Rational Solutions of First Order Algebraic Ordinary Differential Equations
Abstract
Let be a first order ordinary differential equation with polynomial coefficients. Eremenko in 1999 proved that there exists a constant such that every rational solution of is of degree not greater than . Examples show that this degree bound depends not only on the degrees of in but also on the coefficients of viewed as polynomial in . In this paper, we show that if
then the degree bound only depends on the degrees of , and furthermore we present an explicit expression for in terms of the degrees of .
keywords:
first order AODE, rational solution, degree bound, heightMSC:
[2010] 34A05, 68W301 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 of has branch points , where is an arbitrary constant, so 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 be an integer. Then is a rational solution of . The degree of depends on the constant coefficient of .
Let be an irreducible first order AODE. Set
Fuchs’ criterion (see Remark on page 14 of [14]) implies that has movable singularities if . On the other hand, it was proved in [5] that if has movable singularities then it can be transferred into an AODE with positive . This motivates us to focus on first order AODEs with positive . We prove that for an irreducible first order AODE with the degrees of rational solutions of are independent of the constant coefficients of and furthermore we present an explicit degree bound in terms of the degrees of . 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, stands for the ring of integers, for algebraically closed fields of characteristic zero, and and for algebraic function fields over and respectively. denotes the projective space of dimension over a field and 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 is a map satisfying that for all ,
A ring (resp. field) equipped with a derivation is called a differential ring (resp. differential field). An ideal is called a differential ideal if .
The field of rational functions in can be endowed with a structure of differential field whose derivation is the usual derivation with respect to , i.e. . Set and denote
where are indeterminates. One can extend the derivation on to a derivation on by assigning so that becomes a differential ring. For the sake of notations, we use in place of . Elements in are called differential polynomials over . Let be a differential polynomial not in . Then there is a unique integer such that . This integer is called the order of . We shall use (resp. ) to stand for the differential (resp. algebraic) ideal generated by a set of differential polynomials (resp. polynomials) respectively. Suppose that is irreducible viewed as an algebraic polynomial. Set
where and is the order of . It was proved on page 30 of [15] that is a prime differential ideal and so is a differential domain. Lemma 2.2 of [7] implies that the field of fractions of is isomorphic to that of . Under this isomorphism, the field of fractions of can be endowed with a structure of differential field. We shall still use , or ′ in short, to denote the induced derivation on the field of fractions of .
In this paper, the first order AODEs under consideration are differential equations of the following form
(1) |
where .
Definition 2.2.
An element satisfying is called a rational solution of .
Remark that the derivation in can be uniquely extended to a derivation in which we shall still denote by . Assume that viewed as a polynomial in , is irreducible over . Then the field of fractions of is not only an algebraic function field over but also a differential field.
2.2 Algebraic function fields of one variable
Let be an algebraically closed field of characteristic zero and an extension field of . We say is an algebraic function field of one variable over if satisfies the following conditions: there is an element of which is transcendental over , and is algebraic of finite degree over . Assume is an algebraic function field of one variable over . A valuation ring of over is a subring satisfying that
-
1.
; and
-
2.
if , then .
All non-invertible elements of form a maximal ideal which is called a place of , and is called the corresponding valuation ring of . Let be a valuation ring with as place. There is an element , called a local uniformizer of or , such that and . The factor ring is equal to since is algebraically closed. For every valuation ring with place , we define a map
satisfying if then , otherwise . It is well-known that admits infinitely many places, and there is one-to-one correspondence between places and valuation rings.
Let be a place of and the corresponding valuation ring of . Let be a local uniformizer of . Then for every non-zero element of , there is a unique integer such that
for some invertible element . It is easy to see that the integer is independent of the choice of local uniformizers. Such is called the order of at and denoted by . We make the convention to write . Then the place induces a map from to sending to . This map is called the order function at . For , we have
where the equality in the later formula holds if . Let and be a place. We say is a zero of if , and a pole of if . Every non-zero element of admits only finitely many zeros and poles.
A divisor in is a formal sum
for all the places of , where and for all but finitely many . It is easy to see that the set of divisors in forms an abelian group. is effective if for all . The degree of , denoted by , is defined to be and the support of , denoted by , is defined to be . For brief, we denote
Let and be two divisors in , we write provided is effective. For every non-zero element of , we denote
where ranges over all places of . Then is a divisor of degree . For a divisor , we denote
which is called the Riemann-Roch space of . It is well-known that each Riemann-Roch space is a -vector space of finite dimension. The Riemann-Roch Theorem implies that if is a divisor whose degree is not less than the genus of then is of positive dimension.
Let be irreducible. One sees that the field of fractions of is an algebraical function field of one variable over which is called the algebraic function field of . For an irreducible homogeneous polynomial in , the corresponding algebraic function field is defined to be the algebraic function field of . Remark that the algebraic function fields of and are all isomorphic.
2.3 Models of algebraic function fields of one variable
Let be an algebraic function field of one variable over . The set of all places of can be viewed as a nonsingular model of . On the other hand, let be an irreducible homogeneous polynomial whose algebraic function field is . Then the projective curve is another model of . There is a surjective map from a nonsingular model of to the curve . To describe this map precisely, let be three nonzero elements of satisfying that
Set . Let be a place of with as local uniformizer. Denote by . One sees that and moreover not all are zero. Therefore defines a point of . Remark that this point does not depend on the choice of .
Definition 2.3.
We call the center of with respect to . Denote by the set of centers with respect to .
We claim that is the plane projective curve in defined by and the map sending to the center of with respect to is the required map. It is easy to verify that for all . Conversely, let be a point of . Without loss of generality, we may assume that . Then . Remark tha . As , due to Corollary 2 on page 8 of [4], there is a place containing and . For this place, one has that and furthermore . Write . Then the center of with respect to is
This implies that .
Definition 2.4.
We call or a plane projective model of .
The plane projective models of usually have singularities. Let be an irreducible projective curve in defined by a homogeneous polynomial . A point of is said to be of multiplicity , if all derivatives of up to and including the -th vanish at but not all the -th derivatives vanish at . Suppose that is a point of with multiplicity . If , then is called a simple point of , otherwise a singular point of . A point of multiplicity is ordinary if the tangents to at this point are distinct, otherwise it is non-ordinary. Due to Propositon on page of [9], has always a plane projective model with only ordinary singularities.
Let be an invertible transformation, where are homogeneous polynomials in of the same degree and they have no common factors. We further assume that for all . Then
Proposition 2.5.
Let be as above and a place of . Assume that is the center of with respect to . If , then is the center of with respect to .
Proof.
Let be a local uniformizer of and . One has that
for some nonzero . Denote . Then
This implies that In other words,
Then the center of with respect to is
Hence is the center of with respect to . ∎
2.4 Heights
All algebraic function fields under consideration in this subsection are finite extensions of . They are algebraic function fields of one variable over 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 .
-
1.
Given , let be a finite extension of containing all . We define the height of , denoted by , to be
where ranges over all places of .
-
2.
For , we define
-
3.
For , we define the height of to be , denoted by .
-
4.
Let be a polynomial in . Suppose that contains at least two terms. We define the height of , denoted by , to be where is the point in a suitable projective space formed by the coefficients of . For convention, when only contains one term, we defined to be zero.
-
5.
Let be a hypersurface in defined by . We define the height of , denoted by , to be .
Remark 2.7.
Assume that .
-
1.
One sees that is independent of the choice of homogeneous coordinates and the choice of . Without loss of generality, we suppose , then
-
2.
Assume is a finite extension of containing all and . Then one sees that if for all places of then .
-
3.
Suppose that and . Then
To see this, let . Then
where ranges over all the places of . Note that every place of has a local uniformizer of the form or for some . Suppose the place has as a local uniformizer. Then if and only if . Since , there is some such that . This implies that for places with as local uniformizers,
For the place with as local uniformizer, one has that . So for this place,
Consequently, .
-
4.
Let and . Let be a nonzero irreducible polynomial over such that . It is clear that if . Assume that and are all distinct poles of in , then
In particular, if then which is defined to be the maximun of the degrees of the denominator and numerator of .
From the above remark, it is easy to see that for and
Proposition 2.8.
Let , with . Then
-
1.
if ;
-
2.
;
-
3.
for all .
Proof.
If then the assertion is obvious. Suppose that . Let . Then . The assertion follows from Remark 2.7 and the fact that
Use an argument similar to that in 2. and the fact that
∎
The following examples show that the equalities may hold in and of Proposition 2.8.
Example 2.9.
Let and . Then
Moreover both of them are greater than the maximun of .
In the following, stands for the vector with indeterminates and denotes for .
Proposition 2.10.
Let and be polynomials in , then
-
1.
-
2.
If both and have 1 as a coefficient, then
Proof.
Write and with . Let be a finite extension of containing all , and a place of .
1. Each coefficient of is of the form , where . Since
we have
where is the set of all coefficients of . It follows that
2. The assertion follows from the fact that
∎
Proposition 2.11.
Let where and . Suppose that is a zero of in and is a finite extension of containing and all . Then for each place of ,
Proof.
The assertion is clear if or is not a pole of . Assume and is a pole of . Then
for some with . This together with the fact that implies that
i.e.
Consequently, one has that
∎
Corollary 2.12.
Let be a polynomial in and a zero of in . Then .
The equality in Corollary 2.12 may hold as shown in the following example.
Example 2.13.
Let
then .
Lemma 2.14.
Assume and is a factor of . Then
Proof.
Without loss of generality, we may assume both and are monic. Write
We first show that if and then . Suppose that
where . Let . For each place of , denote
Then for all and all places of . From , one has that
We claim that for all and all places of . For , one has that
This implies that . By Proposition 2.11, we have that . Hence for all places of . Now assume that there is a place of and with such that but . Then one has that and for all . On the other hand, since , i.e. . From , one has that
The last equality holds because for all . Thus which implies that
for all . As , one has that and furthermore for all . In other words, . Applying a similar argument to the equalities for successively yields that and for all . However, one has that . This implies that and thus . That is to say, , a contradiction. This proves the claim. This claim and Remark 2.7 imply that .
Now we prove the assertion by induction on . The base case is obvious. Suppose that the assertion holds for . Consider the case . If then there is nothing to prove. Suppose that . Then there is such that divides . Let for some . Then and divides . By induction hypothesis, one has that . ∎
Corollary 2.15.
Assume and is a factor of . Then
Proof.
Suppose that where . Let be an integer greater than . One has that
Note that for , if and only if . This implies that the set of the coefficients of coincides with that of . Hence . Similarly, . It is clear that is still a factor of . By Lemma 2.14, . Thus . ∎
Proposition 2.16.
Assume , then
where is the resultant of and with respect to .
Proof.
The assertion is clear that if . Consider the case . Assume . Write
where . Denote further by the sets of the coefficients in of respectively. Then
By the definitions of determinant, we can write
where , is a monomial in with total degree and is a monomial in with total degree . Let . For each place of , we have
Therefore by Remark 2.7,
∎
3 Degrees and Heights on Riemann-Roth spaces
Throughout this section, denotes an algebraic function field of one variable over . Let be such that a plane projective model of , i.e. . Each can be presented by where are two homogeneous polynomials in of the same degree and having no common factors, and . We call a representation of . 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 and where 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, always stands for an irreducible homogeneous polynomial in of degree which defines a plane projective model of , i.e. there is such that and .
3.1 Quadratic transformations
Let be a divisor in . Due to Proposition on page of [9], there is a birational transformation such that the transformation of under is a plane projective curve with only ordinary singularties, moreover 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 under a quadratic transformation.
Definition 3.1.
-
1.
stands for a projective change of coordinates on that is defined as where , and denotes the standard quadratic transformation that is defined as
where .
-
2.
The height of , denoted by , is defined as .
Notation 3.2.
-
1.
stands for .
-
2.
stands for the irreducible polynomial satisfying
where .
One sees that (resp. ) is the variaty (resp. the projective closure of ).
Remark 3.3.
is bijective on and .
Let us first bound the heights of the common points of two algebraic curves in .
Proposition 3.4.
Let be two homogenenous polynomials in of degree respectively. Suppose that and have no common factor, and is a common point of and . Then
Furthermore, if with then .
Proof.
Let where . Proposition 2.16 implies that Without loss of generality, suppose . Since is a common point of and ,
It follows from Corollary 2.12 that for all . Let and a place of . Then
Whence
It remains to pove the second assertion. Since , not all are zero. Without loss of generality, assume that . Substituting into yields that
This implies that . On the other hand, one sees that
So
which results in . ∎
Corollary 3.5.
If is a singular point of , then
Lemma 3.6.
Suppose that is a projective change of coordinates. Then
-
1.
;
-
2.
for each , .
Proof.
Suppose that and
where and .
1. One has that
Let be a coefficient of viewed as polynomial in . Then is a -linear combination of monomials with . Let be a finite extension of containing all and . Suppose that is a place of . Then one has that
i.e.
Therefore due to Remark 2.7.
2. Suppose that and . Then . Let be a finite extension of containing all and , and a place of . Then
i.e.
So . ∎
Corollary 3.7.
Suppose that . Let be a projective change of coordinates with
(2) |
where . Then
-
1.
;
-
2.
;
-
3.
for each , .
Proof.
The first assertion is obvious. The second and third assertions follows from Lemma 3.6 and the fact that and . ∎
Definition 3.8.
-
1.
The points are called fundamental points.
-
2.
Assume is a singular point of with multiplicity . is said to be in excellent position if it satisfies the following two conditions:
-
(a)
the line intersects in distinct non-fundamental points;
-
(b)
the lines intersect in distinct points other than fundamental points.
-
(a)
Lemma 3.9.
Suppose that is a singular point of of multiplicity . There is a projective change of coordinates with having the form (2) such that and is in excellent position.
Proof.
Denote . Let be the projective change of coordinates with of the form
One sees that there are polynomials such that with satisfies the required conditions if and only if
Note that if is a projective change of coordinates such that then for some . Due to Lemma 1 on page of [9], there are projective changes of coordinates satisfying the above conditions. In other words, . Therefore . For every , is as required. ∎
Lemma 3.10.
-
1.
;
-
2.
For each , .
Proof.
Assume that where . Then
From this, one sees that the set of coefficients of is equal to that of . Hence .
One has that . Let be a finite exntesion of containing all and a place of . Note that
i.e. . So . ∎
Definition 3.11.
-
1.
We call a projective change of coordinates in Lemma 3.9 a projective change of coordinates centered at .
-
2.
Let be a projective change of coordinates centered at and the stand quadratic transformation. We call , the composition of and , a quadratic transformation centered at .
Notation 3.12.
Let be a quadratic transformation centered at . We shall denote .
Corollary 3.13.
Let be a singular point of and a quadratic transformation centered at . Then
-
1.
;
-
2.
for , ;
Proposition 3.14.
Let be a plane projective model of defined by and a place of . Let be the center of with respect to . Assume that is a quadratic transformation centered at for some singular point of and is the center of with respect to . Then
Proof.
We first claim that if then . Otherwise assume that . Then is a fundamental point of . Since neither nor is a point of . One has that . Hence . This proves our claim.
Suppose that . Then and thus by Proposition 2.5. Corollary 3.13 then implies that Now suppose that . Denote . By Proposition 2.5 again, is the center of with respect to . Suppose that is a local uniformizer of and . From the definition of center, one sees that for all . Write where , with . One then has that
where with . Set
Since both and are greater than , or . In the case that , one has that . So . By Propositon 3.4, . In other cases, one sees that is a fundamental point and so . Hence, in each case, one has that
∎
3.2 Degrees and Heights for Riemann-Roch Spaces
Let be a disivor in where . Let defined by be a plane projective model of . Suppose that and where are two homogeneous polynomials of the same degree in . In this subsection, we shall estimate and in terms of and . For this, we introduce the following notations and definitions.
Notation 3.15.
Let be a plane projective model of and a divisor.
-
1.
.
-
2.
.
Definition 3.16.
Let be two homogeneous polynomials in of the same degree. Write .
-
Define
-
Define
where the sum ranges over all places of . Furthermore, define
It is easy to see that and if and only if where is the center of with respect to . Futhermore for all nonzero .
Remark 3.17.
On page 182 of [9], is defined to be the order at of the image of in the valutaion ring of . Remark that
where satisfies that . Under the map sending to for all , is sent to which lies in the valuation ring of . Therefore given in Definition 3.16 concides with the one given in [9] and is nothing else but the intersection cycle of and (see page 119 of [9]).
The lemma below follows easily from the definition.
Lemma 3.18.
Suppose that are two homogeneous polynomials of the same degree. Then
-
1.
; and
-
2.
.
Lemma 3.19.
Suppose that is an ordinary singular point of of multiplicity , and is a finite set of points of .
-
1.
Let . Then for all but a finite number of , intersects in distinct points other than the points in .
-
2.
Write
where is a homogeneous polynomial in of degree . Assume that . Let
where satisfies that . Then for all but a finite number of , intersects in distinct points other than the points in .
Proof.
Under the projective change of coordinates with and , we may assume that . Set where . Substituting into yields that
Set . For every root of , one sees that is a common point of and other than . Moreover if has distinct roots then intersects in distinct points other than . So it suffices to prove that for all but a finite number of , has distinct roots. Note tht substituting into yields that
From this, one sees that if is a common root of and then is a common point of and . Since and have only finitely many common points, there are only finitely many such that has multiple roots. In other words, there is only a finite number of such that intersects in less than distinct points other than . It remains to show that there are only finitely many such that . Assume that which lies in the line . If then
If then and thus . In other words, or , which is impossible.
Similarly, we may assume that . Then
Applying the standard quadratic transformation to and , one obtains that
Since is an ordinary singular point and , by on page of [9] is a simple point of . Moveover, as , neither nor is a point of and so . Thus
For every common point of and with , and then it is a common point of and other than . Therefore it suffices to show that for all but a finite number of , and have distinct common points with . Let be the projective change of coordinates such that . Then . Note that which is a simple point of . Thus
By , for all but a finite number of , intersects in distinct points with . Remark that if is a common point of and with then is a common point of and with . These imply that for all but a finite number of , intersects in distinct points with .
Finally, we need to prove that there are only finitely many such that . Assume that which lies in . We claim that . Suppose on the contrary that . Then by , one sees that either or both and are zero. This implies that must be one of three points . This is impossible and then our claim holds. It follows from the claim that is uniquely determined by . ∎
Proposition 3.20.
Suppose that has only ordinary singularities and is an effective divisor in . Let be a divisor in .
-
1.
Assume further that where all have the same center which is a point of with multiplicity . Then there is a linear homogeneous polynomial in such that
where is a very simple and effective divisor of degree , , and
-
2.
Assume that where the center of is a singular point of . Then there are two homogeneous polynomials of degree two such that
where is a very simple divisor, and
Proof.
Suppose is the center of with respect to . Without loss of generality, we assume that and . Set . Due to Lemma 3.19, for all but a finite number of , intersects in distincet points other than the points in . Let be such that satisfies the above condition. Then
where is an effective divisor of degree and . It is clear that is very simple since intersects in distinct points other than . Finally, one easily sees that . As the points in are the intersection points of and , by Proposition 3.4.
Suppose that is the center of with respect to , and is of multiplicity . Since is an ordinary singular point, there are exactly places of with as the center with respect to . Denote these places by . Without loss of generality, we may assume that and . Write
where is a homogeneous polynomial of degree . Choose a projective change of coordinates with such that
satisfies that , where and . By Lemma 3.6, . Denote
Then is the center of with respect to . For , write
where is a local uniformizer of , , not all zero, and . Furthermore,
where . This implies that . Since , . Set and
Due to Lemma 3.19, for all but a finite number of , intersects in distinct points other than the points . Let be the very simple divisor consisting of the places whose centers with respect to are the intersection points of and other than respectively. Then . We claim that for the above ,
Note that
where . This implies that . Hence
On the other hand, since , one has that
This proves our claim. Now set . As , one sees that
This implies that
Therefore . Note that we can choose . For such , one has that
Since , one sees that . This implies that and
Now applying to the case that and , one gets a linear homogeneous polynomial such that , where is a very simple divisor satisfying that . Moreover . Let be a linear homogeneous polynomial in such that intersects in distinct points other than the points in . For such , one has that is a very simple divisor satisfying that . Set and . Then and we obtain two polynomials as required. Note that and . Hence ∎
Definition 3.21.
Suppose that has only ordinary singularities, say , and is the multiplicity of . Suppose further that for each , are all places of with as center with respect to . Set
A homogeneous polynomial such that is called an adjoint of .
We have the following two corollaries of Proposition 3.20.
Corollary 3.22.
Suppose that is a simple and effective divisor in . Let be a divisor in . Then there is a homogeneous polynomial of degree not greater than such that
where is a very simple and effective divisor of degree not greater than
such that and . Moreover
Proof.
Denote where is given as in Definition 3.21. Note that , . Write
where the centers of is a simple point of , for some . Applying succesively Proposition 3.20 to and , one obtains linear homogeneous polynomials such that if , or if , where is a very simple and effective divisor such that
Set and . Then one has that
Moreover by Proposition 3.20, for all and then Proposition 2.10 implies that . It is obvious that is not greater than because so is for all . ∎
Corollary 3.23.
Suppose that are two divisors in . Then there are two homogeneous polynomials of the same degree such that is very simple and . Moreover and
where .
Proof.
We first show the case that is effective. Denote and write
where the center of (resp. ) with respect to is a simple (resp. singular) point of . Applying Proposition 3.20 to yields a homogenenous polynomial of degree such that where is a very simple and effective divisor such that . Moreover . Construct linear homogeneous polynomials in such that is very simple and . It is easy to see that . Set . By Proposition 3.20 again, one obtains pairs of homogeneous polynomials of degree two such that where is a very simple divisor such that
and , . Set and . Then
which is very simple. It is clear that , and by Proposition 3.20
Similarly, . Furthermore, one has that
and
For the general case, write . The previous discussion implies that we can obtain such that and are very simple. Moreover
Set and . Then satisfies the required condition. Furthermore and
∎
Now we are ready to pove the main results of this section. Let us start with two lemmas.
Lemma 3.24.
Suppose that has only ordinary singularities and is a divisor in . Let be a homogeneous polynomial in such that , where is an effective divisor. Then
(3) |
Proof.
Lemma 3.25.
Assume that is an matrix with . Assume further that for each place of , where and for all but finite number of , . Then there is a basis of the solution space of satisfying that
for all .
Proof.
Assume that . Then . Without loss of generality, we may assume the first -rows of are linearly independent and denote by the matrix formed by them. Then the solution space of is the same as that of . Hence it suffices to consider the system . We may further assume that the matrix formed by the first -columns of is invertible. For every and , set to be the determinant of the matrix obtained from by replacing the -th column of by the -th column of . For each , denote
where denotes the transpose of a vector. Then by Cramer’s rule, the are solutions of and thus they form a basis of the solution space of . Note that as well as is an integer combination of the monomials in the entries of of total degree . So for all and ,
where is a place of . This together Remark 2.7 implies the lemma. ∎
Theorem 3.26.
Suppose that has only ordinary singularities. Let be a divisor in . Denote and Then there is a -basis of such that every element of can be represented by where are two homogeneous polynomials of the same degree not greater than and
Proof.
By Corollary 3.23, there are two homogeneous polynomials of the same degree such that is very simple. Moreover
Due to Corollary 3.22, there is a homogeneous polynomial of degree not greater than such that , where is a very simple and effective divisor and . Moreover
and
Denote . By Lemma 3.24, to compute , it suffices to compute all homogeneous polynomials of degree satisfying that
Assume that
where are indeterminates. There are indeterminates in total. For each , if and only if the center of with respect to is a zero of . This imposes linear constraints on . At the same time, if and only if the center of with respect to is a common zero of
for all nonnegative integers satisfying that , where is as in Definition 3.21. This imposes linear constraints on . So there are totally linear constraints on . The problem of finding is reduced to that of solving the system , where is a vector with indeterminates entries and is a matrix. Denote by the center of in . Then the entries in the same row of are monomials of total degree in for some in . Without loss of generality, we may assume that one of is 1. Let be a finite extension of containing all . For each place of , set
Since for all , and
where . Note that
Applying Lemma 3.25 to yields that
The last inequality holds because
Set and . Then
∎
Next, we consider the case that may have non-ordinary singular points. Let be a plane projective model of . Suppose that is defined by and for , is the quadratic transformation of under the quadratic transformation , where is a singular point of . Denote and set , and
(4) |
Proposition 3.27.
Let be a divisor in and the notations as above. One has that
-
1.
;
-
2.
.
Proof.
1. Set . Since every is a singular point, one has that . This implies .
2. Denote by the maximum of the heights of singular points of . We first prove by induction on that for all . Note that for all . Since , it is clear that and by Corollary 3.5. Now assume that for . Consider the case . By Corollary 3.13 and induction hypothesis, one has that
Note that as the curve has singularities. One sees that if and
Consequently, for all . On the other hand, one has already seen that By Corollary 3.13 again,
For the divisor , it is obvious that . Suppose that for . By Corollary 3.13 and the induction hypothesis,
∎
Notation 3.28.
Let be the defining polynomial of . Denote by the number of quadratic transformations such that has only ordinary singularities, where is the image of under these quadratic transformations. By Theorem 2 in Chapter 7 of [9], can be chosen to be an integer not greater than
where , is the number of non-ordinary singularities of , ranges over all singularities of and is the multiplicity of .
Theorem 3.29.
Let be a divisor in . Denote
Then there is a -basis of such that every element of can be represented by where are two homogeneous polynomials of the same degree not greater and
Proof.
If has only ordinary singularities, i.e. , then the assertion is clear by Theorem 3.26. Suppose has non-ordinary singularities, i.e. . Let be the image of under quadratic transformations such that has only ordinary singularities. Let be the defining polynomial of and set
Then by Proposition 3.27
By Theorem 3.26, there is a -basis of satisfying that each element in can be represented by where are homogeneous polynomials of degree not greater than and
It remains to represent elements of in terms of . We use the same notations as in the proof of Propositoin 3.27. Let and . Denote by the degree of the defining polynomial of and the maximum of the heights of singular points of . Let and for all One sees that . By Lemmas 3.6 and 3.10,
where . From the proof of Proposition 3.27, we have
where is given as in (4). Note that . One has that
Similarly, we obtain bounds for and . ∎
4 Heights on plane algebraic curves
Let be an irreducible polynomial over and satisfy , i.e. is a rational parametrization of . The result on parametrization (see [16] for instance) implies that
In other words, . A similar relation holds for points in algebraic curves defined over , i.e there is a constant only depending on such that if is a point of with coordinates in then
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 , Eremenko in 1999 presented another proof which actually provides a procedure to find explicitly. In this section, we shall present an explicit formula for following Eremenko’s proof.
4.1 Heights on plane projective curves
Throughout this subsection, is an irreducible polynomial in and is the algebraic function field over associated to . Let us start with a refinement of Lemma 1 of [5].
Lemma 4.1.
Assume that is irreducible over and satisfying . If , then for every place of with , we have that
Proof.
Since , by Proposition 2 of [5], can be written to be of the form
where with . Write with . Let be a finite extension of containing all and . Suppose that is a place of . Then
for some . Equivalently,
Therefore
This implies that . ∎
Lemma 4.2.
Let be a finite set of places in and . Then there are with such that
Proof.
Set
Then is a finite set in . Let satisfy that and for all . For with , one has that
This implies that . On the other hand, for with , one has that
In both cases, is not a pole of . Thus 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 be an irreducible polynomial in of degree with respect to and of degree with respect to . Suppose that and . Then for every satisfying , one has that
where
(5) |
and is the number of quadratic transformations which are applied to resolve the singularities of .
Proof.
If one of is in then the height of the other one is not greater than . The inequalities then obviously hold. In the following, we assume that neither nor is in .
Let be the algebraic function field associated to and satisfy that . Choose such that and
Such exist due to Lemma 4.2. Set . Consider the divisor
Note that , and
So
This implies that is greater than the genus of and thus . Denote and by the homogenization of . We claim that . Note that For each , is of the form where satisfies that . So and then . If then each point in is of the form too and so . Otherwise, for each , one has that
Moreover . This implies that each point of is of the form whose height is not greater than . Thus . Our claim is proved. Note that
Suppose that . Due to Theorem 3.29, where are two homogeneous polynomials of degree not greater than
and
Set
Without loss of generality, we assume that and have no common factor. Otherwise, by Corollary 2.15, we may replace and by and where is the greatest common factor of and . Moreover, multiplying by suitable elements in if necessary, we can assume that both and have 1 as a coefficient. Let be a place of containing and . Then and . As , . If , then and so . Consequently, is a common point of and . Proposition 3.4 implies that . It is easy to verify that in this case satisfy the required inequalities. Therefore we only need to prove the case .
Set
where denotes the resultant of and with respect to . Note that , because and have no common factor. As is not effective, . Furthermore, as , . It is easy to see that and . Let be an irreducible factor of in such that . Propositions 2.16 and 2.10 imply that
Remark that
Note that . By Lemma 4.1,
(6) | ||||
Similarly, let and
Then and . Let be an irreducible factor of in such that . Applying Propositions 2.16 and 2.10 again yields that
An argument similar to the above implies that . Since , one has that and thus
Furthermore since , . By Lemma 4.1,
which together with (6) gives
Proposition 2.8 implies that
To prove the inequality in the opppsite direction, consider
Remark that and . We have the same bounds for elements in . A similar argument then implies that
Set . Then one gets the required inequalities. ∎
5 Main results
In this section, we always ssume that is irreducible over and
Pick such that . Set . Then . Set
where . Then an easy calculation yields that
As , This implies that
Then because
We claim that is irreducible over . First of all, assume that for some . Then we have that
If then the have common zeroes and none of common zeroes is zero. It is easy to see that is a common zero of all if is a common zero of all . This contradicts with the fact that . Secondly, if has a factor with positive degree in then will have a factor with positive degree in , a contradiction. This proves our claim. Remark that is a nontrivial rational solution of if and only if is a nontrivial rational solution of . The main result of this paper is the following theorem.
Theorem 5.1.
Assume that is a first order AODE with positive and assume further that is irreducible over . Then if is a rational solution of then
where .
Proof.
We shall use the notations as above. Due to the above discussion, we only need to consider the differential equation . Denote and . One sees that and
Suppose that
where is irreducible over . Since is irreducible over , one has that all are conjugate to each other and then
By Corollary 2.15, . Assume that is a rational solution of then is a rational solution of all . In particular, . Denote and . Then
Set . By Theorem 4.3 and Remark 2.7, one has that
where
Note that is the number of quadratic transformations applied to transfer to an algebraic curve with only ordinary singularities. Due to Theorem 2 in Chapter 7 of [9], can be chosen to be an integer not greater than
Remark that . Thus
(7) |
As divides both and , divides . Set . Then
This together with (7) implies that
The second inequality holds because . ∎
Remark 5.2.
Theorem 5.1 implies that an autonomous first order AODE with positive has no nontrival rational solutions, because . In fact, suppose that has a nontrival rational solution. Then it will have infinitely many rational solutions. By Corollary 4.6 of [7], has no movable singularities. However, as has positive , Fuchs’ criterion implies that 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 is a differential polynomial over . Denote
If there is satisfying that and for every , then we say that is maximally comparable. The following examples shows that their algorithms can not deal with all first order AODEs with positive .
Example 5.3.
Let
where . Then . Since but , is not maximally comparable. On the other hand, we have that
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 th 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.