bf c}). This implies that T(Gλ)T(F)+nT(𝐜)T(G_{\lambda})\leq T(F)+nT({\bf c}) and

T(Aλ)2(2T(F)+nT(Gλ))(2n+4)T(F)+2n2T(𝐜).T(A_{\lambda})\leq 2(2T(F)+nT(G_{\lambda}))\leq(2n+4)T(F)+2n^{2}T({\bf c}).

Now applying 1.1. to the case that D=i=1r𝔓iD=\sum_{i=1}^{r}{\mathfrak{P}}_{i} and D=AλD^{\prime}=A_{\lambda}, one gets a linear homogeneous polynomial L1L_{1} such that div𝝃(L1)=i=1r𝔓i+C{\rm div}_{\bm{\xi}}(L_{1})=\sum_{i=1}^{r}{\mathfrak{P}}_{i}+C, where CC is a very simple divisor satisfying that supp(C){supp(Aλ){𝔓1,,𝔓r}}={\rm supp}(C)\cap\{{\rm supp}(A_{\lambda})\cup\{{\mathfrak{P}}_{1},\dots,{\mathfrak{P}}_{r}\}\}=\emptyset. Moreover T(L1)T𝝃(𝔓1)=T(𝐜)T(L_{1})\leq T_{\bm{\xi}}({\mathfrak{P}}_{1})=T({\bf c}). Let L2L_{2} be a linear homogeneous polynomial in k[x0,x1,x2]k[x_{0},x_{1},x_{2}] such that L2=0L_{2}=0 intersects F=0F=0 in nn distinct points other than the points in T𝝃(Aλ+C)T_{\bm{\xi}}(A_{\lambda}+C). For such L2L_{2}, one has that div𝝃(L2){\rm div}_{\bm{\xi}}(L_{2}) is a very simple divisor satisfying that supp(div𝝃(L2))supp(A1+C+D)={\rm supp}({\rm div}_{\bm{\xi}}(L_{2}))\cap{\rm supp}(A_{1}+C+D^{\prime})=\emptyset. Set H=L1L2H=L_{1}L_{2} and A=A1Cdiv𝝃(L2)A=A_{1}-C-{\rm div}_{\bm{\xi}}(L_{2}). Then T(H)T(𝐜)T(H)\leq T({\bf c}) and we obtain two polynomials G,HG,H as required. Note that T𝝃(C)2(T(F)+nT𝝃(D))T_{\bm{\xi}}(C)\leq 2(T(F)+nT_{\bm{\xi}}(D)) and T𝝃(div𝝃(L2))2T(F)T_{\bm{\xi}}({\rm div}_{\bm{\xi}}(L_{2}))\leq 2T(F). Hence T(Gλ),T(H)T(F)+nT(𝐜)andT𝝃(A)(2n+4)T(F)+2n2T(𝐜).T(G_{\lambda}),T(H)\leq T(F)+nT({\bf c})\,\,\mbox{and}\,\,T_{\bm{\xi}}(A)\leq(2n+4)T(F)+2n^{2}T({\bf c}).

Definition 3.21.

Suppose that FF has only ordinary singularities, say 𝐪1,,𝐪{\bf q}_{1},\dots,{\bf q}_{\ell}, and rir_{i} is the multiplicity of 𝐪i{\bf q}_{i}. Suppose further that for each i=1,,i=1,\dots,\ell, 𝔔i,1,,𝔔i,ri{\mathfrak{Q}}_{i,1},\dots,{\mathfrak{Q}}_{i,r_{i}} are all places of {\mathcal{R}} with 𝐪i{\bf q}_{i} as center with respect to 𝛏{\bm{\xi}}. Set

E𝝃=i=1(ri1)j=1ri𝔔i,j.E_{\bm{\xi}}=\sum_{i=1}^{\ell}(r_{i}-1)\sum_{j=1}^{r_{i}}{\mathfrak{Q}}_{i,j}.

A homogeneous polynomial GG such that div𝛏(G)E𝛏{\rm div}_{\bm{\xi}}(G)\geq E_{\bm{\xi}} is called an adjoint of FF.

We have the following two corollaries of Proposition 3.20.

Corollary 3.22.

Suppose that DD is a simple and effective divisor in {\mathcal{R}}. Let DD^{\prime} be a divisor in {\mathcal{R}}. Then there is a homogeneous polynomial GG of degree not greater than deg(D)+(n1)2/2\deg(D)+(n-1)^{2}/2 such that

div𝝃(G)=D+E𝝃+A{\rm div}_{\bm{\xi}}(G)=D+E_{\bm{\xi}}+A

where AA is a very simple and effective divisor of degree not greater than

deg(D)(n1)+n(n1)2/2\deg(D)(n-1)+n(n-1)^{2}/2

such that supp(A)(supp(D+E𝛏)supp(D))={\rm supp}(A)\cap({\rm supp}(D+E_{\bm{\xi}})\cup{\rm supp}(D^{\prime}))=\emptyset and T𝛏(A)2(T(F)+nT𝛏(D+E𝛏))T_{\bm{\xi}}(A)\leq 2(T(F)+nT_{\bm{\xi}}(D+E_{\bm{\xi}})). Moreover

T(G)(deg(D)+(n1)2/2)T𝝃(D+E𝝃).T(G)\leq\left(\deg(D)+(n-1)^{2}/2\right)T_{\bm{\xi}}(D+E_{\bm{\xi}}).
Proof.

Denote μ=deg(D)+i=1(ri1)\mu=\deg(D)+\sum_{i=1}^{\ell}(r_{i}-1) where rir_{i} is given as in Definition 3.21. Note that i=1(ri1)(n1)2/2\sum_{i=1}^{\ell}(r_{i}-1)\leq(n-1)^{2}/2, μdeg(D)+(n1)2/2\mu\leq\deg(D)+(n-1)^{2}/2. Write

D+E𝝃=i=1deg(D)𝔓i+s=deg(D)+1μDsD+E_{\bm{\xi}}=\sum_{i=1}^{\deg(D)}{\mathfrak{P}}_{i}+\sum_{s=\deg(D)+1}^{\mu}D_{s}

where the centers of 𝔓i{\mathfrak{P}}_{i} is a simple point of F=0F=0, Ds=j=1ri𝔔i,jD_{s}=\sum_{j=1}^{r_{i}}{\mathfrak{Q}}_{i,j} for some 1i1\leq i\leq\ell. Applying succesively Proposition 3.20 to 𝔓i{\mathfrak{P}}_{i} and DsD_{s}, one obtains μ\mu linear homogeneous polynomials L1,,LμL_{1},\dots,L_{\mu} such that div𝝃(Li)=𝔓i+Ai{\rm div}_{\bm{\xi}}(L_{i})={\mathfrak{P}}_{i}+A_{i} if imi\leq m, or div𝝃(Li)=Di+Ai{\rm div}_{\bm{\xi}}(L_{i})=D_{i}+A_{i} if i>mi>m, where AiA_{i} is a very simple and effective divisor such that

supp(Ai)(supp(D)supp(D+E𝝃+A1++Ai1))=.{\rm supp}(A_{i})\cap({\rm supp}(D^{\prime})\cup{\rm supp}(D+E_{\bm{\xi}}+A_{1}+\dots+A_{i-1}))=\emptyset.

Set G=i=1μLiG=\sum_{i=1}^{\mu}L_{i} and A=i=1μAiA=\sum_{i=1}^{\mu}A_{i}. Then one has that

div𝝃(G)=D+E𝝃+A.{\rm div}_{\bm{\xi}}(G)=D+E_{\bm{\xi}}+A.

Moreover by Proposition 3.20, T(Li)T𝝃(D+E𝝃)T(L_{i})\leq T_{\bm{\xi}}(D+E_{\bm{\xi}}) for all i=1,,μi=1,\dots,\mu and then Proposition 2.10 implies that T(G)μT𝝃(D+E𝝃)T(G)\leq\mu T_{\bm{\xi}}(D+E_{\bm{\xi}}). It is obvious that T𝝃(A)T_{\bm{\xi}}(A) is not greater than 2(T(F)+nT𝝃(D+E𝝃))2(T(F)+nT_{\bm{\xi}}(D+E_{\bm{\xi}})) because so is T𝝃(Ai)T_{\bm{\xi}}(A_{i}) for all i=1,,μi=1,\dots,\mu. ∎

Corollary 3.23.

Suppose that D,DD,D^{\prime} are two divisors in {\mathcal{R}}. Then there are two homogeneous polynomials G,HG,H of the same degree 2deg(D++D)\leq 2\deg(D^{+}+D^{-}) such that D^=div𝛏(G/H)+D\hat{D}={\rm div}_{\bm{\xi}}(G/H)+D is very simple and supp(div𝛏(G/H)+D)supp(D)={\rm supp}({\rm div}_{\bm{\xi}}(G/H)+D)\cap{\rm supp}(D^{\prime})=\emptyset. Moreover deg(D^+),deg(D^)2n(deg(D++D))\deg(\hat{D}^{+}),\deg(\hat{D}^{-})\leq 2n(\deg(D^{+}+D^{-})) and

T𝝃(div𝝃(G/H)+D)\displaystyle T_{\bm{\xi}}({\rm div}_{\bm{\xi}}(G/H)+D) (2n+4)T(F)+2n2T𝝃(D)\displaystyle\leq(2n+4)T(F)+2n^{2}T_{\bm{\xi}}(D)
T(G),T(H)\displaystyle T(G),T(H) deg(D++D)(T(F)+nT𝝃(D)),\displaystyle\leq\deg(D^{+}+D^{-})(T(F)+nT_{\bm{\xi}}(D)),

where n=deg(F)n=\deg(F).

Proof.

We first show the case that D-D is effective. Denote μ=deg(D)\mu=\deg(-D) and write

D=i=1s𝔓i+i=s+1μ𝔔i-D=\sum_{i=1}^{s}{\mathfrak{P}}_{i}+\sum_{i=s+1}^{\mu}{\mathfrak{Q}}_{i}

where the center of 𝔓i{\mathfrak{P}}_{i} (resp. 𝔔j{\mathfrak{Q}}_{j}) with respect to 𝝃{\bm{\xi}} is a simple (resp. singular) point of F=0F=0. Applying Proposition 3.20 to i=1s𝔓i\sum_{i=1}^{s}{\mathfrak{P}}_{i} yields a homogenenous polynomial G0G_{0} of degree ss such that div𝝃(G0)=i=1s𝔓i+A0{\rm div}_{\bm{\xi}}(G_{0})=\sum_{i=1}^{s}{\mathfrak{P}}_{i}+A_{0} where A0A_{0} is a very simple and effective divisor such that supp(A0)(supp(D)supp(D))={\rm supp}(A_{0})\cap({\rm supp}(D)\cup{\rm supp}(D^{\prime}))=\emptyset. Moreover deg(A0)=nsdeg(i=1s𝔓i)\deg(A_{0})=ns-\deg(\sum_{i=1}^{s}{\mathfrak{P}}_{i}). Construct ss linear homogeneous polynomials L1,,LsL_{1},\dots,L_{s} in k[x0,x1,x2]k[x_{0},x_{1},x_{2}] such that div𝝃(L1Ls){\rm div}_{\bm{\xi}}(L_{1}\cdots L_{s}) is very simple and supp(div𝝃(L1Ls))(supp(div𝝃(G0)supp(D))={\rm supp}({\rm div}_{\bm{\xi}}(L_{1}\cdots L_{s}))\cap({\rm supp}({\rm div}_{\bm{\xi}}(G_{0})\cup{\rm supp}(D^{\prime}))=\emptyset. It is easy to see that deg(div𝝃(L1Ls))=ns\deg({\rm div}_{\bm{\xi}}(L_{1}\cdots L_{s}))=ns. Set H0=L1LsH_{0}=L_{1}\cdots L_{s}. By Proposition 3.20 again, one obtains μs\mu-s pairs (G1,H1),,(Gμs,Hμs)(G_{1},H_{1}),\dots,(G_{\mu-s},H_{\mu-s}) of homogeneous polynomials of degree two such that div𝝃(Gi/Hi)=𝔔i+Ai{\rm div}_{\bm{\xi}}(G_{i}/H_{i})={\mathfrak{Q}}_{i}+A_{i} where AiA_{i} is a very simple divisor such that

supp(Ai)(supp(D)supp(D+A0++Ai1))={\rm supp}(A_{i})\cap({\rm supp}(D^{\prime})\cup{\rm supp}(D+A_{0}+\cdots+A_{i-1}))=\emptyset

and deg(Ai+)=2ndeg(𝔔i)\deg(A_{i}^{+})=2n-\deg({\mathfrak{Q}}_{i}), deg(Ai)=2n\deg(A_{i}^{-})=2n. Set G~=G0G1Gμs\tilde{G}=G_{0}G_{1}\cdots G_{\mu-s} and H~=H0H1Hμs\tilde{H}=H_{0}H_{1}\cdots H_{\mu-s}. Then

D^=div𝝃(G~/H~)+D=A0+A1++Aμs\hat{D}={\rm div}_{\bm{\xi}}(\tilde{G}/\tilde{H})+D=A_{0}+A_{1}+\dots+A_{\mu-s}

which is very simple. It is clear that deg(G~)=deg(H~)2deg(D)\deg(\tilde{G})=\deg(\tilde{H})\leq 2\deg(-D), and by Proposition 3.20

T(G~)T(G0)+(μs)T(Gi)μ(T(F)+nT𝝃(D)).T(\tilde{G})\leq T(G_{0})+(\mu-s)T(G_{i})\leq\mu(T(F)+nT_{\bm{\xi}}(D)).

Similarly, T(H~)μ(T(F)+nT𝝃(D))T(\tilde{H})\leq\mu(T(F)+nT_{\bm{\xi}}(D)). Furthermore, one has that

T𝝃(div𝝃(G~/H~)+D)(2n+4)T(F)+2n2T𝝃(D)T_{\bm{\xi}}\left({\rm div}_{\bm{\xi}}(\tilde{G}/\tilde{H})+D\right)\leq(2n+4)T(F)+2n^{2}T_{\bm{\xi}}(D)

and

deg(D^+)=(2μs)ndeg(D)2nμ,deg(D^)=(2μs)n2nμ.\deg(\hat{D}^{+})=(2\mu-s)n-\deg(-D)\leq 2n\mu,\,\deg(\hat{D}^{-})=(2\mu-s)n\leq 2n\mu.

For the general case, write D=D+DD=D^{+}-D^{-}. The previous discussion implies that we can obtain G~i,H~i\tilde{G}_{i},\tilde{H}_{i} such that div𝝃(G~1/H1~)D+{\rm div}_{\bm{\xi}}(\tilde{G}_{1}/\tilde{H_{1}})-D^{+} and div𝝃(G~2/H2~)D{\rm div}_{\bm{\xi}}(\tilde{G}_{2}/\tilde{H_{2}})-D^{-} are very simple. Moreover

supp(div𝝃(G~1/H1~)D+)supp(div𝝃(G~2/H2~)D)\displaystyle{\rm supp}\left({\rm div}_{\bm{\xi}}(\tilde{G}_{1}/\tilde{H_{1}})-D^{+}\right)\bigcap{\rm supp}\left({\rm div}_{\bm{\xi}}(\tilde{G}_{2}/\tilde{H_{2}})-D^{-}\right) =,\displaystyle=\emptyset,
supp(div𝝃(G~2/H2~)D+div𝝃(G~1/H1~)D+)(supp(D)supp(D))\displaystyle{\rm supp}\left({\rm div}_{\bm{\xi}}(\tilde{G}_{2}/\tilde{H_{2}})-D^{-}+{\rm div}_{\bm{\xi}}(\tilde{G}_{1}/\tilde{H_{1}})-D^{+}\right)\bigcap\left({\rm supp}(D)\cup{\rm supp}(D^{\prime})\right) =.\displaystyle=\emptyset.

Set G=G~2H~1G=\tilde{G}_{2}\tilde{H}_{1} and H=G~1H~2H=\tilde{G}_{1}\tilde{H}_{2}. Then div𝝃(G/H)+D{\rm div}_{\bm{\xi}}(G/H)+D satisfies the required condition. Furthermore deg(D^+),deg(D^)2n(deg(D++D))\deg(\hat{D}^{+}),\deg(\hat{D}^{-})\leq 2n(\deg(D^{+}+D^{-})) and

T(G),T(H)\displaystyle T(G),T(H) deg(D++D)(T(F)+nT𝝃(D)),\displaystyle\leq\deg(D^{+}+D^{-})(T(F)+nT_{\bm{\xi}}(D)),
T𝝃(div𝝃(G/H)+D)\displaystyle T_{\bm{\xi}}({\rm div}_{\bm{\xi}}(G/H)+D) (2n+4)T(F)+2n2T𝝃(D).\displaystyle\leq(2n+4)T(F)+2n^{2}T_{\bm{\xi}}(D).

Now we are ready to pove the main results of this section. Let us start with two lemmas.

Lemma 3.24.

Suppose that F=0F=0 has only ordinary singularities and DD is a divisor in {\mathcal{R}}. Let HH be a homogeneous polynomial in k(t)¯[x0,x1,x2]{\overline{k(t)}}[x_{0},x_{1},x_{2}] such that div𝛏(H)=D++E𝛏+A{\rm div}_{\bm{\xi}}(H)=D^{+}+E_{\bm{\xi}}+A, where AA is an effective divisor. Then

𝔏(D)={G(𝝃)H(𝝃)|G are homogeneous polynomials of deg(H)with div𝝃(G)D+E𝝃+A}.{\mathfrak{L}}(D)=\left\{\left.\frac{G({\bm{\xi}})}{H({\bm{\xi}})}\,\right|\,\begin{array}[]{c}\mbox{$G$ are homogeneous polynomials of $\deg(H)$}\\ \mbox{with ${\rm div}_{\bm{\xi}}(G)\geq D^{-}+E_{\bm{\xi}}+A$}\end{array}\right\}. (3)
Proof.

Note that div(G(𝝃)/H(𝝃))=div𝝃(G)div𝝃(H){\rm div}(G({\bm{\xi}})/H({\bm{\xi}}))={\rm div}_{\bm{\xi}}(G)-{\rm div}_{\bm{\xi}}(H). It is obvious that the right hand side of (3) is a subspace of 𝔏(D){\mathfrak{L}}(D). Suppose that h𝔏(D){0}h\in{\mathfrak{L}}(D)\setminus\{0\}, i.e. D=div(h)+DD^{\prime}={\rm div}(h)+D is effective. Then

div𝝃(H)+div(h)=D+D+E𝝃+A.{\rm div}_{\bm{\xi}}(H)+{\rm div}(h)=D^{-}+D^{\prime}+E_{\bm{\xi}}+A.

By the Residuce Theorem (see page of [9]), there is a homogeneous polynomial GG of degree deg(H)\deg(H) such that

div𝝃(G)=D+D+E𝝃+AD+E+A.{\rm div}_{\bm{\xi}}(G)=D^{-}+D^{\prime}+E_{\bm{\xi}}+A\geq D^{-}+E+A.

One sees that

div(hH(𝝃)/G(𝝃))=div(h)+div𝝃(H/G)=div(h)+div𝝃(H)div𝝃(G)=0.{\rm div}(hH({\bm{\xi}})/G({\bm{\xi}}))={\rm div}(h)+{\rm div}_{\bm{\xi}}(H/G)={\rm div}(h)+{\rm div}_{\bm{\xi}}(H)-{\rm div}_{\bm{\xi}}(G)=0.

Thus hH(𝝃)/G(𝝃)k(t)¯hH({\bm{\xi}})/G({\bm{\xi}})\in{\overline{k(t)}}, i.e. hh belongs to the right hand side of (3). ∎

Lemma 3.25.

Assume that M=(ai,j)M=(a_{i,j}) is an l×ml\times m matrix with ai,jk(t)¯a_{i,j}\in{\overline{k(t)}}. Assume further that for each place 𝔭{\mathfrak{p}} of k(t,a1,1,,al,m)k(t,a_{1,1},\dots,a_{l,m}), ν𝔭(ai,j)m𝔭-\nu_{\mathfrak{p}}(a_{i,j})\leq m_{\mathfrak{p}} where m𝔭0m_{\mathfrak{p}}\geq 0 and for all but finite number of 𝔭{\mathfrak{p}}, m𝔭0m_{\mathfrak{p}}\neq 0. Then there is a basis BB of the solution space of MY=0MY=0 satisfying that

T(𝐛)min{l,m}𝔭m𝔭[k(t,a1,1,,al,m):k(t)]T({\bf b})\leq\frac{\min\{l,m\}\sum_{{\mathfrak{p}}}m_{\mathfrak{p}}}{[k(t,a_{1,1},\dots,a_{l,m}):k(t)]}

for all 𝐛B{\bf b}\in B.

Proof.

Assume that r=rank(M)r={\rm rank}(M). Then rmin{l,m}r\leq\min\{l,m\}. Without loss of generality, we may assume the first rr-rows of MM are linearly independent and denote by M~\tilde{M} the matrix formed by them. Then the solution space of M~Y=0\tilde{M}Y=0 is the same as that of MY=0MY=0. Hence it suffices to consider the system M~Y=0\tilde{M}Y=0. We may further assume that the matrix M~1\tilde{M}_{1} formed by the first rr-columns of M~\tilde{M} is invertible. For every i=1,,ri=1,\dots,r and j=r+1,,mj=r+1,\dots,m, set di,jd_{i,j} to be the determinant of the matrix obtained from M~1\tilde{M}_{1} by replacing the ii-th column of M~1\tilde{M}_{1} by the jj-th column of M~\tilde{M}. For each j=r+1,,mj=r+1,\cdots,m, denote

𝐜j=(d1,j,,dr,j,0,,0,det(M1)j,0,,0)t{\bf c}_{j}=(d_{1,j},\dots,d_{r,j},0,\dots,0,\underbrace{\det(M_{1})}_{j},0,\dots,0)^{t}

where ()t(\cdot)^{t} denotes the transpose of a vector. Then by Cramer’s rule, the 𝐜j{\bf c}_{j} are solutions of M~Y=0\tilde{M}Y=0 and thus they form a basis of the solution space of M~Y=0\tilde{M}Y=0. Note that di,jd_{i,j} as well as det(M~1)\det(\tilde{M}_{1}) is an integer combination of the monomials in the entries of M~\tilde{M} of total degree rr. So for all i=1,,ri=1,\dots,r and j=r+1,,mj=r+1,\dots,m,

ν𝔭(M~1),ν𝔭(di,j)rm𝔭min{l,m}m𝔭-\nu_{\mathfrak{p}}(\tilde{M}_{1}),-\nu_{\mathfrak{p}}(d_{i,j})\leq rm_{\mathfrak{p}}\leq\min\{l,m\}m_{\mathfrak{p}}

where 𝔭{\mathfrak{p}} is a place of k(t,a1,1,,al,m)k(t,a_{1,1},\dots,a_{l,m}). This together Remark 2.7 implies the lemma. ∎

Theorem 3.26.

Suppose that F=0F=0 has only ordinary singularities. Let DD be a divisor in {\mathcal{R}}. Denote μ=deg(D++D)\mu=\deg(D^{+}+D^{-}) and N=max{T𝛏(D),T(F)}.N=\max\{T_{\bm{\xi}}(D),T(F)\}. Then there is a k(t)¯{\overline{k(t)}}-basis BB of 𝔏(D){\mathfrak{L}}(D) such that every element of BB can be represented by G(𝛏)/H(𝛏)G({\bm{\xi}})/H({\bm{\xi}}) where G,HG,H are two homogeneous polynomials of the same degree not greater than 2(n+1)μ+(n1)2/2\leq 2(n+1)\mu+(n-1)^{2}/2 and

T(G),T(H)4n5(n+1)3(2μ+(n1)/2)3N.T(G),T(H)\leq 4n^{5}(n+1)^{3}(2\mu+(n-1)/2)^{3}N.
Proof.

By Corollary 3.23, there are two homogeneous polynomials G1,H1G_{1},H_{1} of the same degree 2μ\leq 2\mu such that D^=div𝝃(G1/H1)+D\hat{D}={\rm div}_{\bm{\xi}}(G_{1}/H_{1})+D is very simple. Moreover

T(G1),T(H1)\displaystyle T(G_{1}),T(H_{1}) μ(n+1)N,\displaystyle\leq\mu(n+1)N,
T𝝃(D^)(2n+4)T(F)+2n2T𝝃(D)\displaystyle T_{\bm{\xi}}(\hat{D})\leq(2n+4)T(F)+2n^{2}T_{\bm{\xi}}(D) (2n2+2n+4)N,\displaystyle\leq(2n^{2}+2n+4)N,
deg(D^+),deg(D^)\displaystyle\deg(\hat{D}^{+}),\deg(\hat{D}^{-}) 2nμ.\displaystyle\leq 2n\mu.

Due to Corollary 3.22, there is a homogeneous polynomial G2G_{2} of degree not greater than 2nμ+(n1)2/22n\mu+(n-1)^{2}/2 such that div𝝃(G2)=D^++E𝝃+A{\rm div}_{\bm{\xi}}(G_{2})=\hat{D}^{+}+E_{\bm{\xi}}+A, where AA is a very simple and effective divisor and supp(A)supp(D^)={\rm supp}(A)\cap{\rm supp}(\hat{D}^{-})=\emptyset. Moreover

T(G2)(2nμ+(n1)2/2)T𝝃(D^++E𝝃)(2nμ+(n1)2/2)(2n2+2n+4)NT(G_{2})\leq(2n\mu+(n-1)^{2}/2)T_{\bm{\xi}}(\hat{D}^{+}+E_{\bm{\xi}})\leq(2n\mu+(n-1)^{2}/2)(2n^{2}+2n+4)N

and

T𝝃(A)2(T(F)+nT𝝃(D^++E𝝃))2(2n3+2n2+4n+1)N.T_{\bm{\xi}}(A)\leq 2(T(F)+nT_{\bm{\xi}}(\hat{D}^{+}+E_{\bm{\xi}}))\leq 2(2n^{3}+2n^{2}+4n+1)N.

Denote d=deg(G2)d=\deg(G_{2}). By Lemma 3.24, to compute 𝔏(D){\mathfrak{L}}(D), it suffices to compute all homogeneous polynomials H2H_{2} of degree dd satisfying that

div𝔓(H2)D^+E𝝃+A.{\rm div}_{\mathfrak{P}}(H_{2})\geq\hat{D}^{-}+E_{\bm{\xi}}+A.

Assume that

H2=i=0dj=0dici,jx0ix1jx2dijH_{2}=\sum_{i=0}^{d}\sum_{j=0}^{d-i}c_{i,j}x_{0}^{i}x_{1}^{j}x_{2}^{d-i-j}

where ci,jc_{i,j} are indeterminates. There are (d+1)(d+2)/2(d+1)(d+2)/2 indeterminates in total. For each 𝔓supp(D^+A){\mathfrak{P}}\in{\rm supp}(\hat{D}^{-}+A), div𝝃(H2)𝔓{\rm div}_{\bm{\xi}}(H_{2})\geq{\mathfrak{P}} if and only if the center of 𝔓{\mathfrak{P}} with respect to 𝝃{\bm{\xi}} is a zero of H2H_{2}. This imposes deg(D^+A)\deg(\hat{D}^{-}+A) linear constraints on H2H_{2}. At the same time, div𝝃(H2)(ri1)j=1ri𝔔i,j{\rm div}_{\bm{\xi}}(H_{2})\geq(r_{i}-1)\sum_{j=1}^{r_{i}}{\mathfrak{Q}}_{i,j} if and only if the center of 𝔔i,1{\mathfrak{Q}}_{i,1} with respect to 𝝃{\bm{\xi}} is a common zero of

j0+j1+j2(H2)x0j0x1j1x2j2\frac{\partial^{j_{0}+j_{1}+j_{2}}(H_{2})}{\partial x_{0}^{j_{0}}x_{1}^{j_{1}}x_{2}^{j_{2}}}

for all nonnegative integers j0,j1,j2j_{0},j_{1},j_{2} satisfying that j0+j1+j2=ri2j_{0}+j_{1}+j_{2}=r_{i}-2, where 𝔔i,j{\mathfrak{Q}}_{i,j} is as in Definition 3.21. This imposes ri(ri1)/2r_{i}(r_{i}-1)/2 linear constraints on H2H_{2}. So there are totally deg(D^+A)+deg(E𝝃)/2\deg(\hat{D}^{-}+A)+\deg(E_{\bm{\xi}})/2 linear constraints on H2H_{2}. The problem of finding H2H_{2} is reduced to that of solving the system MY=0MY=0, where YY is a vector with indeterminates entries and MM is a (deg(D^+A)+deg(E𝝃)/2)×(d+1)(d+2)/2(\deg(\hat{D}^{-}+A)+\deg(E_{\bm{\xi}})/2)\times(d+1)(d+2)/2 matrix. Denote by 𝐜𝔓=(c0,𝔓,c1,𝔓,c2,𝔓){\bf c}_{\mathfrak{P}}=(c_{0,{\mathfrak{P}}},c_{1,{\mathfrak{P}}},c_{2,{\mathfrak{P}}}) the center of 𝔓{\mathfrak{P}} in supp(D^+E𝝃+A){\rm supp}(\hat{D}^{-}+E_{\bm{\xi}}+A). Then the entries in the same row of MM are monomials of total degree d\leq d in c0,𝔓,c1,𝔓,c2,𝔓c_{0,{\mathfrak{P}}},c_{1,{\mathfrak{P}}},c_{2,{\mathfrak{P}}} for some 𝔓{\mathfrak{P}} in supp(D^+E𝝃+A){\rm supp}(\hat{D}^{-}+E_{\bm{\xi}}+A). Without loss of generality, we may assume that one of c0,𝔓,c1,𝔓,c2,𝔓c_{0,{\mathfrak{P}}},c_{1,{\mathfrak{P}}},c_{2,{\mathfrak{P}}} is 1. Let RR be a finite extension of k(t)k(t) containing all ci,𝔓c_{i,{\mathfrak{P}}}. For each place 𝔭{\mathfrak{p}} of RR, set

m𝔭=d𝔓supp(D^+E𝝃+A)max{ν𝔭(c0,𝔓),ν𝔭(c1,𝔓),ν𝔭(c2,𝔓)}.m_{\mathfrak{p}}=d\sum_{{\mathfrak{P}}\in{\rm supp}(\hat{D}^{-}+E_{\bm{\xi}}+A)}\max\{-\nu_{\mathfrak{p}}(c_{0,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{1,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{2,{\mathfrak{P}}})\}.

Since max{ν𝔭(c0,𝔓),ν𝔭(c1,𝔓),ν𝔭(c2,𝔓)}0\max\{-\nu_{\mathfrak{p}}(c_{0,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{1,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{2,{\mathfrak{P}}})\}\geq 0 for all 𝔓{\mathfrak{P}}, m𝔭0m_{\mathfrak{p}}\geq 0 and

ν𝔭(ai,j)dmax𝔓supp(D^+E𝝃+A)max{ν𝔭(c0,𝔓),ν𝔭(c1,𝔓),ν𝔭(c2,𝔓)}m𝔭\displaystyle-\nu_{\mathfrak{p}}(a_{i,j})\leq d\max_{{\mathfrak{P}}\in{\rm supp}(\hat{D}^{-}+E_{\bm{\xi}}+A)}\max\{-\nu_{\mathfrak{p}}(c_{0,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{1,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{2,{\mathfrak{P}}})\}\leq m_{\mathfrak{p}}

where M=(ai,j)M=(a_{i,j}). Note that

deg(D^+E𝝃+A)deg(div𝝃(H2))=nd.\deg(\hat{D}^{-}+E_{\bm{\xi}}+A)\leq\deg({\rm div}_{\bm{\xi}}(H_{2}))=nd.

Applying Lemma 3.25 to MM yields that

T(H2)\displaystyle T(H_{2}) deg(D^+E𝝃+A)𝔭m𝔭[R:k(t)]\displaystyle\leq\deg(\hat{D}^{-}+E_{\bm{\xi}}+A)\sum_{{\mathfrak{p}}}\frac{m_{\mathfrak{p}}}{[R:k(t)]}
nd2𝔓𝔭max{ν𝔭(c0,𝔓),ν𝔭(c1,𝔓),ν𝔭(c2,𝔓)}[R:k(t)]\displaystyle\leq nd^{2}\sum_{{\mathfrak{P}}}\sum_{{\mathfrak{p}}}\frac{\max\{-\nu_{\mathfrak{p}}(c_{0,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{1,{\mathfrak{P}}}),-\nu_{\mathfrak{p}}(c_{2,{\mathfrak{P}}})\}}{[R:k(t)]}
nd2𝔓T(𝐜𝔓)nd2deg(D^+E𝝃+A)max𝔓T(𝐜𝔓)\displaystyle\leq nd^{2}\sum_{{\mathfrak{P}}}T({\bf c}_{\mathfrak{P}})\leq nd^{2}\deg(\hat{D}^{-}+E_{\bm{\xi}}+A)\max_{{\mathfrak{P}}}T({\bf c}_{\mathfrak{P}})
n2d3T𝝃(D^+E𝝃+A)2n2d3(2n3+2n2+4n+1)N\displaystyle\leq n^{2}d^{3}T_{\bm{\xi}}(\hat{D}^{-}+E_{\bm{\xi}}+A)\leq 2n^{2}d^{3}(2n^{3}+2n^{2}+4n+1)N
2n5(2μ+(n1)/2)3(2n3+2n2+4n+1)N.\displaystyle\leq 2n^{5}(2\mu+(n-1)/2)^{3}(2n^{3}+2n^{2}+4n+1)N.

The last inequality holds because

d2nμ+(n1)2/2n(2μ+(n1)/2).d\leq 2n\mu+(n-1)^{2}/2\leq n(2\mu+(n-1)/2).

Set G=H2G1G=H_{2}G_{1} and H=G2H1H=G_{2}H_{1}. Then

deg(G)\displaystyle\deg(G) =deg(H)2(n+1)μ+(n1)2/2,\displaystyle=\deg(H)\leq 2(n+1)\mu+(n-1)^{2}/2,
T(G),T(H)\displaystyle T(G),T(H) T(H2)+T(G1)<T(H2)+μ(n+1)N\displaystyle\leq T(H_{2})+T(G_{1})<T(H_{2})+\mu(n+1)N
2n5(2μ+(n1)/2)3(2n3+2n2+4n+2)N\displaystyle\leq 2n^{5}(2\mu+(n-1)/2)^{3}(2n^{3}+2n^{2}+4n+2)N
4n5(n+1)3(2μ+(n1)/2)3N.\displaystyle\leq 4n^{5}(n+1)^{3}(2\mu+(n-1)/2)^{3}N.

Next, we consider the case that F=0F=0 may have non-ordinary singular points. Let 𝒞(𝝃){\mathcal{C}}({\bm{\xi}}) be a plane projective model of {\mathcal{R}}. Suppose that 𝒞(𝝃){\mathcal{C}}({\bm{\xi}}) is defined by F0F_{0} and for i=1,,si=1,\dots,s, FiF_{i} is the quadratic transformation of Fi1F_{i-1} under the quadratic transformation 𝒬𝐜i1{\mathcal{Q}}_{{\bf c}_{i-1}}, where 𝐜i1{\bf c}_{i-1} is a singular point of Fi1=0F_{i-1}=0. Denote n=deg(F0)n=\deg(F_{0}) and set 𝝃0=𝝃{\bm{\xi}}_{0}={\bm{\xi}}, 𝝃i+1=𝒬𝐜i1(𝝃i){\bm{\xi}}_{i+1}={\mathcal{Q}}_{{\bf c}_{i}}^{-1}({\bm{\xi}}_{i}) and

Ni=2i(i1)2nimax{8nT(F0),T𝝃0(D)}.N_{i}=2^{\frac{i(i-1)}{2}}n^{i}\max\{8nT(F_{0}),T_{{\bm{\xi}}_{0}}(D)\}. (4)
Proposition 3.27.

Let DD be a divisor in {\mathcal{R}} and the notations Fi,𝛏i,NiF_{i},{\bm{\xi}}_{i},N_{i} as above. One has that

  1. 1.

    tdeg(Fi)n2i2i+1+2{\rm tdeg}(F_{i})\leq n2^{i}-2^{i+1}+2;

  2. 2.

    T𝝃i(D),T(Fi)NiT_{{\bm{\xi}}_{i}}(D),T(F_{i})\leq N_{i}.

Proof.

1. Set ni=deg(Fi)n_{i}=\deg(F_{i}). Since every 𝐜i{\bf c}_{i} is a singular point, one has that ni2ni12n_{i}\leq 2n_{i-1}-2. This implies ni2in2i+1+2n_{i}\leq 2^{i}n-2^{i+1}+2.

2. Denote by SiS_{i} the maximum of the heights of singular points of Fi=0F_{i}=0. We first prove by induction on ii that T(Fi),SiNiT(F_{i}),S_{i}\leq N_{i} for all i=0,,si=0,\dots,s. Note that ni+1<2inn_{i}+1<2^{i}n for all i=1,,si=1,\dots,s. Since N0=8nT(F0)N_{0}=8nT(F_{0}), it is clear that T(F0)<N0T(F_{0})<N_{0} and S0<4nT(F0)<N0S_{0}<4nT(F_{0})<N_{0} by Corollary 3.5. Now assume that T(Fi),SiNiT(F_{i}),S_{i}\leq N_{i} for i=0i=\ell\geq 0. Consider the case i=+1i=\ell+1. By Corollary 3.13 and induction hypothesis, one has that

T(F+1)T(F)+nS(1+n)N<2nN=N+1.T(F_{\ell+1})\leq T(F_{\ell})+n_{\ell}S_{\ell}\leq(1+n_{\ell})N_{\ell}<2^{\ell}nN_{\ell}=N_{\ell+1}.

Note that n0=n>2n_{0}=n>2 as the curve F0=0F_{0}=0 has singularities. One sees that 4S4N<2nN=N+14S_{\ell}\leq 4N_{\ell}<2^{\ell}nN_{\ell}=N_{\ell+1} if >0\ell>0 and

4S0<16nT(F0)<8n2T(F0)N1.4S_{0}<16nT(F_{0})<8n^{2}T(F_{0})\leq N_{1}.

Consequently, 4Sj<Nj+14S_{j}<N_{j+1} for all j0j\geq 0. On the other hand, one has already seen that T(F)+nS<N+1.T(F_{\ell})+n_{\ell}S_{\ell}<N_{\ell+1}. By Corollary 3.13 again,

S+1max{4S,T(F)+nS}<N+1.\displaystyle S_{\ell+1}\leq\max\{4S_{\ell},T(F_{\ell})+n_{\ell}S_{\ell}\}<N_{\ell+1}.

For the divisor DD, it is obvious that T𝝃0(D)N0T_{{\bm{\xi}}_{0}}(D)\leq N_{0}. Suppose that T𝝃i(D)NiT_{{\bm{\xi}}_{i}}(D)\leq N_{i} for i=0i=\ell\geq 0. By Corollary 3.13 and the induction hypothesis,

T𝝃+1(D)max{2(S+N),T(F)+nT𝝃(D)}N+1.T_{{\bm{\xi}}_{\ell+1}}(D)\leq\max\{2(S_{\ell}+N_{\ell}),T(F_{\ell})+n_{\ell}T_{{\bm{\xi}}_{\ell}}(D)\}\leq N_{\ell+1}.

Notation 3.28.

Let FF be the defining polynomial of 𝒞(𝛏){\mathcal{C}}({\bm{\xi}}). Denote by s(𝛏)s({\bm{\xi}}) the number of quadratic transformations such that 𝒞(𝛏~){\mathcal{C}}(\tilde{{\bm{\xi}}}) has only ordinary singularities, where 𝛏~\tilde{{\bm{\xi}}} is the image of 𝛏{\bm{\xi}} under these quadratic transformations. By Theorem 2 in Chapter 7 of [9], s(𝛏)s({\bm{\xi}}) can be chosen to be an integer not greater than

m+(n1)(n2)2r𝐜(r𝐜1)2(n1)(n2)2m+\frac{(n-1)(n-2)}{2}-\sum\frac{r_{\bf c}(r_{\bf c}-1)}{2}\leq\frac{(n-1)(n-2)}{2}

where n=deg(F)n=\deg(F), mm is the number of non-ordinary singularities of F=0F=0, 𝐜{\bf c} ranges over all singularities of F=0F=0 and r𝐜r_{\bf c} is the multiplicity of 𝐜{\bf c}.

Theorem 3.29.

Let DD be a divisor in {\mathcal{R}}. Denote

n=deg(F),s=s(𝝃),μ=deg(D++D).n=\deg(F),s=s({\bm{\xi}}),\mu=\deg(D^{+}+D^{-}).

Then there is a k(t)¯{\overline{k(t)}}-basis BB of 𝔏(D){\mathfrak{L}}(D) such that every element of BB can be represented by G(𝛏)/H(𝛏)G({\bm{\xi}})/H({\bm{\xi}}) where G,HG,H are two homogeneous polynomials of the same degree not greater 22s+1(n+1)(μ+2s2n)2^{2s+1}(n+1)(\mu+2^{s-2}n) and

T(G),T(H)2s22+15s2+5ns+5(n+1)3(μ+2s2n)3max{8nT(F),T𝝃(D)}.T(G),T(H)\leq 2^{\frac{s^{2}}{2}+\frac{15s}{2}+5}n^{s+5}(n+1)^{3}(\mu+2^{s-2}n)^{3}\max\{8nT(F),T_{\bm{\xi}}(D)\}.
Proof.

If F=0F=0 has only ordinary singularities, i.e. s=0s=0, then the assertion is clear by Theorem 3.26. Suppose F=0F=0 has non-ordinary singularities, i.e. s1s\geq 1. Let 𝝃~\tilde{{\bm{\xi}}} be the image of 𝝃{\bm{\xi}} under ss quadratic transformations 𝒬𝐜01,,𝒬𝐜s11{\mathcal{Q}}^{-1}_{{\bf c}_{0}},\dots,{\mathcal{Q}}^{-1}_{{\bf c}_{s-1}} such that 𝒞(𝝃~){\mathcal{C}}(\tilde{{\bm{\xi}}}) has only ordinary singularities. Let F~\tilde{F} be the defining polynomial of 𝒞(𝝃~){\mathcal{C}}(\tilde{{\bm{\xi}}}) and set

κ=2s(s1)/2nsmax{8nT(F),T𝝃(D)}.\kappa=2^{s(s-1)/2}n^{s}\max\{8nT(F),T_{\bm{\xi}}(D)\}.

Then by Proposition 3.27

n~=deg(F~)2s(n2)+2andT𝝃~(D),T(F~)κ.\tilde{n}=\deg(\tilde{F})\leq 2^{s}(n-2)+2\,\,\mbox{and}\,\,T_{\tilde{{\bm{\xi}}}}(D),T(\tilde{F})\leq\kappa.

By Theorem 3.26, there is a k(t)¯{\overline{k(t)}}-basis BB of 𝔏(D){\mathfrak{L}}(D) satisfying that each element in BB can be represented by G~(𝝃~)/H~(𝝃~)\tilde{G}(\tilde{{\bm{\xi}}})/\tilde{H}(\tilde{{\bm{\xi}}}) where G~,H~\tilde{G},\tilde{H} are homogeneous polynomials of degree not greater than 2(n~+1)μ+(n~1)2/22(\tilde{n}+1)\mu+(\tilde{n}-1)^{2}/2 and

T(G~),T(H~)4n~5(n~+1)3(2μ+(n~1)/2)3κ.T(\tilde{G}),T(\tilde{H})\leq 4\tilde{n}^{5}(\tilde{n}+1)^{3}(2\mu+(\tilde{n}-1)/2)^{3}\kappa.

It remains to represent elements of BB in terms of 𝝃{\bm{\xi}}. We use the same notations as in the proof of Propositoin 3.27. Let 𝝃0=𝝃{\bm{\xi}}_{0}={\bm{\xi}} and 𝝃i=𝒬𝐜i11(𝝃i1){\bm{\xi}}_{i}={\mathcal{Q}}_{{\bf c}_{i-1}}^{-1}({\bm{\xi}}_{i-1}). Denote by nin_{i} the degree of the defining polynomial of 𝒞(𝝃i){\mathcal{C}}({\bm{\xi}}_{i}) and SiS_{i} the maximum of the heights of singular points of 𝒞(𝝃i){\mathcal{C}}({\bm{\xi}}_{i}). Let Gs=G~G_{s}=\tilde{G} and Gi1=Gi(𝒬𝐜i11((x0,x1,x2)))G_{i-1}=G_{i}({\mathcal{Q}}_{{\bf c}_{i-1}}^{-1}((x_{0},x_{1},x_{2}))) for all i=1,,s.i=1,\dots,s. One sees that deg(Gi)=2sideg(G~)\deg(G_{i})=2^{s-i}\deg(\tilde{G}). By Lemmas 3.6 and 3.10,

T(Gi1)T(G¯i)+deg(G¯i)T(𝐜i1)=T(Gi)+2deg(Gi)T(𝐜i1)T(G_{i-1})\leq T(\bar{G}_{i})+\deg(\bar{G}_{i})T({\bf c}_{i-1})=T(G_{i})+2\deg(G_{i})T({\bf c}_{i-1})

where G¯i=Gi(𝒬1((x0,x1,x2)))\bar{G}_{i}=G_{i}({\mathcal{Q}}^{-1}((x_{0},x_{1},x_{2}))). From the proof of Proposition 3.27, we have

T(G0)\displaystyle T(G_{0}) T(G~)+i=0s12deg(Gi+1)T(𝐜i)T(G~)+(i=0s12si)deg(G~)Ns1\displaystyle\leq T(\tilde{G})+\sum_{i=0}^{s-1}2\deg(G_{i+1})T({\bf c}_{i})\leq T(\tilde{G})+(\sum_{i=0}^{s-1}2^{s-i})\deg(\tilde{G})N_{s-1}
T(G~)+2s+1deg(G~)Ns1,\displaystyle\leq T(\tilde{G})+2^{s+1}\deg(\tilde{G})N_{s-1},

where Ns1N_{s-1} is given as in (4). Note that n~n2s\tilde{n}\leq n2^{s}. One has that

deg(G0)\displaystyle\deg(G_{0}) 2sdeg(G~)2s(2(n~+1)μ+(n~1)2/2)\displaystyle\leq 2^{s}\deg(\tilde{G})\leq 2^{s}(2(\tilde{n}+1)\mu+(\tilde{n}-1)^{2}/2)
2s(n~+1)(2μ+(n~1)/4)22s+1(n+1)(μ+n2s2);\displaystyle\leq 2^{s}(\tilde{n}+1)(2\mu+(\tilde{n}-1)/4)\leq 2^{2s+1}(n+1)(\mu+n2^{s-2});
T(G0)\displaystyle T(G_{0}) 4n~5(n~+1)3(2μ+n~12)3κ+2s+1(2(n~+1)μ+(n~1)22)Ns1\displaystyle\leq 4\tilde{n}^{5}(\tilde{n}+1)^{3}\left(2\mu+\frac{\tilde{n}-1}{2}\right)^{3}\kappa+2^{s+1}\left(2(\tilde{n}+1)\mu+\frac{(\tilde{n}-1)^{2}}{2}\right)N_{s-1}
4n~5(n~+1)3(2μ+n~12)3κ+2s+1(n~+1)(2μ+n~12)κ\displaystyle\leq 4\tilde{n}^{5}(\tilde{n}+1)^{3}\left(2\mu+\frac{\tilde{n}-1}{2}\right)^{3}\kappa+2^{s+1}(\tilde{n}+1)\left(2\mu+\frac{\tilde{n}-1}{2}\right)\kappa
(4n~5(2μ+n~12)+2s+1)(n~+1)3(2μ+n~12)2κ\displaystyle\leq\left(4\tilde{n}^{5}\left(2\mu+\frac{\tilde{n}-1}{2}\right)+2^{s+1}\right)(\tilde{n}+1)^{3}\left(2\mu+\frac{\tilde{n}-1}{2}\right)^{2}\kappa
(4n525s(2μ+n2s12)+2s+1)(n~+1)3(2μ+n~12)2κ\displaystyle\leq\left(4n^{5}2^{5s}\left(2\mu+\frac{n2^{s}-1}{2}\right)+2^{s+1}\right)(\tilde{n}+1)^{3}\left(2\mu+\frac{\tilde{n}-1}{2}\right)^{2}\kappa
8n525s(μ+n2s2)(n~+1)3(2μ+n~12)2κ\displaystyle\leq 8n^{5}2^{5s}(\mu+n2^{s-2})(\tilde{n}+1)^{3}\left(2\mu+\frac{\tilde{n}-1}{2}\right)^{2}\kappa
28s+5n5(n+1)3(μ+n2s2)3κ\displaystyle\leq 2^{8s+5}n^{5}(n+1)^{3}(\mu+n2^{s-2})^{3}\kappa
2s22+15s2+5ns+5(n+1)3(μ+2s2n)3max{8nT(F),T𝝃(D)}.\displaystyle\leq 2^{\frac{s^{2}}{2}+\frac{15s}{2}+5}n^{s+5}(n+1)^{3}(\mu+2^{s-2}n)^{3}\max\{8nT(F),T_{\bm{\xi}}(D)\}.

Similarly, we obtain bounds for deg(H0)\deg(H_{0}) and T(H0)T(H_{0}). ∎

4 Heights on plane algebraic curves

Let fk[x0,x1]f\in k[x_{0},x_{1}] be an irreducible polynomial over kk and a,bk(t){0}a,b\in k(t)\setminus\{0\} satisfy f(a,b)=0f(a,b)=0, i.e. (a,b)(a,b) is a rational parametrization of f=0f=0. The result on parametrization (see [16] for instance) implies that

deg(a)=mdeg(f,x1),deg(b)=mdeg(f,x0).\deg(a)=m\deg(f,x_{1}),\,\,\deg(b)=m\deg(f,x_{0}).

In other words, T(a)deg(f,x0)=T(b)deg(f,x1)T(a)\deg(f,x_{0})=T(b)\deg(f,x_{1}). A similar relation holds for points in algebraic curves defined over k(t)¯{\overline{k(t)}}, i.e there is a constant CC only depending on ff such that if (a,b)(a,b) is a point of f(x0,x1)=0f(x_{0},x_{1})=0 with coordinates in k(t)¯{\overline{k(t)}} then

deg(f,x0)T(a)Cdeg(f,x1)T(b)deg(f,x0)T(a)+C.\deg(f,x_{0})T(a)-C\leq\deg(f,x_{1})T(b)\leq\deg(f,x_{0})T(a)+C.

This is a special case of a general result for points in complete nonsingular varieties over a field with valuations. In the case of algebraic curves defined over k(t)¯{\overline{k(t)}}, Eremenko in 1999 presented another proof which actually provides a procedure to find CC explicitly. In this section, we shall present an explicit formula for CC following Eremenko’s proof.

4.1 Heights on plane projective curves

Throughout this subsection, ff is an irreducible polynomial in k(t)¯[x0,x1]{\overline{k(t)}}[x_{0},x_{1}] and {\mathcal{R}} is the algebraic function field over k(t)¯{\overline{k(t)}} associated to ff. Let us start with a refinement of Lemma 1 of [5].

Lemma 4.1.

Assume that fk(t)¯[x0,x1]f\in{\overline{k(t)}}[x_{0},x_{1}] is irreducible over k(t)¯{\overline{k(t)}} and α,βk(t)¯\alpha,\beta\in{\mathcal{R}}\setminus{\overline{k(t)}} satisfying f(α,β)=0f(\alpha,\beta)=0. If div(α)div(β){\rm div}(\alpha)^{-}\leq{\rm div}(\beta)^{-}, then for every place 𝔓{\mathfrak{P}} of {\mathcal{R}} with ν𝔓(β)0\nu_{{\mathfrak{P}}}(\beta)\geq 0, we have that

T(π𝔓(α))T(π𝔓(β))+T(f).T(\pi_{\mathfrak{P}}(\alpha))\leq T(\pi_{\mathfrak{P}}(\beta))+T(f).
Proof.

Since div(α)div(β){\rm div}(\alpha)^{-}\leq{\rm div}(\beta)^{-}, by Proposition 2 of [5], ff can be written to be of the form

f=x0n+an1(x1)x0n1++a1(x1)x0+a0(x1),f=x_{0}^{n}+a_{n-1}(x_{1})x_{0}^{n-1}+\dots+a_{1}(x_{1})x_{0}+a_{0}(x_{1}),

where aik(t)¯[x1]a_{i}\in{\overline{k(t)}}[x_{1}] with deg(ai)ni\deg(a_{i})\leq n-i. Write ai=j=0niai,jx1ja_{i}=\sum_{j=0}^{n-i}a_{i,j}x_{1}^{j} with ai,jk(t)¯a_{i,j}\in{\overline{k(t)}}. Let RR be a finite extension of k(t)k(t) containing all ai,ja_{i,j} and π𝔓(α),π𝔓(β)\pi_{\mathfrak{P}}(\alpha),\pi_{\mathfrak{P}}(\beta). Suppose that 𝔭{\mathfrak{p}} is a place of RR. Then

ν𝔭(π𝔓(αn))\displaystyle\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\alpha^{n})) =ν𝔭(i=0n1j=0niai,jπ𝔓(β)jπ𝔓(α)i)\displaystyle=\nu_{\mathfrak{p}}\left(-\sum_{i=0}^{n-1}\sum_{j=0}^{n-i}a_{i,j}\pi_{\mathfrak{P}}(\beta)^{j}\pi_{\mathfrak{P}}(\alpha)^{i}\right)
min0in1,0jni{ν𝔭(ai,j)+jν𝔭(π𝔓(β))+iν𝔭(π𝔓(α))}\displaystyle\geq\min_{0\leq i\leq n-1,0\leq j\leq n-i}\{\nu_{\mathfrak{p}}(a_{i,j})+j\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta))+i\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\alpha))\}
=ν𝔭(ai,j)+jν𝔭(π𝔓(β))+iν𝔭(π𝔓(α))\displaystyle=\nu_{\mathfrak{p}}(a_{i^{\prime},j^{\prime}})+j^{\prime}\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta))+i^{\prime}\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\alpha))

for some 0in1,0jni0\leq i^{\prime}\leq n-1,0\leq j^{\prime}\leq n-i^{\prime}. Equivalently,

ν𝔭(π𝔓(α))1niν𝔭(ai,j)+jniν𝔭(π𝔓(β)).\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\alpha))\geq\frac{1}{n-i^{\prime}}\nu_{\mathfrak{p}}(a_{i^{\prime},j^{\prime}})+\frac{j^{\prime}}{n-i^{\prime}}\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta)).

Therefore

max{0,ν𝔭(π𝔓(α))}\displaystyle\max\{0,-\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\alpha))\} max{0,ν𝔭(ai,j)nijν𝔭(π𝔓(β))ni}\displaystyle\leq\max\left\{0,-\frac{\nu_{\mathfrak{p}}(a_{i^{\prime},j^{\prime}})}{n-i^{\prime}}-\frac{j^{\prime}\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta))}{n-i^{\prime}}\right\}
max{0,ν𝔭(ai,j)}+max{0,ν𝔭(π𝔓(β))}\displaystyle\leq\max\left\{0,-\nu_{\mathfrak{p}}(a_{i^{\prime},j^{\prime}})\right\}+\max\left\{0,-\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta))\right\}
maxi,j{0,ν𝔭(ai,j)}+max{0,ν𝔭(π𝔓(β))}.\displaystyle\leq\max_{i,j}\left\{0,-\nu_{\mathfrak{p}}(a_{i,j})\right\}+\max\left\{0,-\nu_{\mathfrak{p}}(\pi_{\mathfrak{P}}(\beta))\right\}.

This implies that T(π𝔓(α))T(π𝔓(β))+T(f)T(\pi_{\mathfrak{P}}(\alpha))\leq T(\pi_{\mathfrak{P}}(\beta))+T(f). ∎

Lemma 4.2.

Let SS be a finite set of places in {\mathcal{R}} and α\alpha\in{\mathcal{R}}. Then there are a1,a2ka_{1},a_{2}\in k with a20a_{2}\neq 0 such that

supp(div(αa1α+a2))S=.{\rm supp}\left({\rm div}\left(\frac{\alpha}{a_{1}\alpha+a_{2}}\right)^{-}\right)\cap S=\emptyset.
Proof.

Set

M={π𝔓(α)𝔓S with ν𝔓(α)0}.M=\left\{\pi_{\mathfrak{P}}(\alpha)\mid\mbox{$\forall\,{\mathfrak{P}}\in S$ with $\nu_{\mathfrak{P}}(\alpha)\geq 0$}\right\}.

Then MM is a finite set in k(t)¯{\overline{k(t)}}. Let a1,a2ka_{1},a_{2}\in k satisfy that a20a_{2}\neq 0 and a1c+a20a_{1}c+a_{2}\neq 0 for all cMc\in M. For 𝔓S{\mathfrak{P}}\in S with ν𝔓(α)0\nu_{\mathfrak{P}}(\alpha)\geq 0, one has that

π𝔓(a1α+a2)=a1π𝔓(α)+a20,i.e.ν𝔓(a1α+a2)=0.\pi_{\mathfrak{P}}(a_{1}\alpha+a_{2})=a_{1}\pi_{\mathfrak{P}}(\alpha)+a_{2}\neq 0,\,{\mathrm{i}.e.}\,\,\nu_{\mathfrak{P}}(a_{1}\alpha+a_{2})=0.

This implies that ν𝔓(α/(a1α+a2))=ν𝔓(α)0\nu_{\mathfrak{P}}(\alpha/(a_{1}\alpha+a_{2}))=\nu_{\mathfrak{P}}(\alpha)\geq 0. On the other hand, for 𝔓S{\mathfrak{P}}\in S with ν𝔓(α)<0\nu_{\mathfrak{P}}(\alpha)<0, one has that

ν𝔓(α/(a1α+a2))=ν𝔓(α)ν𝔓(a1α+a2)=ν𝔓(α)ν𝔓(α)=0.\nu_{\mathfrak{P}}(\alpha/(a_{1}\alpha+a_{2}))=\nu_{\mathfrak{P}}(\alpha)-\nu_{\mathfrak{P}}(a_{1}\alpha+a_{2})=\nu_{\mathfrak{P}}(\alpha)-\nu_{\mathfrak{P}}(\alpha)=0.

In both cases, 𝔓{\mathfrak{P}} is not a pole of α/(a1α+a2)\alpha/(a_{1}\alpha+a_{2}). Thus a1,a2a_{1},a_{2} satisfy the requirement. ∎

The main result of this section is the following theorem which is a special case of Lemma 2 of [5]. The original proof of Lemma 2 of [5] contains a small gap. We shall fill in this gap in the proof.

Theorem 4.3.

Let ff be an irreducible polynomial in k(t)¯[x0,x1]{\overline{k(t)}}[x_{0},x_{1}] of degree n0n_{0} with respect to x0x_{0} and of degree n1n_{1} with respect to x1x_{1}. Suppose that n=tdeg(f)n={\rm tdeg}(f) and N1N\geq 1. Then for every c0,c1k(t)¯c_{0},c_{1}\in{\overline{k(t)}} satisfying f(c0,c1)=0f(c_{0},c_{1})=0, one has that

(1nN+n)n0T(c0)Cn1T(c1)(1+nN)n1T(c1)+C\left(1-\frac{n}{N+n}\right)n_{0}T(c_{0})-C\leq n_{1}T(c_{1})\leq\left(1+\frac{n}{N}\right)n_{1}T(c_{1})+C

where

C=2s2/2+15s/2+10(2Nn+n2+2s2)4ns+9(n+1)4T(f)/NC=2^{s^{2}/2+15s/2+10}(2Nn+n^{2}+2^{s-2})^{4}n^{s+9}(n+1)^{4}T(f)/N (5)

and ss is the number of quadratic transformations which are applied to resolve the singularities of f=0f=0.

Proof.

If one of cic_{i} is in kk then the height of the other one is not greater than T(f)T(f). The inequalities then obviously hold. In the following, we assume that neither c0c_{0} nor c1c_{1} is in kk.

Let {\mathcal{R}} be the algebraic function field associated to ff and α,βk(t)¯\alpha,\beta\in{\mathcal{R}}\setminus{\overline{k(t)}} satisfy that f(α,β)=0f(\alpha,\beta)=0. Choose a1,a2ka_{1},a_{2}\in k such that a20a_{2}\neq 0 and

supp(div(α/(a1α+a2)))div(β)=.{\rm supp}({\rm div}(\alpha/(a_{1}\alpha+a_{2}))^{-})\cap{\rm div}(\beta)^{-}=\emptyset.

Such a1,a2a_{1},a_{2} exist due to Lemma 4.2. Set α¯=α/(a1α+a2)\bar{\alpha}=\alpha/(a_{1}\alpha+a_{2}). Consider the divisor

D=(N+n)n1div(β)Nn0div(α¯).D=(N+n)n_{1}{\rm div}(\beta)^{-}-Nn_{0}{\rm div}(\bar{\alpha})^{-}.

Note that deg(div(α¯))=n1\deg({\rm div}(\bar{\alpha})^{-})=n_{1}, deg(div(β))=n0\deg({\rm div}(\beta)^{-})=n_{0} and

n0n1n0+n11n1.n_{0}n_{1}\geq n_{0}+n_{1}-1\geq n-1.

So

deg(D)=nn0n1n(n1).\deg(D)=nn_{0}n_{1}\geq n(n-1).

This implies that deg(D)\deg(D) is greater than the genus of f=0f=0 and thus 𝔏(D){0}{\mathfrak{L}}(D)\neq\{0\}. Denote 𝝃=(α,β,1){\bm{\xi}}=(\alpha,\beta,1) and by F(x0,x1,x2)F(x_{0},x_{1},x_{2}) the homogenization of ff . We claim that T𝝃(D)T(f)T_{\bm{\xi}}(D)\leq T(f). Note that T𝝃(D)=max{T𝝃(div(α¯)),T𝝃(div(β))}.T_{\bm{\xi}}(D)=\max\{T_{\bm{\xi}}({\rm div}(\bar{\alpha})^{-}),T_{\bm{\xi}}({\rm div}(\beta)^{-})\}. For each 𝐚𝒮𝝃(div(β)){\bf a}\in{\mathcal{S}}_{\bm{\xi}}({\rm div}(\beta)^{-}), 𝐚{\bf a} is of the form (b0,b1,0)(b_{0},b_{1},0) where b0,b1b_{0},b_{1} satisfies that F(b0,b1,0)=0F(b_{0},b_{1},0)=0. So T(𝐚)T(F)=T(f)T({\bf a})\leq T(F)=T(f) and then T𝝃(div(β))T(f)T_{\bm{\xi}}({\rm div}(\beta)^{-})\leq T(f). If a1=0a_{1}=0 then each point in 𝒮𝝃(div(α¯)){\mathcal{S}}_{\bm{\xi}}({\rm div}(\bar{\alpha})^{-}) is of the form (b0,b1,0)(b_{0},b_{1},0) too and so T𝝃(div(α¯))T(f)T_{\bm{\xi}}({\rm div}(\bar{\alpha})^{-})\leq T(f). Otherwise, for each 𝔓div(α¯){\mathfrak{P}}\in{\rm div}(\bar{\alpha})^{-}, one has that

ν𝔓(α)=ν𝔓(a2α¯/(1a1α¯))=ν𝔓(α¯)ν𝔓(1a1α¯)=0.\nu_{\mathfrak{P}}(\alpha)=\nu_{\mathfrak{P}}(a_{2}\bar{\alpha}/(1-a_{1}\bar{\alpha}))=\nu_{\mathfrak{P}}(\bar{\alpha})-\nu_{\mathfrak{P}}(1-a_{1}\bar{\alpha})=0.

Moreover π𝔓(α)=a2/a1\pi_{\mathfrak{P}}(\alpha)=-a_{2}/a_{1}. This implies that each point of 𝒮𝝃(div(α¯)){\mathcal{S}}_{\bm{\xi}}({\rm div}(\bar{\alpha})^{-}) is of the form (a2/a1,b,1)(-a_{2}/a_{1},b,1) whose height is not greater than T(f)T(f). Thus T𝝃(div(α¯))T(f)T_{\bm{\xi}}({\rm div}(\bar{\alpha})^{-})\leq T(f). Our claim is proved. Note that

deg(D++D)=2Nn0n1+nn0n1(2N+n)n2.\deg(D^{+}+D^{-})=2Nn_{0}n_{1}+nn_{0}n_{1}\leq(2N+n)n^{2}.

Suppose that γ𝔏(D){0}\gamma\in{\mathfrak{L}}(D)\setminus\{0\}. Due to Theorem 3.29, γ=G(𝝃)/H(𝝃)\gamma=G({\bm{\xi}})/H({\bm{\xi}}) where G,HG,H are two homogeneous polynomials of degree not greater than

22s+1(n+1)(deg(D++D)+2s2n)22s+1n(n+1)(2Nn+n2+2s2)2^{2s+1}(n+1)\left(\deg(D^{+}+D^{-})+2^{s-2}n\right)\leq 2^{2s+1}n(n+1)(2Nn+n^{2}+2^{s-2})

and

T(G),T(H)\displaystyle T(G),T(H) 2s2/2+15s/2+8ns+6(n+1)3(deg(D++D)+2s2n)3T(f)\displaystyle\leq 2^{s^{2}/2+15s/2+8}n^{s+6}(n+1)^{3}\left(\deg(D^{+}+D^{-})+2^{s-2}n\right)^{3}T(f)
2s2/2+15s/2+8ns+9(n+1)3(2Nn+n2+2s2)3T(f).\displaystyle\leq 2^{s^{2}/2+15s/2+8}n^{s+9}(n+1)^{3}\left(2Nn+n^{2}+2^{s-2}\right)^{3}T(f).

Set

C~=2s2/2+15s/2+9ns+9(n+1)4(2Nn+n2+2s2)4T(f).\tilde{C}=2^{s^{2}/2+15s/2+9}n^{s+9}(n+1)^{4}(2Nn+n^{2}+2^{s-2})^{4}T(f).

Without loss of generality, we assume that G(x0,x1,1)G(x_{0},x_{1},1) and H(x0,x1,1)H(x_{0},x_{1},1) have no common factor. Otherwise, by Corollary 2.15, we may replace GG and HH by G/WG/W and H/WH/W where WW is the greatest common factor of GG and HH. Moreover, multiplying by suitable elements in k(t)¯{\overline{k(t)}} if necessary, we can assume that both G(x0,x1,1)G(x_{0},x_{1},1) and H(x0,x1,1)H(x_{0},x_{1},1) have 1 as a coefficient. Let 𝔓{\mathfrak{P}} be a place of {\mathcal{R}} containing αc0\alpha-c_{0} and βc1\beta-c_{1}. Then ν𝔓(α)=0\nu_{\mathfrak{P}}(\alpha)=0 and ν𝔓(β)=0\nu_{\mathfrak{P}}(\beta)=0. As γ𝔏(D)\gamma\in{\mathfrak{L}}(D), ν𝔓(γ)0\nu_{\mathfrak{P}}(\gamma)\geq 0. If ν𝔓(γ)>0\nu_{\mathfrak{P}}(\gamma)>0, then ν𝔓(G(α,β,1))>0\nu_{\mathfrak{P}}(G(\alpha,\beta,1))>0 and so G(c0,c1,1)=0G(c_{0},c_{1},1)=0. Consequently, (c0,c1)(c_{0},c_{1}) is a common point of G(x0,x1,1)=0G(x_{0},x_{1},1)=0 and f(x0,x1)=0f(x_{0},x_{1})=0. Proposition 3.4 implies that T(ci)deg(G)T(f)+nT(G)T(c_{i})\leq\deg(G)T(f)+nT(G). It is easy to verify that in this case T(c0),T(c1)T(c_{0}),T(c_{1}) satisfy the required inequalities. Therefore we only need to prove the case ν𝔓(γ)=0\nu_{\mathfrak{P}}(\gamma)=0.

Set

h1(x1,y)\displaystyle h_{1}(x_{1},y) =resx0(f(x0,x1),H(x0,x1,1)yG(x0,x1,1)\displaystyle={\rm res}_{x_{0}}(f(x_{0},x_{1}),H(x_{0},x_{1},1)y-G(x_{0},x_{1},1)
h2(x2,y)\displaystyle h_{2}(x_{2},y) =resx1(h1(x1,y),x2x1(N+n)n1)\displaystyle={\rm res}_{x_{1}}(h_{1}(x_{1},y),x_{2}-x_{1}^{(N+n)n_{1}})

where resx0(f,g){\rm res}_{x_{0}}(f,g) denotes the resultant of ff and gg with respect to x0x_{0}. Note that h1(x1,y)0h_{1}(x_{1},y)\neq 0, because G(x0,x1,1)G(x_{0},x_{1},1) and H(x0,x1,1)H(x_{0},x_{1},1) have no common factor. As DD is not effective, γk(t)¯\gamma\notin{\overline{k(t)}}. Furthermore, as h1(x1,γ)=0h_{1}(x_{1},\gamma)=0, deg(h1,x1)>0\deg(h_{1},x_{1})>0. It is easy to see that h20h_{2}\neq 0 and h2(β(N+n)n1,γ)=0h_{2}(\beta^{(N+n)n_{1}},\gamma)=0. Let h~2\tilde{h}_{2} be an irreducible factor of h2h_{2} in k(t)¯[x2,y]{\overline{k(t)}}[x_{2},y] such that h~2(β(N+n)n1,γ)=0\tilde{h}_{2}(\beta^{(N+n)n_{1}},\gamma)=0. Propositions 2.16 and 2.10 imply that

T(h~2)\displaystyle T(\tilde{h}_{2}) T(h2)(N+n)n1T(h1)\displaystyle\leq T(h_{2})\leq(N+n)n_{1}T(h_{1})
(N+n)n1(deg(H)T(f)+n(T(G)+T(H)))\displaystyle\leq(N+n)n_{1}\left(\deg(H)T(f)+n(T(G)+T(H))\right)
(N+n)n1(22s+1n(n+1)(2Nn+n2+2s2)T(f)\displaystyle\leq(N+n)n_{1}(2^{2s+1}n(n+1)(2Nn+n^{2}+2^{s-2})T(f)
+2n2s2/2+15s/2+8ns+9(n+1)3(2Nn+n2+2s2)3T(f))\displaystyle+2n2^{s^{2}/2+15s/2+8}n^{s+9}(n+1)^{3}\left(2Nn+n^{2}+2^{s-2}\right)^{3}T(f))
(N+n)n12s2/2+15s/2+9ns+9(n+1)4(2Nn+n2+2s2)3T(f)\displaystyle\leq(N+n)n_{1}2^{s^{2}/2+15s/2+9}n^{s+9}(n+1)^{4}\left(2Nn+n^{2}+2^{s-2}\right)^{3}T(f)
2s2/2+15s/2+9ns+9(n+1)4(2Nn+n2+2s2)4T(f)=C~.\displaystyle\leq 2^{s^{2}/2+15s/2+9}n^{s+9}(n+1)^{4}\left(2Nn+n^{2}+2^{s-2}\right)^{4}T(f)=\tilde{C}.

Remark that

div(β(N+n)n1)=(N+n)n1div(β)D+div(γ).{\rm div}(\beta^{(N+n)n_{1}})^{-}=(N+n)n_{1}{\rm div}(\beta)^{-}\geq D^{+}\geq{\rm div}(\gamma)^{-}.

Note that π𝔓(β)=c1\pi_{\mathfrak{P}}(\beta)=c_{1}. By Lemma 4.1,

T(π𝔓(γ))\displaystyle T(\pi_{\mathfrak{P}}(\gamma)) T(π𝔓(β)(N+n)n1)+T(h~2)\displaystyle\leq T\left(\pi_{\mathfrak{P}}(\beta)^{(N+n)n_{1}}\right)+T(\tilde{h}_{2}) (6)
(N+n)n1T(c1)+C~.\displaystyle\leq(N+n)n_{1}T\left(c_{1}\right)+\tilde{C}.

Similarly, let r1(x0,y)=resx1(f(x0,x1),G(x0,x1,1)yH(x0,x1,1))r_{1}(x_{0},y)={\rm res}_{x_{1}}(f(x_{0},x_{1}),G(x_{0},x_{1},1)y-H(x_{0},x_{1},1)) and

r2(x2,y)=resx0(r1(x0,y),(a1x0+a2)Nn0x2x0Nn0).r_{2}(x_{2},y)={\rm res}_{x_{0}}\left(r_{1}(x_{0},y),(a_{1}x_{0}+a_{2})^{Nn_{0}}x_{2}-x_{0}^{Nn_{0}}\right).

Then r20r_{2}\neq 0 and r2(α¯Nn0,γ1)=0r_{2}(\bar{\alpha}^{Nn_{0}},\gamma^{-1})=0. Let r~2\tilde{r}_{2} be an irreducible factor of r2r_{2} in k(t)¯[x2,y]{\overline{k(t)}}[x_{2},y] such that r~2(α¯Nn0,γ1)=0\tilde{r}_{2}(\bar{\alpha}^{Nn_{0}},\gamma^{-1})=0. Applying Propositions 2.16 and 2.10 again yields that

T(r~)T(r2)\displaystyle T(\tilde{r})\leq T(r_{2}) Nn0T(r1)Nn0(deg(G)T(f)+n(T(H)+T(G))).\displaystyle\leq Nn_{0}T(r_{1})\leq Nn_{0}\left(\deg(G)T(f)+n(T(H)+T(G))\right).

An argument similar to the above implies that T(r~)C~T(\tilde{r})\leq\tilde{C}. Since supp(div(α¯))supp(div(β))={\rm supp}({\rm div}(\bar{\alpha})^{-})\cap{\rm supp}({\rm div}(\beta)^{-})=\emptyset, one has that Nn0δ(α¯)=DNn_{0}\delta(\bar{\alpha})^{-}=D^{-} and thus

div(α¯Nn0)=Nn0div(α¯)=Ddiv(γ)+=div(γ1).{\rm div}(\bar{\alpha}^{Nn_{0}})^{-}=Nn_{0}{\rm div}(\bar{\alpha})^{-}=D^{-}\leq{\rm div}(\gamma)^{+}={\rm div}(\gamma^{-1})^{-}.

Furthermore since a1c0+a20a_{1}c_{0}+a_{2}\neq 0, π𝔓(α¯)=c0/(a1c0+a2)\pi_{\mathfrak{P}}(\bar{\alpha})=c_{0}/(a_{1}c_{0}+a_{2}). By Lemma 4.1,

Nn0T(c0a1c0+a2)\displaystyle Nn_{0}T\left(\frac{c_{0}}{a_{1}c_{0}+a_{2}}\right) =T(π𝔓(α¯)Nn0)T(π𝔓(γ1))+T(r~2)\displaystyle=T(\pi_{\mathfrak{P}}(\bar{\alpha})^{Nn_{0}})\leq T(\pi_{\mathfrak{P}}(\gamma^{-1}))+T(\tilde{r}_{2})
=T(π𝔓(γ))+T(r~2)T(π𝔓(γ))+C~.\displaystyle=T(\pi_{\mathfrak{P}}(\gamma))+T(\tilde{r}_{2})\leq T(\pi_{\mathfrak{P}}(\gamma))+\tilde{C}.

which together with (6) gives

Nn0T(c0a1c0+a2)(N+n)n1T(c1)+2C~.Nn_{0}T\left(\frac{c_{0}}{a_{1}c_{0}+a_{2}}\right)\leq(N+n)n_{1}T\left(c_{1}\right)+2\tilde{C}.

Proposition 2.8 implies that

(1nN+n)n0T(c0a1c0+a2)2C~N+n\displaystyle\left(1-\frac{n}{N+n}\right)n_{0}T\left(\frac{c_{0}}{a_{1}c_{0}+a_{2}}\right)-\frac{2\tilde{C}}{N+n} (1nN+n)n0T(c0)2C~N+n\displaystyle\leq\left(1-\frac{n}{N+n}\right)n_{0}T\left(c_{0}\right)-\frac{2\tilde{C}}{N+n}
n1T(c1).\displaystyle\leq n_{1}T(c_{1}).

To prove the inequality in the opppsite direction, consider

D~=(N+n)n0div(α¯)Nn1div(β).\tilde{D}=(N+n)n_{0}{\rm div}(\bar{\alpha})^{-}-Nn_{1}{\rm div}(\beta)^{-}.

Remark that deg(D++D)=deg(D~++D~)\deg(D^{+}+D^{-})=\deg(\tilde{D}^{+}+\tilde{D}^{-}) and T𝝃(D)=T𝝃(D~)T_{\bm{\xi}}(D)=T_{\bm{\xi}}(\tilde{D}). We have the same bounds for elements in (D~){\mathcal{L}}(\tilde{D}). A similar argument then implies that

n1T(c1)N+nNn0T(c0)+2C~N(1+nN)n1T(c0)+2C~N.n_{1}T(c_{1})\leq\frac{N+n}{N}n_{0}T(c_{0})+\frac{2\tilde{C}}{N}\leq\left(1+\frac{n}{N}\right)n_{1}T(c_{0})+\frac{2\tilde{C}}{N}.

Set C=2C~/NC=2\tilde{C}/N. Then one gets the required inequalities. ∎

5 Main results

In this section, we always ssume that f(y,y)=i=0dai(y)yf(y,y^{\prime})=\sum_{i=0}^{d}a_{i}(y)y^{\prime} is irreducible over k(t)k(t) and

=m.s.index(f)=maxi=0d{deg(ai)2(di)}>0.\ell={\rm m.s.index}(f)=\max_{i=0}^{d}\{\deg(a_{i})-2(d-i)\}>0.

Pick ckc\in k such that a0(c)0a_{0}(c)\neq 0. Set y=(cz+1)/zy=(cz+1)/z. Then y=z/z2y^{\prime}=-z^{\prime}/z^{2}. Set

bi(z)=ai((cz+1)/z)z+2d2i(1)ib_{i}(z)=a_{i}((cz+1)/z)z^{\ell+2d-2i}(-1)^{i}

where i=0,,di=0,\dots,d. Then an easy calculation yields that

g(z,z)=i=0dbi(z)zi=z2d+f(cz+1z,zz2).\displaystyle g(z,z^{\prime})=\sum_{i=0}^{d}b_{i}(z)z^{\prime i}=z^{2d+\ell}f\left(\frac{cz+1}{z},\frac{-z^{\prime}}{z^{2}}\right).

As a0(c)0a_{0}(c)\neq 0, deg(a0(cz+1)/z)zdeg(a0))=deg(a0).\deg(a_{0}(cz+1)/z)z^{\deg(a_{0})})=\deg(a_{0}). This implies that

deg(b0)=+2d>2d.\deg(b_{0})=\ell+2d>2d.

Then tdeg(g)=2d+{\rm tdeg}(g)=2d+\ell because

2d+tdeg(g)=max{deg(bi)+i}max{2d+2i}=2d+.2d+\ell\leq{\rm tdeg}(g)=\max\{\deg(b_{i})+i\}\leq\max\{2d+\ell-2i\}=2d+\ell.

We claim that g(z,z)g(z,z^{\prime}) is irreducible over k(t)k(t). First of all, assume that =deg(ai0)2(di0)\ell=\deg(a_{i_{0}})-2(d-i_{0}) for some 0i0d0\leq i_{0}\leq d. Then we have that

bi0(0)=(1)i0the leading coefficient of ai0(y)0.b_{i_{0}}(0)=(-1)^{i_{0}}\cdot\mbox{the leading coefficient of $a_{i_{0}}(y)$}\neq 0.

If gcd(b0,,bd)1\gcd(b_{0},\dots,b_{d})\neq 1 then the bi(z)b_{i}(z) have common zeroes and none of common zeroes is zero. It is easy to see that (cη+1)/η(c\eta+1)/\eta is a common zero of all ai(y)a_{i}(y) if η\eta is a common zero of all bib_{i}. This contradicts with the fact that gcd(a0,,ad)=1\gcd(a_{0},\dots,a_{d})=1. Secondly, if g(z,z)g(z,z^{\prime}) has a factor with positive degree in zz^{\prime} then f(y,y)f(y,y^{\prime}) will have a factor with positive degree in yy^{\prime}, a contradiction. This proves our claim. Remark that r(t)r(t) is a nontrivial rational solution of g(z,z)=0g(z,z^{\prime})=0 if and only if (cr(t)+1)/r(t)(cr(t)+1)/r(t) is a nontrivial rational solution of f(y,y)=0f(y,y^{\prime})=0. The main result of this paper is the following theorem.

Theorem 5.1.

Assume that f(y,y)=0f(y,y^{\prime})=0 is a first order AODE with positive m.s.index{\rm m.s.index} and assume further that f(y,y)f(y,y^{\prime}) is irreducible over k(t)k(t). Then if r(t)r(t) is a rational solution of f(y,y)=0f(y,y^{\prime})=0 then

deg(r(t))(54n3+9n2+25n2)4n5n2+12211n4+43n2+34T(f).\deg(r(t))\leq(54n^{3}+9n^{2}+2^{5n^{2}})^{4}n^{5n^{2}+12}2^{11n^{4}+43n^{2}+34}T(f).

where n=tdeg(f)n={\rm tdeg}(f).

Proof.

We shall use the notations as above. Due to the above discussion, we only need to consider the differential equation g(z,z)=0g(z,z^{\prime})=0. Denote n=tdeg(f)n={\rm tdeg}(f) and d=deg(f,y)d=\deg(f,y^{\prime}). One sees that T(g)T(f),d=deg(g,z)T(g)\leq T(f),d=\deg(g,z^{\prime}) and

deg(g,z)=2d+=tdeg(g)3n.\deg(g,z)=2d+\ell={\rm tdeg}(g)\leq 3n.

Suppose that

g=h1h1hmg=h_{1}h_{1}\dots h_{m}

where hih_{i} is irreducible over k(t)¯{\overline{k(t)}}. Since gg is irreducible over k(t)k(t), one has that all hih_{i} are conjugate to each other and then

deg(hi,z)=deg(g,z)/m,deg(hi,z)=deg(g,z)/m=d/m.\displaystyle\deg(h_{i},z)=\deg(g,z)/m,\,\,\deg(h_{i},z^{\prime})=\deg(g,z^{\prime})/m=d/m.

By Corollary 2.15, T(hi)T(g)T(f)T(h_{i})\leq T(g)\leq T(f). Assume that r(t)r(t) is a rational solution of g(z,z)=0g(z,z^{\prime})=0 then r(t)r(t) is a rational solution of all hi=0h_{i}=0. In particular, h1(r(t),r(t))=0h_{1}(r(t),r^{\prime}(t))=0. Denote n~=tdeg(h1)\tilde{n}={\rm tdeg}(h_{1}) and d~=deg(h1,z)\tilde{d}=\deg(h_{1},z^{\prime}). Then

n~\displaystyle\tilde{n} =tdeg(g)/m=(2d+)/m3n/m\displaystyle={\rm tdeg}(g)/m=(2d+\ell)/m\leq 3n/m
d~\displaystyle\tilde{d} =deg(g,z)/m=d/m.\displaystyle=\deg(g,z^{\prime})/m=d/m.

Set N=n~2N=\tilde{n}^{2}. By Theorem 4.3 and Remark 2.7, one has that

NN+n~deg(g,z)deg(g,z)deg(r(t))Cd~deg(r(t))\frac{N}{N+\tilde{n}}\frac{\deg(g,z)}{\deg(g,z^{\prime})}\deg(r(t))-\frac{C}{\tilde{d}}\leq\deg(r^{\prime}(t))

where

C=(2n~3+n~2+2s2)4n~s+7(n~+1)42s22+15s2+10T(f).\displaystyle C=(2\tilde{n}^{3}+\tilde{n}^{2}+2^{s-2})^{4}\tilde{n}^{s+7}(\tilde{n}+1)^{4}2^{\frac{s^{2}}{2}+\frac{15s}{2}+10}T(f).

Note that ss is the number of quadratic transformations applied to transfer h1=0h_{1}=0 to an algebraic curve with only ordinary singularities. Due to Theorem 2 in Chapter 7 of [9], ss can be chosen to be an integer not greater than

(n~1)(n~2)/2(3n1)(3n2)/29n2/2.(\tilde{n}-1)(\tilde{n}-2)/2\leq(3n-1)(3n-2)/2\leq 9n^{2}/2.

Remark that deg(r(t))2deg(r(t))\deg(r^{\prime}(t))\leq 2\deg(r(t)). Thus

(NN+n~deg(g,z)deg(g,z)2)deg(r(t))Cd~.\left(\frac{N}{N+\tilde{n}}\frac{\deg(g,z)}{\deg(g,z^{\prime})}-2\right)\deg(r(t))\leq\frac{C}{\tilde{d}}. (7)

As mm divides both deg(g,z)\deg(g,z) and deg(g,z)\deg(g,z^{\prime}), mm divides \ell. Set =m¯\ell=m\bar{\ell}. Then

NN+n~deg(g,z)deg(g,z)2=n~2(2d+)(n~2+n~)d2=n~2d(n~+1)d\displaystyle\frac{N}{N+\tilde{n}}\frac{\deg(g,z)}{\deg(g,z^{\prime})}-2=\frac{\tilde{n}^{2}(2d+\ell)}{(\tilde{n}^{2}+\tilde{n})d}-2=\frac{\tilde{n}\ell-2d}{(\tilde{n}+1)d} ¯deg(g,z)2d(n~+1)d\displaystyle\geq\frac{\bar{\ell}\deg(g,z)-2d}{(\tilde{n}+1)d}
1(n~+1)d.\displaystyle\geq\frac{1}{(\tilde{n}+1)d}.

This together with (7) implies that

deg(r(t))\displaystyle\deg(r(t)) m(n~+1)C\displaystyle\leq m(\tilde{n}+1)C
4n(2n~3+n~2+2s2)4n~s+7(n~+1)42s22+15s2+10T(f)\displaystyle\leq 4n(2\tilde{n}^{3}+\tilde{n}^{2}+2^{s-2})^{4}\tilde{n}^{s+7}(\tilde{n}+1)^{4}2^{\frac{s^{2}}{2}+\frac{15s}{2}+10}T(f)
n(54n3+9n2+2s2)4(4n)s+112s22+15s2+12T(f)\displaystyle\leq n(54n^{3}+9n^{2}+2^{s-2})^{4}(4n)^{s+11}2^{\frac{s^{2}}{2}+\frac{15s}{2}+12}T(f)
(54n3+9n2+2s2)4ns+122s22+19s2+34T(f)\displaystyle\leq(54n^{3}+9n^{2}+2^{s-2})^{4}n^{s+12}2^{\frac{s^{2}}{2}+\frac{19s}{2}+34}T(f)
(54n3+9n2+25n2)4n5n2+12211n4+43n2+34T(f).\displaystyle\leq(54n^{3}+9n^{2}+2^{5n^{2}})^{4}n^{5n^{2}+12}2^{11n^{4}+43n^{2}+34}T(f).

The second inequality holds because m(n~+1)3n+m4nm(\tilde{n}+1)\leq 3n+m\leq 4n. ∎

Remark 5.2.

Theorem 5.1 implies that an autonomous first order AODE f=0f=0 with positive m.s.index{\rm m.s.index} has no nontrival rational solutions, because T(f)=0T(f)=0. In fact, suppose that f=0f=0 has a nontrival rational solution. Then it will have infinitely many rational solutions. By Corollary 4.6 of [7], f=0f=0 has no movable singularities. However, as f=0f=0 has positive m.s.index{\rm m.s.index}, Fuchs’ criterion implies that f=0f=0 has movable singularities, a contradiction.

In [21], the authors developed two algorithms to compute rational solutions of maximally comparable first-order AODEs and first order quasi-linear AODEs respectively. Let us first recall the definition of maximally comparable first order AODEs. Suppose that f=i,jai,jyiyjf=\sum_{i,j}a_{i,j}y^{i}y^{\prime j} is a differential polynomial over k(t)k(t). Denote

S(f)={(i,j)2ai,j0}.S(f)=\{(i,j)\in\mathbb{N}^{2}\mid a_{i,j}\neq 0\}.

If there is (i0,j0)S(f)(i_{0},j_{0})\in S(f) satisfying that i0+j0i+ji_{0}+j_{0}\geq i+j and i0+2j0>i+2ji_{0}+2j_{0}>i+2j for every (i,j)S(f)(i,j)\in S(f), then we say that ff is maximally comparable. The following examples shows that their algorithms can not deal with all first order AODEs with positive m.s.index{\rm m.s.index}.

Example 5.3.

Let

f=yym+y2m+1+tf=yy^{\prime m}+y^{2m+1}+t

where m1m\geq 1. Then S(f)={(1,m),(2m+1,0),(0,0)}S(f)=\{(1,m),(2m+1,0),(0,0)\}. Since 2m+1+01+m2m+1+0\geq 1+m but 2m+1+20=1+2m2m+1+2\cdot 0=1+2\cdot m, ff is not maximally comparable. On the other hand, we have that m.s.index(f)=1>0.{\rm m.s.index}(f)=1>0.

References

References

  • [1] J. M. Aroca, J. Cano, R. Feng, and X. S. Gao. Algebraic general solutions of algebraic ordinary differential equations. In ISSAC’05, pages 29–36. ACM, New York, 2005.
  • [2] Moulay A. Barkatou. On rational solutions of systems of linear differential equations. volume 28, pages 547–567. 1999. Differential algebra and differential equations.
  • [3] L. X. Châu Ngô and Franz Winkler. Rational general solutions of first order non-autonomous parametrizable ODEs. J. Symbolic Comput., 45(12):1426–1441, 2010.
  • [4] Claude Chevalley. Introduction to the theory of algebraic functions of one variable. Mathematical Surveys, No. VI. American Mathematical Society, Providence, R.I., 1963.
  • [5] A. Eremenko. Rational solutions of first-order differential equations. Ann. Acad. Sci. Fenn. Math., 23(1):181–190, 1998.
  • [6] Ruyong Feng and Xiao-Shan Gao. A polynomial time algorithm for finding rational general solutions of first order autonomous ODEs. J. Symbolic Comput., 41(7):739–762, 2006.
  • [7] Shuang Feng and Ruyong Feng. Descent of ordinary differential equations with rational general solutions. J. Syst. Sci. Complex., 2020.
  • [8] James Freitag and Rahim Moosa. Finiteness theorems on hypersurfaces in partial differential-algebraic geometry. Adv. Math., 314:726–755, 2017.
  • [9] William Fulton. Algebraic curves. Advanced Book Classics. Addison-Wesley Publishing Company, Advanced Book Program, Redwood City, CA, 1989. An introduction to algebraic geometry, Notes written with the collaboration of Richard Weiss, Reprint of 1969 original.
  • [10] F. Hess. Computing Riemann-Roch spaces in algebraic function fields and related topics. J. Symbolic Comput., 33(4):425–445, 2002.
  • [11] Ming-Deh Huang and Doug Ierardi. Efficient algorithms for the Riemann-Roch problem and for addition in the Jacobian of a curve. J. Symbolic Comput., 18(6):519–539, 1994.
  • [12] Jerald J. Kovacic. An algorithm for solving second order linear homogeneous differential equations. J. Symbolic Comput., 2(1):3–43, 1986.
  • [13] Serge Lang. Fundamentals of Diophantine geometry. Springer-Verlag, New York, 1983.
  • [14] Michihiko Matsuda. First-order algebraic differential equations, volume 804 of Lecture Notes in Mathematics. Springer, Berlin, 1980. A differential algebraic approach.
  • [15] Joseph Fels Ritt. Differential algebra. Dover Publications, Inc., New York, 1966.
  • [16] J. Rafael Sendra and Franz Winkler. Tracing index of rational curve parametrizations. Comput. Aided Geom. Design, 18(8):771–795, 2001.
  • [17] Michael F. Singer. Liouvillian solutions of nnth order homogeneous linear differential equations. Amer. J. Math., 103(4):661–682, 1981.
  • [18] Marius van der Put and Michael F. Singer. Galois theory of linear differential equations, volume 328 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 2003.
  • [19] Mark van Hoeij, Jean-François Ragot, Felix Ulmer, and Jacques-Arthur Weil. Liouvillian solutions of linear differential equations of order three and higher. volume 28, pages 589–609. 1999. Differential algebra and differential equations.
  • [20] N. Thieu Vo, Georg Grasegger, and Franz Winkler. Deciding the existence of rational general solutions for first-order algebraic ODEs. J. Symbolic Comput., 87:127–139, 2018.
  • [21] Thieu N. Vo, Georg Grasegger, and Franz Winkler. Computation of all rational solutions of first-order algebraic ODEs. Adv. in Appl. Math., 98:1–24, 2018.
  • [22] Franz Winkler. The algebro-geometric method for solving algebraic differential equations—a survey. J. Syst. Sci. Complex., 32(1):256–270, 2019.