On fewnomials, integral points, and a toric version of Bertini’s theorem

By Clemens Fuchs, Vincenzo Mantova, and Umberto Zannier

Abstract

An old conjecture of Erdős and Rényi, proved by Schinzel, predicted a bound for the number of terms of a polynomial when its square has a given number of terms. Further conjectures and results arose, but some fundamental questions remained open.

In this paper, with methods which appear to be new, we achieve a final result in this direction for completely general algebraic equations , where is monic of arbitrary degree in and has boundedly many terms in : we prove that the number of terms of such a is necessarily bounded. This includes the previous results as extremely special cases.

We shall interpret polynomials with boundedly many terms as the restrictions to 1-parameter subgroups or cosets of regular functions of bounded degree on a given torus . Such a viewpoint shall lead to some best-possible corollaries in the context of finite covers of , concerning the structure of their integral points over function fields (in the spirit of conjectures of Vojta) and a Bertini-type irreducibility theorem above algebraic multiplicative cosets. A further natural reading occurs in non-standard arithmetic, where our result translates into an algebraic and integral-closedness statement inside the ring of non-standard polynomials.

1. Introduction

This paper is concerned with algebraic equations involving fewnomials, also sometimes called sparse, or lacunary polynomials. By this we mean that the number of terms is thought as being fixed, or bounded, whereas the degrees of these terms may vary, and similarly for the coefficients (though they are sometimes supposed to be fixed as well).

This context traces back to several different viewpoints and motivations. For instance, there are issues of reducibility (as in the well-known old theory of Capelli for binomials, and in more recent investigations for -nomials, e.g. by Schinzel Reference 18). Sparse polynomials also occur when thinking of complexity in writing down an algebraic expression; see for instance Davenport’s paper Reference 8 (which also mentions issues related to the ones considered below). In turn, low complexity affects important geometrical or topological aspects (as in Khovanskiĭ’s theory Reference 13).

One perspective and series of relevant questions appeared when Erdős and Rényi raised independently the following attractive conjecture: Suppose that is a (complex) polynomial such that has at most terms. Then the number of terms of is bounded dependently only on Reference 9. It turned out that this problem was not innocuous as it might appear; indeed, for infinitely many ’s the number of terms of may be much larger than that of , in fact for a , as was pointed out by Erdős himself Reference 9Reference 18.

The conjecture was proved by Schinzel Reference 17, actually for for any given . Schinzel also extended the conjecture to compositions for any given , which could not be dealt with by his methods. In turn, this was settled in Reference 20.

1.1. Main results

One of the main purposes of the present paper is to achieve a “final” result in the said direction, by treating general algebraic equations of the form , assuming that is a “fewnomial” in and has an arbitrary degree in ; we then seek a bound for the number of terms of . We shall indeed prove that such a bound exists and that it is actually uniform in the coefficients of , recovering the above mentioned conclusions related to the Erdős-Rényi conjecture (in sharper form) as very special cases. For instance, we prove the following.

Theorem 1.1.

Let have terms in and be monic of degree in . If satisfies , then has at most terms.

The Erdős-Rényi conjecture is re-obtained on taking and also Schinzel’s subsequent conjecture with (moreover uniformly in the coefficients of ).

Results of this type are strongly related to other (apparently far) issues of arithmetic and geometric nature, as we now illustrate. First, we remark that a convenient point of view, adopted here, is to think of a (Laurent) fewnomial as the restriction of a given regular function on a torus to a -parameter subgroup or coset. Indeed, a regular function on is just a Laurent polynomial , whereas any connected -parameter subgroup (resp. coset) may be parametrized as (resp. ) for integers (resp. and nonzero constants ). Hence, by substitution inside , we obtain a Laurent polynomial in whose number of terms is bounded independently of the subgroup or coset.⁠Footnote1

1

Naturally, a similar interpretation holds for multivariate fewnomials; however, the issues may usually be reduced to the basic case of a single variable by substitution.

In this view, the above theorem can be rephrased in the following equivalent form.

Theorem 1.2.

If is monic in and of degree at most in each variable, if are natural numbers, and if satisfies

then has at most terms.

The numbers are actually effective, although we skip the details of such calculation. This leads, as we shall see, to a complete algorithmic description of all the possible solutions . Note that moreover the bound is independent of the coefficients of , so that the conclusion remains valid if we use the substitution for some arbitrary numbers .

This viewpoint for instance suggests a generalization of the concept of fewnomial to the case of powers of abelian varieties.⁠Footnote2 A relevant result in this direction can be found in Reference 14. But, more important here, this is useful in the development of the proofs, and it also suggests a number of links with other topics. We will discuss in a moment those which appear to us more relevant.

2

None of the results of this paper are known in that case, and it seems of interest to ask whether an analogue of the Rényi-Erdős or Schinzel’s conjecture is true in that context, already replacing with an elliptic curve.

We also point out the following dichotomy between polynomials and rational functions: we shall state a version of Theorem 1.2 for rational functions (see Theorem 2.2), where we drop the assumption that is monic, and the conclusion will say that can be written as the ratio of two polynomials with a bounded number of terms that are not necessarily coprime. An instance of this behavior is the cyclotomic polynomial , which solves the equation without being a fewnomial, but that in fact can be written as (observe that the equation is indeed not monic). This phenomenon is intrinsic to the problem, and in fact many results here will be stated twice to account for both polynomials and rational functions.

The suitable statement for rational functions can be deduced straight away from Theorem 1.2; however, in this paper we shall actually proceed in the opposite direction, first proving a theorem for rational functions, and then recovering Theorem 1.2 (see the final argument in Section 8).

1.2. Integral points on varieties over function fields

Many attractive Diophantine problems concern the -integers and the -units in a number field .⁠Footnote3 The latter may also be described as just the -integral points for . For instance, the Mordell-Lang conjecture for tori yields a description of the -integral points on subvarieties of , that is the points on with -unit coordinates. Such a description follows from the -unit theorem of Evertse, Schlickewei, and van der Poorten, while the general conjecture for tori has been a theorem of Laurent since the 1980s; see Reference 2, Theorem 7.4.7.

3

We recall that ; for instance, for , the -units are those rationals with numerator and denominator made up only of primes in the finite set .

Instead, much less is known for -integral points on finite covers of (except for the case of curves). Take for instance the simple-looking equation , to be solved with and . This represents a double cover of , on which we seek the -integral points. Alternatively, they may be described as the -integral points for the affine variety obtained as the complement in of two lines and a suitable conic (see Reference 5). Now, this is a divisor of degree with normal crossings, so a celebrated conjecture of Vojta predicts that the solutions are not Zariski dense, but this has not yet been proved (see Reference 2, Section 14.3; this special case was proposed explicitly by Beukers in Reference 1).⁠Footnote4

4

This is indeed a “borderline” case of Vojta’s conjecture on integral points, one of the simplest but yet unsolved ones. See Reference 5 for a proof in the function field context.

A related form of this problem has recently been proposed by Ghioca and Scanlon while studying the dynamical Mordell-Lang conjecture in positive characteristic. Specifically, for a given prime , they ask about the integer solutions of , in the unknowns , where the polynomial and the constants are given. Since are -units, this is in turn a special case of seeking the integral points on the cover of given by .

The methods so far known do not suffice even to treat the former equation (see Reference 7 for some special cases). Actually, the problem arises even in writing down what is expected to be the most general form of the solution. Note that any identity of the shape , for a polynomial , would produce solutions simply by setting . Hence, it is a primary task to write down all such identities. Note also that such an identity (considered now over ) represents an -integral point on the said cover, but now relative to the function field and set Footnote5: in fact, the -units of are precisely the monomials .

5

Here, in accordance with quite a general principle, the integral points over a function field may be used to parametrize the integral points over a number field.

This example makes evident the connection of these topics on integral points with the topic of fewnomials (and with Erdős-Rényi and Schinzel’s mentioned conjectures); indeed, in the case of the problem of Ghioca and Scanlon a complete description in finite terms of the relevant identities follows from Theorem 2 of Reference 20.

The results of the present paper yield a corresponding description in a rather more general situation. Namely, in dealing with an arbitrary finite cover , they allow us to parametrize all the regular maps (i.e., the -integral points on , with respect to the function field and set ).⁠Footnote6

6

The case of more general function fields or even more general sets is not known to us and seems to present subtle difficulties; this happens already by taking . See Reference 6 for some cases related to surfaces.

Theorem 1.3.

Let be a finite map. Then there exists a finite set of regular maps , with an integer and an affine algebraic variety, such that for every regular map there exist a , a point , and a regular map with .

Here denotes the restriction of to . The special case of this theorem appears as Theorem 5.1 in Reference 4, in different phrasing and with a completely different (and somewhat involved) proof.

We therefore see that any -integral point” factors through a map of bounded degree, in the sense that the inverse image of a hyperplane section of has a bounded degree in . This can be expressed in terms of boundedness of the heights of the integral points. Such a conclusion, which is in a sense the best possible, proves Vojta’s conjectures for and the integral points in question.⁠Footnote7 As an application, we can prove the following corollary.

7

See e.g. Reference 2, Section 14 for a general formulation of Vojta’s conjectures, especially over number fields. For brevity we omit here any further detail or example.

Corollary 1.4.

Suppose that the union of the images of all regular non-constant maps is Zariski dense. Then the branch locus of in is invariant by translation by an algebraic subgroup of positive dimension.

This is a useful condition which fits within a classification of Kawamata (see, for instance, the remark after Theorem 2 of Reference 6 or Section 5.5.5 of the recent book by Noguchi and Winkelmann Reference 16).

Moreover, this language also makes it more obvious how to prove the special case (we thank one of the anonymous referees for pointing out this argument). Indeed, for an integral point is a regular map such that the composition is the map , that is an isogeny composed with a translation. Suppose that one such point exists. It follows easily that the normalization of is in fact isomorphic to itself, and all the integral points can then be easily classified.

1.3. A “Bertini Theorem” for covers of tori

Consider again a (ramified) cover , by which we mean a dominant map of finite degree from the irreducible algebraic variety . When is replaced by the affine space , a version of the Bertini irreducibility theorem asserts that for , if is a “general” hyperplane in , the fiber is still irreducible. In the present context one may replace by a general algebraic subgroup (or coset) of and ask about the same conclusion. Of course, a marked contrast with the Bertini case is that the algebraic subgroups now form a discrete family, which prevents standard methods from working in this context. In Reference 21, Theorem 3 a positive result was obtained, however, concerning irreducibility only above components of 1-parameter subgroups, and not above arbitrary cosets.

Now, the arguments and results of this paper (completely independent of Reference 21) directly lead to a toric analogue of Bertini’s theorem without the said restriction.

Theorem 1.5.

Let be a quasi-projective variety and be a (complex) dominant rational map of finite degree , and suppose that the pullback is irreducible. Let be a proper algebraic subvariety such that is finite onto its image. Then there exists a finite union of proper algebraic subgroups of such that if is a connected algebraic subgroup not contained in , then for all , is irreducible.

Note that if is already finite onto its image, then may be omitted from the statement. However, as pointed out by an anonymous referee, to whom we are grateful for the correction, in the general case the subvariety must be included in the statement; for instance, if is the blowup of at a point, then the preimage of a coset passing through the point always contains the exceptional divisor. The hypothesis of irreducibility of the pullback is also a necessary condition.⁠Footnote8

8

For instance, when is an isogeny of , the cover becomes reducible above every subgroup , for any torus not containing the kernel of , since .

As for Theorem 1.2, the set can be given a complete algorithmic description, which is rather uniform in the data. Consider the following particular case. Suppose that the variety can be represented as the hypersurface , with given by the projection on the first coordinates. Assume moreover that is a (Laurent) polynomial in the ’s and monic in . Under these assumptions, the map is finite, and the conclusion of Theorem 1.2 gives the following strengthening.

Addendum to Theorem 1.5.

If is the hypersurface defined by , where is a Laurent polynomial in and monic in , and is the projection onto the first coordinates, then and the set may be chosen dependently only on .

As an application, we immediately obtain the following corollary, in which for a given integer we let denote the Kronecker substitution .

Corollary 1.6.

Let be a complex polynomial of degree in and such that is irreducible over . Then is irreducible over for all integers large enough in terms of .

This had been obtained in Reference 21 (with a completely different proof), however, without this uniformity, which was left as an open question.

1.4. An application to composite rational functions

One may propose an analogue for rational functions of the already mentioned conjectures of Erdős-Rényi and subsequent ones by Schinzel. Namely, let be a given rational function and suppose that for a rational function , the composition may be written as a ratio of two polynomials (not necessarily coprime) with at most terms. Is there a such that may be represented as the ratio of the polynomials with at most terms? The present methods allow a positive solution of this problem as well, as follows.

Theorem 1.7.

If are such that the composition can be written as the ratio , where have altogether at most terms, then there exist polynomials with at most terms such that .

Again, we stress that the pairs and are not necessarily coprime. We also remark that we actually have full uniformity here in the rational function , as the number only depends on and not on (this dependency can be removed thanks to a previous theorem proved by the first and last authors Reference 12; in the same paper, the function is also described, on bounding its degree in terms of only, unless is of certain special shapes listed therein).

1.5. Non-standard polynomials

The notion of fewnomial and our main theorems can be translated naturally in the language of Robinson’s non-standard analysis. We refer the reader to Reference 10 for an introduction to the subject.

Here we just recall that in non-standard analysis one has a map which sends the standard objects, such as or , to their non-standard counterparts, in a way that preserves all first-order formulas. The easiest example of (non-trivial) map is the one that sends any set into the set of sequences with values in (i.e., ) modulo the equivalence relation defined by a fixed non-principal ultrafilter on (i.e., if is in the ultrafilter). This introduces new, non-standard elements; for instance, the non-standard contains an element , the equivalence class of the sequence , which is different from any standard natural number.

Concerning our context, we note that the non-standard contains “polynomials with infinitely many terms,” such as

In fact, this is exactly the equivalence class of the sequence .

We now define the ring of fewnomials in to be the subring of polynomials whose number of terms is actually finite,

In this language, statements about fewnomials become quite compact. As an instance of this phrasing, the Erdős-Rényi conjecture proved by Schinzel becomes the following: if for some , then . Likewise, Theorem 1.1 translates to the following quite short statement.

Theorem 1.1.

The ring is integrally closed in .

This statement was proposed by Fornasiero before the results of this paper, together with its following immediate corollary (which is, in turn, a non-standard translation of Theorem 2.2, stated in the following section).

Corollary (Theorem 2.2).

The fraction field of is relatively algebraically closed in .

The latter conclusion is another example of the aforementioned behavior of rational functions, and indeed it corresponds to dropping the assumption that the polynomial is monic in Theorem 1.1.

It is rather easy to see that Theorems 1.1 and 1.1 are indeed equivalent. For example, assume Theorem 1.1 and suppose by contradiction that Theorem 1.1 is false. Then for some there should be a sequence of polynomials whose number of terms grows to infinity, while they also satisfy

where is a sequence of polynomials with at most terms, of degree at most , and monic in the last variable.

But then the equivalence classes and of the above sequences satisfy

which means that is integral over , while it lies in and not in , a contradiction.

Although there are details to be worked out, we believe that also our proof of Theorem 1.1 can be translated rather naturally to a shorter argument in the non-standard language; this is being investigated and may appear in a future paper. On the one hand, we would lose effectivity, but on the other, we may be able to avoid the use of the resolution of singularities and directly use the construction of the Puiseux series.

The main potential simplification comes from the fact that many notions, which in the proof depend on appropriately chosen parameters, become absolute. For example, the notion of being “small” with respect to a “large” number, which in our proof depends on a parameter to be chosen carefully, translates to being infinitesimal with respect to the second number.

1.6. Fewnomials and unlikely intersections

This instance does not directly use results of the present paper, but we still discuss it because it is far from being unrelated.

As already mentioned at the end of Section 1.1, several results here contain a dichotomy lacunary polynomials lacunary rational functions, where by the latter terminology we mean rational functions which may be represented as ratios of fewnomials, possibly non-coprime, as in Theorem 1.7. Recall the standard example , which shows that a lacunary rational function which is a polynomial is not necessarily a fewnomial. This gives rise to the following problem, also posed independently by Zieve.

Suppose that a rational function can be represented as the ratio , where the integers vary, while are fixed coprime polynomials in . (In accordance with the viewpoint illustrated above, we are viewing as the restriction of a fixed rational function on to a one-dimensional algebraic subgroup which may vary.) One may ask the following question.

Question 1.8.

For which one-dimensional algebraic subgroups does become a (Laurent)polynomial?

For instance, the above example comes from , on ; in this case it is easy to check that the only one-dimensional algebraic subgroups which make a (Laurent) polynomial are given by for integer (as in the example).

Therefore, in particular, we have two coprime polynomials such that they become non-coprime (or such that becomes invertible) along the one-dimensional subtorus of parametrized by . This kind of problem also appeared in a conjecture of Schinzel, which was later recognized as a special case of the more recent Zilber-Pink conjecture in the realm of the so-called unlikely intersections. See Reference 22 for a discussion of this topic, especially Chapter 2. This conjecture of Schinzel was confirmed by Bombieri and the third author (see Reference 18, Appendix), and it was later refined with other methods, in collaboration also with Masser, in Reference 3, Theorem 1.5, in a work proving the Zilber-Pink conjecture for intersections with one-dimensional subgroups.

These last results give an answer to the above question, showing that the relevant algebraic subgroups are contained in a finite union of proper algebraic subgroups of . Given this, one may restrict to the subgroups in and continue by induction to write down all the possibilities: it turns out that the relevant one-dimensional algebraic subgroups are precisely those contained in a certain finite union of proper algebraic subgroups on which becomes regular.

It is to be remarked that the more general question in which one-dimensional algebraic subgroups are replaced by one-dimensional algebraic cosets does not admit a similar solution. This corresponds to the ratio being a polynomial, for integers and nonzero constants . We do not know of any method able to deal with such a question in full generality.

Another connection to integral points was pointed out to us by one of the referees, and we are grateful for it. Starting from the rational function , one can blow up the codimension-two subvariety of defined by the simultaneous vanishing of and , and remove the strict transform of the subvariety of defined by . Let be the resulting variety. Then the one-dimensional (translates of) algebraic subgroups which are solutions to Question 1.8 correspond to regular maps . With this interpretation a solution for the problem of translates can possibly be given for (the case of surfaces) under some normal-crossings conditions which will depend on and are generically satisfied.

1.7. Proof methods and quantitative issues

The strategy of the proofs here follows only in part the pattern of Reference 20; this shall be outlined in more detail in Section 3 (before the formal arguments). The main technical issue is finding an appropriate way of expanding as a kind of multivariate Puiseux series. This is done here by using first the theory of resolution of singularities to reduce to a rather regular case in which one can use multivariate analytic expansions. An earlier version of the proofs involved a different, more complicated construction of certain Puiseux-type expansions, but no use of resolution of singularities. The approach was dropped in favor of the present one for the sake of simplicity, but it may be of independent interest, and it can still be found in an earlier draft of this paper Reference 11.

A by-product is a completely effective output of the proofs: one can obtain effective estimates for the involved quantities, and effective parametrizations (provided of course one deals with cases in which the fields and the equations which occur are finitely presented). However, we do not give here explicit bounds, which in any case would have the shape of highly iterated exponentials.⁠Footnote9

9

In the original cases of the Erdős-Rényi conjecture, doubly exponential bounds had been obtained by Schinzel Reference 17, reduced later to a single exponential by Schinzel and the third author Reference 19.

2. Variations and reductions

2.1. Variations of Theorem 1.2

The following three statements are variations regarding irreducible factors and the dichotomy rational functions polynomials mentioned in the Introduction.

The first one concerns factorizations.

Theorem 2.1.

If is monic in and of degree at most in each variable, if are natural numbers, and if are polynomials monic in such that

then each coefficient of (as a polynomial in ) has at most terms.

(By symmetry, a similar conclusion holds automatically for the coefficients of .)

Note that we recover Theorem 1.2 on taking as the first factor. The converse deduction is also not difficult but shall be explained later. The other variations concern rational functions.

Warning.

We stress again the point that whenever we write a rational function (even when it is a polynomial) as a quotient of two polynomials, we are not usually assuming that the numerator and denominator are coprime (recall the example ). This issue is related to not requiring that is monic in , as in the following statements.

Theorem 2.2.

If is a polynomial of degree at most in each variable, if are integers, and if is such that

then is the ratio of two polynomials in with at most terms.

Theorem 2.3.

If is a polynomial of degree at most in each variable, if are integers, and if are such that

with monic in , then each coefficient of (as a polynomial in ) is the ratio of two polynomials in with at most terms.

It is easy to see that Theorem 1.2 implies Theorem 2.2, but the converse deduction does not appear as straightforward (we stress yet again the point that the polynomial in the conclusion of Theorem 2.2 is represented as a quotient of two polynomials which need not be coprime). In this paper, we actually prove Theorem 2.2 first, and then deduce Theorem 1.2 (and Theorem 2.1) via a general integrality argument.

Remark 2.4.

In all of the above statements, we may actually allow to be negative and , with a similar conclusion.

We may also deduce that the fewnomials which arise can be parametrized with the same exponents. For instance, in Theorem 1.2, we can say that there are and , with and bounded in terms of and only, such that .

For the sake of simplicity, we shall omit details about these further assertions.

2.2. Reductions

We are going to prove Theorem 2.2 first, and we can use some standard arguments to reduce the theorem to a simpler situation. In a moment, we shall reduce both Theorems 2.2 and 2.3 about rational functions to the case where is monic in , and are non-negative. We obtain the following statements, in which the assumption is as in Theorem 1.2 (or Theorem 2.1), but the conclusion is as in Theorem 2.2 (resp. Theorem 2.3).

Proposition 2.5.

If is monic in and of degree at most in each variable, if are natural numbers, and if is such that

then is the ratio of two polynomials in with at most terms.

Note that since is monic in , it follows that is actually a polynomial, but the conclusion only says that is represented by a quotient of two polynomials which need not be coprime. Thus this proposition is a weak form of Theorem 1.2. It is similar for its corollary.

Proposition 2.6.

If is monic in and of degree at most in each variable, if are natural numbers, and if are such that

with monic in , then each coefficient of as a polynomial in is the ratio of two polynomials in with at most terms.

Both are clearly special cases of the original Theorems 2.2 and 2.3. As we now show, it is not difficult to deduce the latter statements from them.

Note 2.7.

It is important to note that the following deductions are valid for each single (whereas the number is changed in the course of the deductions). This is crucial in all our proofs that proceed by induction on ; namely, if we assume that one statement is true for a certain value of and all possible ’s, the other statements will follow as well for the same value of and all possible ’s.

Deduction of Theorem 2.2 from Proposition 2.5.

Note first that Proposition 2.5 requires to be natural numbers rather than integers. We may reduce to the case by replacing, when necessary, by and multiplying the resulting polynomial by ; after this transformation, the degree of in each variable is still bounded by . Therefore, we may assume that for all .

Write as

where the ’s are polynomials of degree at most in each variable. Let be the maximum integer such that is not identically zero, and let .

We now consider the polynomial . Note that is monic in , and it has degree at most in each variable. Assuming Proposition 2.5, each rational root of is the ratio of two polynomials with at most terms. Multiplying each such root by we obtain all the rational roots of , and therefore the rational solutions of (Equation 2.2). In particular, the solutions are ratios of polynomials with at most terms, as desired.

Deduction of Theorem 2.3 from Proposition 2.6.

We proceed as in the previous deduction to show that is a suitable value for .

Moreover, as promised earlier, we can easily deduce Theorem 2.3 from Theorem 2.2. Thanks to the above reductions, it is sufficient to deduce Proposition 2.6 from Proposition 2.5.

Deduction of Proposition 2.6 from Proposition 2.5.

Suppose that is a coefficient of a monic irreducible factor of the polynomial,

Let us call the roots of this polynomial in an algebraic closure of , with repetitions, where . The polynomial is, up to sign, an elementary symmetric polynomial in some of the roots. Let us denote the elementary symmetric polynomials as .

Up to reordering the roots, we may write

for some . This implies that , up to sign, is a root of the monic polynomial

But the coefficients of are now symmetric polynomials in the roots , which implies that they are actually polynomials in the ’s, i.e., the coefficients of . A rough estimate shows that the degree of each such polynomial in each is bounded by .

This implies that we may find monic in and of degree at most in each variable such that

Assuming Proposition 2.5, since is a root of , it must be a ratio of two polynomials with at most terms, as desired.

The exact same argument can also be used to show that Theorem 2.1 follows from Theorem 1.2.

Deduction of Theorem 2.1 from Theorem 1.2.

We proceed as in the previous proof to show that is a suitable value for .

2.3. Further lemmas

In the course of our proof, we will need on a few occasions to replace with an auxiliary variable such that . In the next lemma, we show that these substitutions do not affect our statements, so they may be considered as immaterial.

Lemma 2.8.

Let be a polynomial such that can be written as the ratio of two polynomials with at most terms. Then is the ratio of two polynomials with at most terms.

Proof.

Suppose that , where and are polynomials with at most terms. Grouping the monomials whose degrees in are in the same congruence class modulo we may (uniquely) write

with , polynomials with at most terms as well.

But then, since , we must have for all , and in particular . As at least one is non-zero, we have found a representation of as the ratio of two polynomials with at most terms, as desired.

Another easy reduction shows that if we find a -linear relation with bounded coefficients between the exponents , then we may actually remove one of the exponents. This is also crucial for our induction on .

Lemma 2.9.

Suppose that we are under the hypothesis of Theorem 2.2, and that there are integers , not all zero, and some such that

Assume moreover that Theorem 2.2 has been proved for and any degree . Then is the ratio of two polynomials with at most terms.

Proof.

Without loss of generality, we may assume that . In this case, we take new variables , we replace in with for and with , and we multiply the result by . The resulting polynomial has degree at most in each variable, and it vanishes at and .

Now, using the assumption about Theorem 2.2 and Lemma 2.8, is the ratio of two polynomials with at most terms.

3. Introduction to the proof

In order to prove Theorem 2.2, we build up on the same technique of Reference 20 but with the additional use of the theory of resolution of singularities to reduce to a sufficiently regular case. Indeed, the underlying expansions depend not quite on the variable , but on the variables ; it is well known that expansions of algebraic functions of several variables often depend on subtle geometric features.

For the sake of illustration, we explain the strategy of the proof in a simpler example where this combinatorial aspect is missing. We work by induction on .

Say that, as in the original Erdős-Rényi conjecture (a special case of Theorem 1.2), we start with the polynomial

For simplicity, we also assume that .

If we want to prove that a rational root of

is the ratio of two polynomials with few terms, we may expand with the binomial series; namely, letting , we may easily obtain the multinomial expansion

It is crucial that run through natural numbers. Assuming that , if for some fixed , each exponent is at least . Since the degree of must be , we find that all terms must eventually cancel except possibly for those such that , leading to the bound for the number of terms.

This consideration always works for (with ), and in particular we obtain the base case of our induction. However, in general we have no lower bound at all for . To cope with this difficulty, the principle in Reference 20 is that if some terms are very small compared to , we can group together these small contributions as follows: we define

and we expand as

As before, we can expand the powers of , which involve the large exponents only; however, the new coefficients will not be constants, as before, but actually functions in the hyperelliptic function field . Despite this radically new feature, a theorem in Diophantine approximation over function fields (see Section 6) allows one to reduce to the inductive hypothesis at , provided is large enough, by which we mean that it is greater than for an absolute . Of course, for some we must indeed have that is small, whereas is large, concluding the argument.

In the general case, we wish to apply the same approximation technique. However, a direct attempt at expanding as a kind of multivariate series might fail; The main issue is that we may have monomials involving exponents that are combinations of with negative coefficients, in which case a combination of large exponents may become small, and it is not as easy any more to separate the big ones from the small ones. These obstacles appear when is a non-simple root of .⁠Footnote10

10

These issues are entirely avoided in the cases considered in Reference 20, where multinomial expansions suffice.

We shall overcome these obstacles by applying a suitable monoidal transformation to our original equation. Although might still be a non-simple root of , the transformation will guarantee that can still be expanded as in the original case. The choice of the monoidal transformation relies on the theory of resolution of singularities.⁠Footnote11

11

We recall that an earlier draft of this paper contained a different proof based on a careful construction of a Puiseux-type expansion rather than the resolution of singularities Reference 11.

4. Reduction to the regular case

In order to obtain our desired expansion of as a “pseudo-analytic series,” we prove that Proposition 2.5 can be further reduced to a special case in which the polynomial is sufficiently regular.

Let be as in Proposition 2.5. Assume, as we may, that is irreducible. Let be the function field generated by the independent variables and an algebraic function such that .

Let be a projective non-singular model of the function field . For each , let be the set of the irreducible components of the divisor of , and let . By the known theory of resolution of singularities, after applying some blowups, we may further assume that the divisors appearing in are non-singular and have normal crossings.

Let be the unique non-constant map such that

for all ;

,

where is the standard coordinate function . Let .

Definition 4.1.

Under the above notations, we say that a solution to (Equation 2.2), namely , is regular if are local parameters at .

Note that one might reformulate the above notion in a more compact and geometric way by only referring to the map . However, we prefer to keep an explicit reference to the polynomial and the numbers .

The purpose of this section is to show that it suffices to prove the conclusion of Proposition 2.5 in the special case in which the solution is regular.

In what follows, given a divisor and a regular function , we let denote the order of at . Note that since has degree at most in each variable, we have for all .

Lemma 4.2.

Let be the divisors in on which lies. Then either

for some integers not all zero such that or the matrix is invertible (and in particular ).

Proof.

Since the divisors in have normal crossings, we know at once that . Assume that the matrix is not invertible; otherwise we are done. Then there are integers , not all zero, such that

for all . Since for all , we may choose the integers so that by Siegel’s lemma.

Let . By the above observation, none of is a component of the divisor of . On the other hand, the components of the divisor of are in . It follows that no such component contains , so is regular at and . Note moreover that . Therefore,

Since , it immediately follows that , reaching the desired conclusion.

Thanks to the above observation, in order to prove Proposition 2.5, it shall be sufficient to prove the following special version which has a few more hypotheses.

Proposition 4.3.

If is monic in , is irreducible, and is of degree at most in each variable, and if and form a regular solution of

then is the ratio of two polynomials in with at most terms.

Note 4.4.

As in Note 2.7, the following deduction is valid at every single .

Deduction of Proposition 2.5 from Proposition 4.3.

We work by induction on . In particular, if for some , we may specialize the variable to ; if , we conclude by inductive hypothesis, while if , we simply note that we actually have . Therefore, we may assume that for all . Moreover, we may replace by an irreducible factor, and therefore assume directly that is irreducible.

By Lemma 4.2, either lies on distinct divisors such that the matrix is invertible, or there is a relation with integers not all zero and such that . In the latter case, we must have , and we may conclude by Lemma 2.9 and the inductive hypothesis. Therefore, we may assume to be in the former case.

Let be new independent variables, and set

Note that is a polynomial of degree at most in each variable.

Let be the inverse matrix of multiplied by its determinant , so that its coefficients are all in . Let . By construction, we have

In turn, we choose an irreducible factor of such that

We claim that form a regular solution of this equation in the sense of Definition 4.1. Indeed, we may now assume that are algebraic functions in some algebraic closure of such that

Let be a projective non-singular model of the function field , equipped with a surjective, finite map . As at the beginning of the section, we may apply some blowups and assume that all the components of the divisors of the functions are non-singular and have normal crossings.

Let be the unique non-constant map such that and . Let . Since by construction , it follows at once that . For , let be a component of on which lies.

Finally, note that . In particular, can be factored as , where is a function on . By construction, we have , where is the Kronecker delta. Let be local parameters of , so that (in the sense of analytic equivalence). It follows at once that ; since the function field extension is generated by the functions , each is also a local parameter of , and in particular, is a local parameter of . Since the divisors have normal crossings, this means that are local parameters at .

In turn, form a regular solution of (Equation 4.1), so can be written as the ratio of two polynomials with at most terms. By Lemma 2.8, can also be written as the ratio of two polynomials with at most terms, concluding the argument.

5. From multivariate expansions to algebraic approximations

We now start our argument toward the proof of Proposition 4.3. From now up to Section 8, assume that we are working under the assumptions of Proposition 4.3, and in particular that and form a regular solution of , where is irreducible, is monic in , and has degree at most in each variable.

Under the notation of Section 4, regularity means that are local parameters at , which means that there is an embedding of the regular functions at into . Since moreover the function is integral over , we obtain an embedding

The fairly trivial, but crucial observation, is that for any we can also rewrite

Therefore, fix one such . We write for the vector . If is a vector of integers , we write . With this notation, the above embedding yields a (unique) expansion

where . Recall that

We now specialize the above expansion along the curve and pull it back to . Recall that the ring of functions on that are regular at the origin can be embedded (uniquely) into . Under this embedding, the specialization and the pullback simply mean that we specialize at term by term.

We first specialize at for . Since for all , each series converges at to a series , and we have

Likewise, we can further specialize at for . We then obtain the expansion

where and denotes the usual scalar product. Note that such expansion is convergent again since for all .

Equation (Equation 5.3) yields an expansion of resembling an analytic expansion, but with coefficients that are themselves functions of , providing the first ingredient toward the proof of Proposition 4.3. We now use (Equation 5.1) and (Equation 5.2) to deduce some bounds on the coefficients and .

Proposition 5.1.

The coefficients generate a finite extension of degree at most of . Similarly, the coefficients generate a finite extension of degree at most of .

Proof.

Suppose that the coefficients generate either an algebraic extension of degree greater than or a non-algebraic extension. In both cases, we can apply Galois automorphisms over to find at least distinct sequences of coefficients. In turn, such automorphisms extend naturally to by leaving fixed. Applying the automorphisms to (Equation 5.1), we find that the degree of in the last variable should be at least , a contradiction. The conclusion for the coefficients can be proved with a similar argument applied to (Equation 5.2) (just recall that is monic in , so it does not become trivial when specializing ).

For , let be the field generated by over , where is the -norm of , and . By Proposition 5.1, . Similarly, let be the field generated by over , and . Again, . Let be the logarithmic height of the function field normalized so that .

Lemma 5.2.

Let . Let be an irreducible polynomial such that

If , then the degree of in each variable is at most for a suitable .

Proof.

This is a classical result, although usually only stated for the case , and either with or with additional assumptions on the derivative in of the polynomial . For the sake of completeness, we sketch an argument that reduces the general case to , .

Suppose and . In this case, the vector is just a single natural number and the degree of in coincides with the logarithmic height of in the function field (upon choosing an appropriate normalization). Then the height of is at most for a suitable ; moreover, there is a finite set of places , whose size can be bounded in terms of only, such that all the coefficients are -integral. (See for instance Reference 15, Lemma V.5, with the additional observation that the order of at can also be bounded in terms of only.)

The general case can be reduced to the above special case by specializing for , and for , where are algebraically independent over the field of definition of . If is the specialization of , the coefficient of in the specialized expansion is

where . By the previous argument, the height of is bounded by , and is -integral for a suitable of size bounded in terms of only. Since is contained among the poles of the functions , upon varying one recovers that does not depend on . Using sufficiently many independent values of , one can eliminate and obtain that the height of each is bounded by . It now suffices to note that

satisfies , and it is the product of at most conjugate irreducible factors (using Proposition 5.1), so . Since the degree of in coincides with , the conclusion follows at once.

Proposition 5.3.

For all , there is a primitive element of such that, if is an irreducible polynomial such that

then the degree of in each variable is at most . Moreover, we may assume that is monic in .

Proof.

By a classical argument of Galois theory, for sufficiently generic coefficients , the element

generates over . The conclusion then follows by Lemma 5.2 and some elementary algebra.

Proposition 5.4.

For all , if , then either or there are integers , not all zero, such that and .

Proof.

By Lemma 5.2, we have

where has degree at most in each variable. Therefore, the specialized polynomial has degree at most in . If , then is bounded by the height of the polynomial , and the first conclusion follows by an easy estimate of such height.

Otherwise, if vanishes, then (at least) two distinct terms of become terms of the same degree in when specialized. This immediately implies the second conclusion.

Corollary 5.5.

For all , either the genus of is bounded by for some or there are integers , not all zero, such that and .

Proof.

This follows at once from the estimate of Proposition 5.4 and the basic theory of function fields (for instance, by counting the ramification points).

Remark 5.6.

Recall that for all sufficiently large integers . One can prove that this happens for all integers larger than a number dependent on only, showing that the above bound can be made independent of (as for Lemma 5.2, this is usually proven only with additional assumptions on ). However, we will not need this additional uniformity.

6. Diophantine approximation

Now that we have found a suitable expansion of as a convergent sum of algebraic functions, we proceed as in Reference 20. Recall the following lemma.

Lemma 6.1 (Reference 20, Proposition 1).

Let be a function field in one variable, of genus , let be linearly independent over , and let . Let be a finite set of places of containing all the poles of and also all the zeros of . Further, put . Then

where .

A rather straightforward application of the above lemma to (Equation 5.3) yields the following (using the notations of Section 5).

Proposition 6.2.

Suppose that and that for some given . Then at least one of the following holds:

(1)

is -linearly dependent on the set ,

(2)

for some ,

(3)

there are , not all zero, such that and .

Proof.

We apply Lemma 6.1 with the following data.

Let . Let be a -linear basis of the set , and let (note that is at most the number of vectors such that , which can be bounded in terms of and only). Let . Let be the set of zeros and poles of the functions and of the function . Let be the place at induced by the embedding of into . Thanks to the observations of Section 5, either we get conclusion (3) or we have the following bounds:

the zeros and poles of each are at most by Proposition 5.4;

the zeros and poles of , which include the poles of , are at most by Proposition 5.1;

the genus of is at most by Corollary 5.5;

by the expansion (Equation 5.3).

Finally, note that for all , is non-negative, and that by simple degree considerations, . In particular, .

Applying Lemma 6.1 and using these bounds, we either reach conclusion (1), or we obtain the following inequality:

which in turn yields

proving conclusion (2).

7. The case of linear dependence

Note that outcome (1) of Proposition 6.2 is that is -linearly dependent on for a certain . In this section, we study what happens when this is the case.

Let . Fix to be a primitive element of given by Proposition 5.3, with the corresponding irreducible polynomial . Let also . Then for all we can write

where .

Lemma 7.1.

The degree of in each variable is at most for some .

Proof.

Let be an embedding of into an algebraic closure of . Then

Since the matrix is invertible, the desired bound follows from the bound of Proposition 5.3 and some elementary algebra.

Proposition 7.2.

Suppose that is -linearly dependent on . Assume that Proposition 4.3 is true for . Then either for some integers , not all zero, such that , or can be written as the ratio of two polynomials with at most terms.

Proof.

Assume first that for some with , is not well defined, by which we mean that some denominator of vanishes on . In turn, two distinct terms of such a denominator must have the same degree when specialized, which means that

for some integers such that for all , and for at least one . We thus reach the former conclusion.

Otherwise, we specialize at for , which is possible since is by construction integral over . This yields a . By the above argument, we can also specialize each , yielding the following:

In particular, is a primitive element of over .

By assumption of linear dependence, we have

for some numbers .

Let . If , then are -linearly independent, so we must have

and the latter conclusion follows.

Otherwise, if are not -linearly independent, it means that . Note that is a root of

By hypothesis, we may assume that Proposition 4.3 holds for , and in particular we may apply Theorem 2.3. It follows that the irreducible factor of of which is a root has coefficients that can be written as ratios of polynomials with at most terms. In turn, we may rewrite each power , for , as

where each can be written as the ratio of two polynomials whose number of terms is bounded in terms of and only.

Therefore, we have

Since are -linearly independent, we may now conclude as in the case .

8. Proof of the main theorem

Finally, we can prove Proposition 4.3. We shall then prove that it implies Theorem 1.2.

Proof of Proposition 4.3.

First of all, we may directly assume that : indeed, we may simply rearrange as required.

We then work by primary induction on and secondary reverse induction on . Our inductive hypothesis at stage reads as follows:

(1)

either is the ratio of two polynomials with at most terms

(2)

or for some .

Note that in the case , the conclusion is trivial: the expansion (Equation 5.3) is just

where . Since the degree of in is at most , by simple degree considerations, is a polynomial with at most terms, reaching conclusion (1).

Similarly, in the case , the expansion is of the type

where . Since again the degree of in is at most , the only terms appearing on the right hand side satisfy . On the other hand, . Therefore,

In turn, this implies that the number of terms of can be bounded in terms of and , so in terms of and only, reaching again conclusion (1).

We now wish to prove the general case for arbitrary . Assume either that or that we have already proven stage . Proposition 6.2 yields three possible conclusions: in the first case, we obtain that is -linearly dependent on for a suitably chosen ; in the second case, we reach conclusion (2) straightaway; in the third case, we obtain that for some integers of bounded size, in which case we reach conclusion (1) by Lemma 2.9 and the inductive hypothesis.

For the first case, we then apply Proposition 7.2, and either we reach conclusion (1) immediately or we find again that for integers of bounded size, so we reach conclusion (1) again by Lemma 2.9 and the inductive hypothesis. This concludes our induction.

Chasing back the series of deductions, this finally proves Theorem 2.2. The proof of Theorem 1.2 now follows the same argument found in Reference 20.

Proof of Theorem 1.2.

By Theorem 2.2, a rational function such that

can always be written as the ratio of two polynomials, say and , with at most terms.

As in Reference 20, we may exploit this information to show that we may explicitly parametrize all such polynomials , . Indeed, for , let be the maximum absolute value of its entries; for , if

and, writing for ,

we have that

We now expand all the involved products to get monomials of the shape , where is a monomial in the coefficients and , and is a positive -linear combination of the exponents and . In order to satisfy (Equation 8.1), we can recognize two types of conditions.

(I) The first type concerns the exponents of . We can partition the monomials by grouping the ones with the same . For each set of the partition, the corresponding expressions of must have the same value, producing several vanishing homogeneous linear forms with integer coefficients in the . Note that the coefficients of such linear forms are bounded in terms of only. Moreover, since the number of possible partitions is bounded in terms of and , there is a bound on the number of resulting linear equations.

(II) For a fixed partition of the monomials with the same as in (I), the sum of their coefficients must be zero. This yields an affine algebraic variety whose coordinates correspond to the coefficients .

Each solution of (Equation 8.1) yields a solution to a linear equation as in (I) and a point on the corresponding algebraic variety given in (II). Vice versa, each solution to a linear equation as in (I) and a point on the corresponding algebraic variety in (II) yield two polynomials satisfying (Equation 8.1).

Suppose now that we fix a set of linear equations as in (I), given by a partition of the exponents, and a point in the algebraic variety found in (II), but we let the exponents vary among all the possible solutions. Since the (vector) solutions of such a system of linear equations span a subgroup of , we may in fact find a -basis, say with elements, whose entries are bounded only in terms of and ; we may then write each solution as linear combinations of these basis vectors, with integer coefficients . After this substitution, we may rewrite the resulting polynomials and as

and as

where , , and are certain Laurent polynomials in . Note that moreover their degrees are bounded in terms of the basis vectors and hence may be bounded in terms of and only.

Now, the equality

is satisfied for all in , and therefore we actually have that

Since is monic in , this implies that is integral over and therefore it is a Laurent polynomial in ; moreover, the number of terms as a Laurent polynomial is bounded dependently on and because the degree of is likewise bounded.

Therefore, since any satisfying (Equation 1.1) can be obtained using the above procedure, we have that must be a Laurent polynomial in with a number of terms bounded dependently on and . Now, since is integral over , then all of its monomials have non-negative degree, and therefore it is a polynomial with a bounded number of terms, as desired.

9. Proofs of the remaining assertions

From Theorem 2.2 we can now deduce the various statements given in Section 1 with a relatively small effort.

We first prove Theorem 1.3 and its Corollary 1.4 on integral points, i.e., regarding the regular maps for a given finite cover .

Proof of Theorem 1.3.

We first note that it suffices to prove the conclusion for a finite set of regular functions on . Therefore, we may assume that may be represented as the hypersurface , where is monic in , it is Laurent in the ’s, and is the projection onto the first coordinates.

With this proviso, we go to the proof. A regular map may be represented in the form , where , , and . Thus .

By Theorem 2.2, and using the same argument of the proof of Theorem 1.2, we see that each choice of the coefficients and of the polynomial corresponds to an integer solution of a system of linear equations (I) and to a point on an algebraic variety (II).

Now, for each system (I), let be the rank of its solution space, and let be the corresponding algebraic variety (II). By construction, we obtain a map . The above comment on and implies that there is map , given by the solution of the system (I) corresponding to , and a point corresponding to the coefficients of and the ’s, such that in fact . Since the number of possible systems, and therefore of maps , is bounded in terms of and , this yields the desired conclusion.

Proof of Corollary 1.4.

We first remark a few things about the conclusion of Theorem 1.3. First, we observe that since a regular map from to is a monomial, each is of the shape , for non-vanishing functions on and pure monomials in the . Also, since the map is a homomorphism, after an automorphism of it factors as a projection times a homomorphism with a finite kernel; hence, , and we may in fact take . (Indeed, the map sends to a fiber of , which is finite; hence this image is constant, and we may remove from the picture.)

Then, after pullback of by an isogeny, we may assume that embeds in on the first coordinates. Therefore, we can see that the map yields a family of translates of parametrized by , and corresponding regular sections of over each of them.

Turning back to the proof, we note that the hypothesis combined with Theorem 1.3 imply immediately that one of the maps is dominant. Therefore, the composition is regular and dominant, and (by the previous remarks) we may even suppose that it is expressed in the shape where are monomials in , are non-vanishing regular functions on , and .

If we fix a point , the restriction of to is an isogeny, and therefore unramified. We define as the ramification divisor of , and as the branch locus. Let, for , . Note that may be reducible, even for all . However, the image of restricted to is of the shape (where is a certain regular map ), and the map is essentially an isogeny and is finite. Then we have that consists of a finite union of components of such that .

Since is dominant, it follows easily (by counting dimensions) that can miss a whole component of only for in a proper closed subset of . On the other hand, since the said map is essentially an isogeny, cannot meet its image, so is contained in the components missed by .

Therefore can be nonempty only for , and then the projection of to is contained in . Since has pure codimension in , it follows that is a union of cosets of and is therefore invariant by multiplication by .

The proof of the toric version of Bertini’s Theorem 1.5 follows a similar pattern.

Proof of Theorem 1.5.

Let us assume first that is representable as an open dense subset of the hypersurface , where is an irreducible complex polynomial, and is the projection onto the first coordinates.

Let us analyze a factorization with integers and polynomials (Laurent in ) , monic in . By Theorem 2.2, and proceeding as in the proof of Theorem 1.2, we can see that the pairs correspond to solutions of suitable linear systems (I) and to points on the corresponding affine algebraic varieties (II).

Now, fix a system (I) and a point on the algebraic variety (II). As before, if is the rank of the solution space, we can easily obtain the following factorization:

where are (Laurent) monomials in , , and , are polynomials (Laurent in the ’s) and monic in .

Now, suppose the monomials are multiplicatively independent. This means that the homomorphism given by is surjective. By simple general theory, it must factor as a composition of a projection and an isogeny of . But then the identity (Equation 9.1) shows that the pullback is reducible; now, it is known and not too difficult to prove that this implies that is already reducible (see Reference 21, Proposition 2.1), against the assumptions.

Therefore, we may assume that in all cases the sets of monomials so obtained are multiplicatively dependent; hence they satisfy an identical relation for integer exponents , not all zero and depending only on the linear form chosen in (I). In particular, the vector takes altogether only finitely many values.

Since the ’s are pure monomials in the , we may assume that the ’s are coprime. The multiplicative relation defines a certain proper connected algebraic subgroup of , while the corresponding factorization implies that is reducible for . Therefore, the original one-dimensional torus parametrized by is contained in . We now let be the union of all finitely many sub-tori which arise in this way. Note that can be chosen dependently only on .

Now, assume that is reducible, for a certain and a certain torus of dimension . If is a parametrization of by monomials in the , then the polynomial is reducible (over ). Hence, simply by specialization, the polynomial must be reducible for all integer vectors such that the torus is contained in . But then any such torus must be contained in some as above; it is now easy to see that itself must be contained in , proving the desired conclusion.

To complete the proof, consider a general quasi-projective variety . After replacing with for a suitable proper subvariety , we may assume that is finite onto its image. We note that we may cover with finitely many (open dense) affine charts, such that for any two points of there is a chart containing both of them; since is finite over its image, we may further assume that each chart can be represented as an open dense subset of the hypersurface for some . We then observe that if has at least two irreducible components, for some subgroup and some , then there is at least one affine chart intersecting both components, and the conclusion follows by the previous case.

Finally, the only remaining statement is the analogue for composite rational functions of Schinzel’s conjecture, namely Theorem 1.7.

Proof of Theorem 1.7.

Let be given and let be as in the statement. We write with . If we put , we know by the main theorem of Reference 12 that unless we are in the exceptional situation of that theorem, where our statement is trivially true. Therefore we may write where are two (coprime) polynomials in . From we therefore get

We then define

This is a polynomial of degree at most in each variable. An application of Theorem 2.2 shows at once that there exists a number such that , which satisfies , is the ratio of two polynomials in with most terms, as desired.

Acknowledgments

The authors express gratitude to A. Fornasiero for raising the question in the non-standard setting, thus renewing interest in this problem, and to D. Ghioca and T. Scanlon for informing them about their conjecture and its link with the problems discussed here. The authors also wish to thank the anonymous referees for the very detailed reading and the various important comments, corrections, and further pointers to the literature.

Mathematical Fragments

Theorem 1.1.

Let have terms in and be monic of degree in . If satisfies , then has at most terms.

Theorem 1.2.

If is monic in and of degree at most in each variable, if are natural numbers, and if satisfies

then has at most terms.

Theorem 1.3.

Let be a finite map. Then there exists a finite set of regular maps , with an integer and an affine algebraic variety, such that for every regular map there exist a , a point , and a regular map with .

Corollary 1.4.

Suppose that the union of the images of all regular non-constant maps is Zariski dense. Then the branch locus of in is invariant by translation by an algebraic subgroup of positive dimension.

Theorem 1.5.

Let be a quasi-projective variety and be a (complex) dominant rational map of finite degree , and suppose that the pullback is irreducible. Let be a proper algebraic subvariety such that is finite onto its image. Then there exists a finite union of proper algebraic subgroups of such that if is a connected algebraic subgroup not contained in , then for all , is irreducible.

Theorem 1.7.

If are such that the composition can be written as the ratio , where have altogether at most terms, then there exist polynomials with at most terms such that .

Theorem 2.1.

If is monic in and of degree at most in each variable, if are natural numbers, and if are polynomials monic in such that

then each coefficient of (as a polynomial in ) has at most terms.

Theorem 2.2.

If is a polynomial of degree at most in each variable, if are integers, and if is such that

then is the ratio of two polynomials in with at most terms.

Theorem 2.3.

If is a polynomial of degree at most in each variable, if are integers, and if are such that

with monic in , then each coefficient of (as a polynomial in ) is the ratio of two polynomials in with at most terms.

Proposition 2.5.

If is monic in and of degree at most in each variable, if are natural numbers, and if is such that

then is the ratio of two polynomials in with at most terms.

Proposition 2.6.

If is monic in and of degree at most in each variable, if are natural numbers, and if are such that

with monic in , then each coefficient of as a polynomial in is the ratio of two polynomials in with at most terms.

Note 2.7.

It is important to note that the following deductions are valid for each single (whereas the number is changed in the course of the deductions). This is crucial in all our proofs that proceed by induction on ; namely, if we assume that one statement is true for a certain value of and all possible ’s, the other statements will follow as well for the same value of and all possible ’s.

Lemma 2.8.

Let be a polynomial such that can be written as the ratio of two polynomials with at most terms. Then is the ratio of two polynomials with at most terms.

Lemma 2.9.

Suppose that we are under the hypothesis of Theorem 2.2, and that there are integers , not all zero, and some such that

Assume moreover that Theorem 2.2 has been proved for and any degree . Then is the ratio of two polynomials with at most terms.

Definition 4.1.

Under the above notations, we say that a solution to (Equation 2.2), namely , is regular if are local parameters at .

Lemma 4.2.

Let be the divisors in on which lies. Then either

for some integers not all zero such that or the matrix is invertible (and in particular ).

Proposition 4.3.

If is monic in , is irreducible, and is of degree at most in each variable, and if and form a regular solution of

then is the ratio of two polynomials in with at most terms.

Equation (4.1)
Equation (5.1)
Equation (5.2)
Equation (5.3)
Proposition 5.1.

The coefficients generate a finite extension of degree at most of . Similarly, the coefficients generate a finite extension of degree at most of .

Lemma 5.2.

Let . Let be an irreducible polynomial such that

If , then the degree of in each variable is at most for a suitable .

Proposition 5.3.

For all , there is a primitive element of such that, if is an irreducible polynomial such that

then the degree of in each variable is at most . Moreover, we may assume that is monic in .

Proposition 5.4.

For all , if , then either or there are integers , not all zero, such that and .

Corollary 5.5.

For all , either the genus of is bounded by for some or there are integers , not all zero, such that and .

Lemma 6.1 (Reference 20, Proposition 1).

Let be a function field in one variable, of genus , let be linearly independent over , and let . Let be a finite set of places of containing all the poles of and also all the zeros of . Further, put . Then

where .

Proposition 6.2.

Suppose that and that for some given . Then at least one of the following holds:

(1)

is -linearly dependent on the set ,

(2)

for some ,

(3)

there are , not all zero, such that and .

Proposition 7.2.

Suppose that is -linearly dependent on . Assume that Proposition 4.3 is true for . Then either for some integers , not all zero, such that , or can be written as the ratio of two polynomials with at most terms.

Equation (8.1)
Equation (9.1)

References

Reference [1]
F. Beukers, Ternary form equations, J. Number Theory 54 (1995), no. 1, 113–133, DOI 10.1006/jnth.1995.1105. MR1352640,
Show rawAMSref \bib{Beukers1995}{article}{ author={Beukers, F.}, title={Ternary form equations}, journal={J.~Number Theory}, volume={54}, date={1995}, number={1}, pages={113\textendash 133}, issn={0022-314X}, review={\MR {1352640}}, doi={10.1006/jnth.1995.1105}, }
Reference [2]
Enrico Bombieri and Walter Gubler, Heights in Diophantine geometry, New Mathematical Monographs, vol. 4, Cambridge University Press, Cambridge, 2006, DOI 10.1017/CBO9780511542879. MR2216774,
Show rawAMSref \bib{Bombieri2006}{book}{ author={Bombieri, Enrico}, author={Gubler, Walter}, title={Heights in Diophantine geometry}, series={New Mathematical Monographs}, volume={4}, publisher={Cambridge University Press, Cambridge}, date={2006}, pages={xvi+652}, isbn={978-0-521-84615-8}, isbn={0-521-84615-3}, review={\MR {2216774}}, doi={10.1017/CBO9780511542879}, }
Reference [3]
E. Bombieri, D. Masser, and U. Zannier, Anomalous subvarieties—structure theorems and applications, Int. Math. Res. Not. IMRN 19 (2007), Art. ID rnm057, 33, DOI 10.1093/imrn/rnm057. MR2359537,
Show rawAMSref \bib{Bombieri2007}{article}{ author={Bombieri, E.}, author={Masser, D.}, author={Zannier, U.}, title={Anomalous subvarieties\textemdash structure theorems and applications}, journal={Int. Math. Res. Not. IMRN}, date={2007}, number={19}, pages={Art. ID rnm057, 33}, issn={1073-7928}, review={\MR {2359537}}, doi={10.1093/imrn/rnm057}, }
Reference [4]
Pietro Corvaja and Umberto Zannier, On the integral points on certain surfaces, Int. Math. Res. Not. IMRN 20 (2006), Art. ID 98623, DOI 10.1155/IMRN/2006/98623. MR2219222,
Show rawAMSref \bib{Corvaja2006}{article}{ author={Corvaja, Pietro}, author={Zannier, Umberto}, title={On the integral points on certain surfaces}, journal={Int. Math. Res. Not. IMRN}, date={2006}, number={20}, pages={Art. ID 98623}, issn={1073-7928}, review={\MR {2219222}}, doi={10.1155/IMRN/2006/98623}, }
Reference [5]
Pietro Corvaja and Umberto Zannier, Some cases of Vojta’s conjecture on integral points over function fields, J. Algebraic Geom. 17 (2008), no. 2, 295–333, DOI 10.1090/S1056-3911-07-00489-4. MR2369088,
Show rawAMSref \bib{Corvaja2008}{article}{ author={Corvaja, Pietro}, author={Zannier, Umberto}, title={Some cases of Vojta's conjecture on integral points over function fields}, journal={J. Algebraic Geom.}, volume={17}, date={2008}, number={2}, pages={295\textendash 333}, issn={1056-3911}, review={\MR {2369088}}, doi={10.1090/S1056-3911-07-00489-4}, }
Reference [6]
Pietro Corvaja and Umberto Zannier, Algebraic hyperbolicity of ramified covers of (and integral points on affine subsets of ), J. Differential Geom. 93 (2013), no. 3, 355–377. MR3024299,
Show rawAMSref \bib{Corvaja2013}{article}{ author={Corvaja, Pietro}, author={Zannier, Umberto}, title={Algebraic hyperbolicity of ramified covers of $\mathbb {G}^2_m$ (and integral points on affine subsets of $\mathbb {P}_2$)}, journal={J. Differential Geom.}, volume={93}, date={2013}, number={3}, pages={355\textendash 377}, issn={0022-040X}, review={\MR {3024299}}, }
Reference [7]
Pietro Corvaja and Umberto Zannier, Finiteness of odd perfect powers with four nonzero binary digits (English, with English and French summaries), Ann. Inst. Fourier (Grenoble) 63 (2013), no. 2, 715–731. MR3112846,
Show rawAMSref \bib{Corvaja2013a}{article}{ author={Corvaja, Pietro}, author={Zannier, Umberto}, title={Finiteness of odd perfect powers with four nonzero binary digits}, language={English, with English and French summaries}, journal={Ann. Inst. Fourier (Grenoble)}, volume={63}, date={2013}, number={2}, pages={715\textendash 731}, issn={0373-0956}, review={\MR {3112846}}, }
Reference [8]
James Harold Davenport and Jacques Carette, The Sparsity Challenges. In 2009 11th International Symposium on Symbolic and Numeric Algorithms for Scientific Computing, 3–7, IEEE, 2009, DOI 10.1109/SYNASC.2009.62.
Reference [9]
Paul Erdős, On the number of terms of the square of a polynomial, Nieuw Arch. Wiskunde (2) 23 (1949), 63–65. MR0027779,
Show rawAMSref \bib{Erdos1949}{article}{ author={Erd\H {o}s, Paul}, title={On the number of terms of the square of a polynomial}, journal={Nieuw Arch. Wiskunde (2)}, volume={23}, date={1949}, pages={63\textendash 65}, review={\MR {0027779}}, }
Reference [10]
Michael D. Fried and Moshe Jarden, Field arithmetic, 3rd ed., Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics], vol. 11, Springer-Verlag, Berlin, 2008. Revised by Jarden. MR2445111,
Show rawAMSref \bib{Fried2008}{book}{ author={Fried, Michael D.}, author={Jarden, Moshe}, title={Field arithmetic}, series={Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]}, volume={11}, edition={3}, note={Revised by Jarden}, publisher={Springer-Verlag, Berlin}, date={2008}, pages={xxiv+792}, isbn={978-3-540-77269-9}, review={\MR {2445111}}, }
Reference [11]
Clemens Fuchs, Vincenzo Mantova, and Umberto Zannier, On fewnomials, integral points and a toric version of Bertini’s theorem, 2014, arXiv 1412.4548v1.
Reference [12]
Clemens Fuchs and Umberto Zannier, Composite rational functions expressible with few terms, J. Eur. Math. Soc. (JEMS) 14 (2012), no. 1, 175–208, DOI 10.4171/JEMS/299. MR2862037,
Show rawAMSref \bib{Fuchs2012}{article}{ author={Fuchs, Clemens}, author={Zannier, Umberto}, title={Composite rational functions expressible with few terms}, journal={J. Eur. Math. Soc. (JEMS)}, volume={14}, date={2012}, number={1}, pages={175\textendash 208}, issn={1435-9855}, review={\MR {2862037}}, doi={10.4171/JEMS/299}, }
Reference [13]
A. G. Khovanskiĭ, Fewnomials, Translations of Mathematical Monographs, vol. 88, American Mathematical Society, Providence, RI, 1991. Translated from the Russian by Smilka Zdravkovska. MR1108621,
Show rawAMSref \bib{Khovanski1991}{book}{ author={Khovanski\u \i , A. G.}, title={Fewnomials}, series={Translations of Mathematical Monographs}, volume={88}, note={Translated from the Russian by Smilka Zdravkovska}, publisher={American Mathematical Society, Providence, RI}, date={1991}, pages={viii+139}, isbn={0-8218-4547-0}, review={\MR {1108621}}, }
Reference [14]
Steven S. Y. Lu, On surfaces of general type with maximal Albanese dimension, J. Reine Angew. Math. 641 (2010), 163–175, DOI 10.1515/CRELLE.2010.032. MR2643929,
Show rawAMSref \bib{Lu2010}{article}{ author={Lu, Steven S. Y.}, title={On surfaces of general type with maximal Albanese dimension}, journal={J. Reine Angew. Math.}, volume={641}, date={2010}, pages={163\textendash 175}, issn={0075-4102}, review={\MR {2643929}}, doi={10.1515/CRELLE.2010.032}, }
Reference [15]
R. C. Mason, Diophantine equations over function fields, London Mathematical Society Lecture Note Series, vol. 96, Cambridge University Press, Cambridge, 1984, DOI 10.1017/CBO9780511752490. MR754559,
Show rawAMSref \bib{Mason1984}{book}{ author={Mason, R. C.}, title={Diophantine equations over function fields}, series={London Mathematical Society Lecture Note Series}, volume={96}, publisher={Cambridge University Press, Cambridge}, date={1984}, pages={x+125}, isbn={0-521-26983-0}, review={\MR {754559}}, doi={10.1017/CBO9780511752490}, }
Reference [16]
Junjiro Noguchi and Jörg Winkelmann, Nevanlinna theory in several complex variables and Diophantine approximation, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 350, Springer, Tokyo, 2014, DOI 10.1007/978-4-431-54571-2. MR3156076,
Show rawAMSref \bib{Noguchi2014}{book}{ author={Noguchi, Junjiro}, author={Winkelmann, J\"org}, title={Nevanlinna theory in several complex variables and Diophantine approximation}, series={Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]}, volume={350}, publisher={Springer, Tokyo}, date={2014}, pages={xiv+416}, isbn={978-4-431-54570-5}, isbn={978-4-431-54571-2}, review={\MR {3156076}}, doi={10.1007/978-4-431-54571-2}, }
Reference [17]
A. Schinzel, On the number of terms of a power of a polynomial, Acta Arith. 49 (1987), no. 1, 55–70. MR913764,
Show rawAMSref \bib{Schinzel1987}{article}{ author={Schinzel, A.}, title={On the number of terms of a power of a polynomial}, journal={Acta Arith.}, volume={49}, date={1987}, number={1}, pages={55\textendash 70}, issn={0065-1036}, review={\MR {913764}}, }
Reference [18]
A. Schinzel, Polynomials with special regard to reducibility, Encyclopedia of Mathematics and Its Applications, vol. 77, Cambridge University Press, Cambridge, 2000. With an appendix by Umberto Zannier, DOI 10.1017/CBO9780511542916. MR1770638,
Show rawAMSref \bib{Schinzel2000}{book}{ author={Schinzel, A.}, title={Polynomials with special regard to reducibility}, series={Encyclopedia of Mathematics and Its Applications}, volume={77}, note={With an appendix by Umberto Zannier}, publisher={Cambridge University Press, Cambridge}, date={2000}, pages={x+558}, isbn={0-521-66225-7}, review={\MR {1770638}}, doi={10.1017/CBO9780511542916}, }
Reference [19]
Andrzej Schinzel and Umberto Zannier, On the number of terms of a power of a polynomial, Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl. 20 (2009), no. 1, 95–98, DOI 10.4171/RLM/534. MR2491571,
Show rawAMSref \bib{Schinzel2009}{article}{ author={Schinzel, Andrzej}, author={Zannier, Umberto}, title={On the number of terms of a power of a polynomial}, journal={Atti Accad. Naz. Lincei Rend. Lincei Mat. Appl.}, volume={20}, date={2009}, number={1}, pages={95\textendash 98}, issn={1120-6330}, review={\MR {2491571}}, doi={10.4171/RLM/534}, }
Reference [20]
Umberto Zannier, On composite lacunary polynomials and the proof of a conjecture of Schinzel, Invent. Math. 174 (2008), no. 1, 127–138, DOI 10.1007/s00222-008-0136-8. MR2430978,
Show rawAMSref \bib{Zannier2008}{article}{ author={Zannier, Umberto}, title={On composite lacunary polynomials and the proof of a conjecture of Schinzel}, journal={Invent. Math.}, volume={174}, date={2008}, number={1}, pages={127\textendash 138}, issn={0020-9910}, review={\MR {2430978}}, doi={10.1007/s00222-008-0136-8}, }
Reference [21]
Umberto Zannier, Hilbert irreducibility above algebraic groups, Duke Math. J. 153 (2010), no. 2, 397–425, DOI 10.1215/00127094-2010-027. MR2667137,
Show rawAMSref \bib{Zannier2010}{article}{ author={Zannier, Umberto}, title={Hilbert irreducibility above algebraic groups}, journal={Duke Math. J.}, volume={153}, date={2010}, number={2}, pages={397\textendash 425}, issn={0012-7094}, review={\MR {2667137}}, doi={10.1215/00127094-2010-027}, }
Reference [22]
Umberto Zannier, Some problems of unlikely intersections in arithmetic and geometry, Annals of Mathematics Studies, vol. 181, Princeton University Press, Princeton, NJ, 2012. With appendixes by David Masser. MR2918151,
Show rawAMSref \bib{Zannier2012}{book}{ author={Zannier, Umberto}, title={Some problems of unlikely intersections in arithmetic and geometry}, series={Annals of Mathematics Studies}, volume={181}, note={With appendixes by David Masser}, publisher={Princeton University Press, Princeton, NJ}, date={2012}, pages={xiv+160}, isbn={978-0-691-15371-1}, review={\MR {2918151}}, }

Article Information

MSC 2010
Primary: 11C08 (Polynomials)
Secondary: 12E05 (Polynomials), 12Y05 (Computational aspects of field theory and polynomials), 14G05 (Rational points), 14J99 (None of the above, but in this section), 11U10 (Nonstandard arithmetic)
Author Information
Clemens Fuchs
Department of Mathematics, University of Salzburg, Hellbrunnerstrasse 34/I, A-5020 Salzburg, Austria
clemens.fuchs@sbg.ac.at
ORCID
MathSciNet
Vincenzo Mantova
School of Science and Technology, Mathematics Division, University of Camerino, Via Madonna delle Carceri 9, IT-62032 Camerino, Italy
Address at time of publication: School of Mathematics, University of Leeds, LS2 9JT Leeds, United Kingdom
v.l.mantova@leeds.ac.uk
ORCID
MathSciNet
Umberto Zannier
Classe di Scienze, Scuola Normale Superiore, Piazza dei Cavalieri 7, IT-56126 Pisa, Italy
umberto.zannier@sns.it
MathSciNet
Additional Notes

The first author was supported by FWF (Austrian Science Fund) grant No. P24574.

The second author was supported by the Italian FIRB 2010 RBFR10V792 “New advances in the Model Theory of exponentiation.”.

The authors were also supported by the ERC-AdG 267273 “Diophantine Problems.”.

Journal Information
Journal of the American Mathematical Society, Volume 31, Issue 1, ISSN 1088-6834, published by the American Mathematical Society, Providence, Rhode Island.
Publication History
This article was received on , revised on , , and published on .
Copyright Information
Copyright 2017 American Mathematical Society
Article References
  • Permalink
  • Permalink (PDF)
  • DOI 10.1090/jams/878
  • MathSciNet Review: 3718452
  • Show rawAMSref \bib{3718452}{article}{ author={Fuchs, Clemens}, author={Mantova, Vincenzo}, author={Zannier, Umberto}, title={On fewnomials, integral points, and a toric version of Bertini's theorem}, journal={J. Amer. Math. Soc.}, volume={31}, number={1}, date={2018-01}, pages={107-134}, issn={0894-0347}, review={3718452}, doi={10.1090/jams/878}, }

Settings

Change font size
Resize article panel
Enable equation enrichment

Note. To explore an equation, focus it (e.g., by clicking on it) and use the arrow keys to navigate its structure. Screenreader users should be advised that enabling speech synthesis will lead to duplicate aural rendering.

For more information please visit the AMS MathViewer documentation.