To Robert Tijdeman on his 75-th birthday

Similar documents
Results and open problems related to Schmidt s Subspace Theorem. Jan-Hendrik Evertse

LINEAR EQUATIONS WITH UNKNOWNS FROM A MULTIPLICATIVE GROUP IN A FUNCTION FIELD. To Professor Wolfgang Schmidt on his 75th birthday

Some remarks on diophantine equations and diophantine approximation

Cover Page. The handle holds various files of this Leiden University dissertation

On Schmidt s Subspace Theorem. Jan-Hendrik Evertse

On the norm form inequality F (x) h.

NOTES ON DIOPHANTINE APPROXIMATION

Effective results for unit equations over finitely generated domains. Jan-Hendrik Evertse

Chapter 8. P-adic numbers. 8.1 Absolute values

Approximation of complex algebraic numbers by algebraic numbers of bounded degree

LINEAR FORMS IN LOGARITHMS

A POSITIVE PROPORTION OF THUE EQUATIONS FAIL THE INTEGRAL HASSE PRINCIPLE

The number of solutions of linear equations in roots of unity

NOTES ON DIOPHANTINE APPROXIMATION

RECENT RESULTS ON LINEAR RECURRENCE SEQUENCES

A QUANTITATIVE VERSION OF THE ABSOLUTE SUBSPACE THEOREM

Chapter 6. Approximation of algebraic numbers by rationals. 6.1 Liouville s Theorem and Roth s Theorem

TC10 / 3. Finite fields S. Xambó

Course 311: Michaelmas Term 2005 Part III: Topics in Commutative Algebra

Unit equations in characteristic p. Peter Koymans

8. Prime Factorization and Primary Decompositions

= 1 2x. x 2 a ) 0 (mod p n ), (x 2 + 2a + a2. x a ) 2

Local Fields. Chapter Absolute Values and Discrete Valuations Definitions and Comments

On prime factors of subset sums

Chapter 1 : The language of mathematics.

ON A PROBLEM OF PILLAI AND ITS GENERALIZATIONS

Cullen Numbers in Binary Recurrent Sequences

Chapter 6. Approximation of algebraic numbers by rationals. 6.1 Liouville s Theorem and Roth s Theorem

MINKOWSKI THEORY AND THE CLASS NUMBER

MULTIPLICITIES OF MONOMIAL IDEALS

THESIS. Presented in Partial Fulfillment of the Requirements for the Degree Master of Science in the Graduate School of The Ohio State University

Mathematics Course 111: Algebra I Part I: Algebraic Structures, Sets and Permutations

Irreducible Polynomials over Finite Fields

Fibonacci numbers of the form p a ± p b

Lower bounds for resultants II.

To Professor W. M. Schmidt on his 60th birthday

be a sequence of positive integers with a n+1 a n lim inf n > 2. [α a n] α a n

Factorization of integer-valued polynomials with square-free denominator

FORMAL GROUPS OF CERTAIN Q-CURVES OVER QUADRATIC FIELDS

1 Adeles over Q. 1.1 Absolute values

PRACTICE PROBLEMS: SET 1

P -adic root separation for quadratic and cubic polynomials

ON THE APPROXIMATION TO ALGEBRAIC NUMBERS BY ALGEBRAIC NUMBERS. Yann Bugeaud Université de Strasbourg, France

THE NUMBER OF SOLUTIONS OF DECOMPOSABLE FORM EQUATIONS. Jan-Hendrik Evertse

A further improvement of the Quantitative Subspace Theorem

Chapter 1 Preliminaries

Introduction to Real Analysis Alternative Chapter 1

Predictive criteria for the representation of primes by binary quadratic forms

Finite Fields and Error-Correcting Codes

#A5 INTEGERS 18A (2018) EXPLICIT EXAMPLES OF p-adic NUMBERS WITH PRESCRIBED IRRATIONALITY EXPONENT

0 Sets and Induction. Sets

On the power-free parts of consecutive integers

Definitions. Notations. Injective, Surjective and Bijective. Divides. Cartesian Product. Relations. Equivalence Relations

Factorization in Integral Domains II

SMALL ZEROS OF QUADRATIC FORMS WITH LINEAR CONDITIONS. Lenny Fukshansky. f ij X i Y j

Polynomials, Ideals, and Gröbner Bases

Hamburger Beiträge zur Mathematik

The Subspace Theorem and twisted heights

Determining elements of minimal index in an infinite family of totally real bicyclic biquadratic number fields

Solutions of exercise sheet 6

NUMBERS WITH INTEGER COMPLEXITY CLOSE TO THE LOWER BOUND

Higher-Dimensional Analogues of the Combinatorial Nullstellensatz

Part II. Number Theory. Year

Discrete Math, Second Problem Set (June 24)

1. Let r, s, t, v be the homogeneous relations defined on the set M = {2, 3, 4, 5, 6} by

Local properties of plane algebraic curves

Codewords of small weight in the (dual) code of points and k-spaces of P G(n, q)

School of Mathematics and Statistics. MT5836 Galois Theory. Handout 0: Course Information

GALOIS THEORY. Contents

Places of Number Fields and Function Fields MATH 681, Spring 2018

How many units can a commutative ring have?

A field F is a set of numbers that includes the two numbers 0 and 1 and satisfies the properties:

Pacific Journal of Mathematics

Math 429/581 (Advanced) Group Theory. Summary of Definitions, Examples, and Theorems by Stefan Gille

D-MATH Algebra I HS18 Prof. Rahul Pandharipande. Solution 6. Unique Factorization Domains

PolynomialExponential Equations in Two Variables

Polynomial Rings. i=0. i=0. n+m. i=0. k=0

On a Theorem of Dedekind

NOTES ON FINITE FIELDS

Quadratic reciprocity (after Weil) 1. Standard set-up and Poisson summation

HASSE-MINKOWSKI THEOREM

MATH 8253 ALGEBRAIC GEOMETRY WEEK 12

w d : Y 0 (N) Y 0 (N)

Algebraic function fields

ZEROS OF SPARSE POLYNOMIALS OVER LOCAL FIELDS OF CHARACTERISTIC p

Fields and Galois Theory. Below are some results dealing with fields, up to and including the fundamental theorem of Galois theory.

Partial cubes: structures, characterizations, and constructions

Jónsson posets and unary Jónsson algebras

1 Basic Combinatorics

6 Lecture 6: More constructions with Huber rings

On prime factors of integers which are sums or shifted products. by C.L. Stewart (Waterloo)

ON A THEOREM OF TARTAKOWSKY

CHAPTER 0 PRELIMINARY MATERIAL. Paul Vojta. University of California, Berkeley. 18 February 1998

Axioms for Set Theory

A MATROID EXTENSION RESULT

Course 2316 Sample Paper 1

Public-key Cryptography: Theory and Practice

Jeong-Hyun Kang Department of Mathematics, University of West Georgia, Carrollton, GA

ALGEBRA PH.D. QUALIFYING EXAM September 27, 2008

Topological properties

Transcription:

S-PARTS OF VALUES OF UNIVARIATE POLYNOMIALS, BINARY FORMS AND DECOMPOSABLE FORMS AT INTEGRAL POINTS YANN BUGEAUD, JAN-HENDRIK EVERTSE, AND KÁLMÁN GYŐRY To Robert Tijdeman on his 75-th birthday 1. Introduction Let S = {p 1,..., p s } be a finite, non-empty set of distinct prime numbers. For a non-zero integer m, write m = p a 1 1... p as s b, where a 1,..., a s are non-negative integers and b is an integer relatively prime to p 1 p s. Then we define the S-part [m] S of m by [m] S := p a 1 1... p as s. The motivation of the present paper was given by the following result, established in 2013 by Gross and Vincent [10]. Theorem A. Let f(x) be a polynomial with integral coefficients with at least two distinct roots and S a finite, non-empty set of prime numbers. Then there exist effectively computable positive numbers κ 1 and κ 2, depending only on f(x) and S, such that for every non-zero integer x that is not a root of f(x) we have [f(x)] S < κ 2 f(x) 1 κ 1. We mention that earlier, Stewart [24] obtained a similar result, but only for polynomials whose zeros are consecutive integers. Gross and Vincent s proof of Theorem A depends on the theory of linear forms in complex logarithms, Under the additional hypotheses that f(x) has degree n 2 and no multiple roots, we deduce an ineffective analogue of Theorem A, with instead of 1 κ 1 an exponent 1 n +ɛ for every ɛ > 0 and instead of κ 2 an ineffective number depending 2010 Mathematics Subject Classification: 11D45,11D57,11D59,11J86,11J87. Keywords and Phrases: S-part, polynomials, binary forms, decomposable forms, Subspace Theorem, Baker theory. February 20, 2018. 1

2 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY on f(x), S and ɛ. This is in fact an easy application of the p-adic Thue- Siegel-Roth Theorem. We show that the exponent 1 n is best possible. Lastly, we give an estimate for the density of the set of integers x for which [f(x)] S is large, i.e., for every small ɛ > 0 we estimate in terms of B the number of integers x with x B such that [f(x)] S f(x) ɛ. We considerably extend both Theorem A, its ineffective analogue, and the density result by proving similar results for the S-parts of values of homogeneous binary forms and, more generally, of values of decomposable forms at integer points, under suitable assumptions. In addition, in the effective results we give an expression for κ 1, which is explicit in terms of S. For our extensions to binary forms and decomposable forms, we use the p-adic Thue-Siegel-Roth Theorem and the p-adic Subspace Theorem of Schmidt and Schlickewei for the ineffective estimates for the S-part. The proof of the effective estimates is based on an effective theorem of Győry and Yu [15] on decomposable form equations whose proof depends on estimates for linear forms in complex and in p-adic logarithms. Lastly, the proofs of our density results on the number of integer points of norm at most B at which the value of the binary form or decomposable form under consideration has large S-value are based on a recent general lattice point counting result of Barroero and Widmer [1] and on work in the PhD-thesis of Junjiang Liu [16]. For simplicity, we have restricted ourselves to univariate polynomials, binary forms and decomposable forms with coefficients in Z. With some extra technical effort, analogous results could have been obtained over arbitrary number fields. In Section 2 we state our results, in Sections 3 6 we give the proofs, in Sections 7 and 8 we present some applications, and in Section 9 we give some additional comments on Theorem A. 2. Results 2.1. Results for univariate polynomials and binary forms. We use notation a,b,..., a,b,... to indicate that the constants implied by the Vinogradov symbols depend only on the parameters a, b,.... Further, we use the notation A a,b,... B to indicate that both A a,b,... B and B a,b,... A hold. We prove the following ineffective analogue of Theorem A mentioned in the previous section.

S-PARTS OF VALUES OF POLYNOMIALS 3 Theorem 2.1. Let f(x) Z[X] be a polynomial of degree n 2 without multiple zeros. (i) Let S = {p 1,..., p s } be a non-empty set of primes. Then for every ɛ > 0 and for every x Z with f(x) 0, [f(x)] S f,s,ɛ f(x) (1/n)+ɛ. (ii) There are infinitely many primes p, and for each of these p, there are infinitely many integers x, such that f(x) 0 and [f(x)] {p} f f(x) 1/n. For completeness, we give here also a more precise effective version of Theorem A, which is a consequence of Theorem 2.5 stated below on the S-parts of values of binary forms. Theorem 2.2. Let f(x) Z[X] be a polynomial with at least two distinct roots and suppose that its splitting field has degree d over Q. Further, let S = {p 1,..., p s } be a non-empty set of primes and put P := max(p 1,..., p s ). Then for every integer x with f(x) 0 we have where κ 1 = [f(x)] S κ 2 f(x) 1 κ 1, ( c s 1( P (log p1 ) (log p s ) ) d ) 1, and c 1, κ 2 are effectively computable positive numbers that depend only on f(x). For variations on this result, and related results, we refer to Section 9. For polynomials X(X + 1) and X 2 + 7 and special sets S, Bennett, Filaseta, and Trifonov [2, 3] have obtained stronger effective results. As is to be expected, for most integers x, the S-part [f(x)] S is small compared with f(x). This is made more precise in the following result. For any finite set of primes S and any ɛ > 0, B > 0, we denote by N(f, S, ɛ, B) the number of integers x such that (2.1) x B, f(x) 0, [f(x)] S f(x) ɛ. Denote by D(f) the discriminant of f and for a prime p, denote by g p the largest integer g such that p g divides D(f).

4 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY Theorem 2.3. Let f(x) Z[X] be a polynomial of degree n 2 with non-zero discriminant. Further, let 0 < ɛ < 1/n, and let S be a finite set of primes. Denote by s the number of primes p S such that f(x) 0 (mod p gp+1 ) is solvable and assume that this number is positive. Then N(f, S, ɛ, B) f,s,ɛ B 1 nɛ (log B) s 1 as B. Remarks. 1. If s = 0 then [f(x)] S is bounded, and so the set of integers x with [f(x)] S f(x) ɛ is finite. 2. There are infinitely many primes p such that f(x) 0 (mod p) is solvable. Removing from those the finitely many that divide D(f), there remain infinitely many primes p such that g p = 0 and f(x) 0 (mod p) is solvable. We now formulate some analogues of the above mentioned results for binary forms. Denote by Z 2 prim the set of pairs (x, y) Z 2 with gcd(x, y) = 1. Theorem 2.4. Let F (X, Y ) Z[X, Y ] be a binary form of degree n 2 with non-zero discriminant. (i) Let S = {p 1,..., p s } be a non-empty set of primes. Then for every ɛ > 0 and every pair (x, y) Z 2 prim with F (x, y) 0, [F (x, y)] S F,S,ɛ F (x, y) (2/n)+ɛ. (ii) There are finite sets of primes S with the smallest prime in S arbitrarily large, and for every one of these sets S infinitely many pairs (x, y) Z 2 prim, such that F (x, y) 0 and [F (x, y)] S F,S,ɛ F (x, y) 2/n. Our next result is an effective analogue of Theorem 2.2 for binary forms. It is an easy consequence of Theorem 2.10 stated below on decomposable forms. The splitting field of a binary form is the smallest extension of Q over which it factors into linear forms. Theorem 2.5. Let F (X, Y ) be a binary form of degree n 3 with coefficients in Z and with splitting field K. Suppose that F has at least

S-PARTS OF VALUES OF POLYNOMIALS 5 three pairwise non-proportional linear factors over K. Let again S = {p 1,..., p s } be a finite set of primes and [K : Q] = d. Then [F (x, y)] S κ 4 F (x, y) 1 κ 3 for every (x, y) Z 2 prim with F (x, y) 0, where κ 3 = ( ( c s 2 (P (log p1 ) (log p s ) ) d) 1 and κ 4, c 2 are effectively computable positive numbers, depending only on F. We obtain Theorem 2.2 on polynomials f(x) Z[X] by applying Theorem 2.5 to the binary form Y 1+deg f f(x/y ) with (x, y) = (x, 1) Z 2 prim. Let again F (X, Y ) Z[X, Y ] be a binary form of degree n 2 and of non-zero discriminant. For any finite set of primes S and any ɛ > 0, B > 0, we denote by N(F, S, ɛ, B) the number of pairs (x, y) Z 2 prim such that (2.2) max( x, y ) B, F (x, y) 0, [F (x, y)] S F (x, y) ɛ. Denote by D(F ) the discriminant of F and for a prime p, denote by g p the largest integer g such that p g divides D(F ). Theorem 2.6. Let F (X, Y ) Z[X, Y ] be a binary form of degree n 3 with non-zero discriminant. Further, let 0 < ɛ < 1 n, and let S be a finite set of primes. Denote by s the number of primes p S such that F (x, y) 0 (mod p gp+1 ) has a solution (x, y) Z 2 prim and assume that this number is positive. Then N(F, S, ɛ, B) F,S,ɛ B 2 nɛ (log B) s 1 as B. Parts (i) of Theorems 2.1 and 2.4 are easy consequences of the p-adic Thue-Siegel-Roth Theorem. Part (ii) of Theorem 2.1 is a consequence of the fact that for a given non-constant polynomial f(x) Z[X] there are infinitely many primes p such that f(x) has a zero in Z p. The proof of part (ii) of Theorem 2.4 uses some geometry of numbers. There are two main tools in the proof of Theorem 2.6. The first is a result of Stewart [25, Thm. 2] on the number of congruence classes x modulo p k of f(x) 0 (mod p k ) for f(x) a polynomial and p k a prime power. The second is a powerful lattice point counting result of Barroero and Widmer [1, Thm. 1.3]. The proof of Theorem 2.3 is very

6 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY similar, but instead of the result of Barroero and Widmer it uses a much more elementary counting argument. 2.2. Ineffective results for decomposable forms. We will state results on the S-parts of values of decomposable forms in m variables at integral points, where m 2. We start with some notation and definitions. Let K be a finite, normal extension of Q. For a linear form l = α 1 X 1 + + α m X m with coefficients in K and for an element σ of the Galois group Gal(K/Q) we define σ(l) := σ(α 1 )X 1 + + σ(α m )X m and then for a set of linear forms L = {l 1,..., l r } with coefficients in K we write σ(l) := {σ(l 1 ),..., σ(l r )}. A set of linear forms L with coefficients in K is called Gal(K/Q)-symmetric if σ(l) = L for each σ Gal(K/Q), and Gal(K/Q)-proper if for each σ L we have either σ(l) = L or σ(l) L =. We denote by [L] the K-vector space generated by L, and define rank L to be the dimension of [L] over K. Finally, we define the sum of two vector spaces V 1, V 2 over K by V 1 + V 2 := {x + y : x V 1, y V 2 }. Recall that a decomposable form in Z[X 1,..., X m ] is a homogeneous polynomial that factors into linear forms in X 1,..., X m over some extension of Q. The smallest extension over which such a factorization is possible is called the splitting field of the decomposable form. This is a finite, normal extension of Q. Let F Z[X 1,..., X m ] be a decomposable form of degree n 3 with splitting field K. Then we can express F as (2.3) F = cl e(l 1) 1 l e(lr) r c a non-zero rational, with L F = {l 1,..., l r } a Gal(K/Q)-symmetric set of pairwise non-proportional linear forms with coefficients in K, e(l 1 ),..., e(l r ) positive integers, with e(l i ) = e(l j ) whenever l j = σ(l i ) for some σ Gal(K/Q). Lastly, define Z m prim to be the set of x = (x 1,..., x m ) Z m with gcd(x 1,..., x m ) = 1 and define x to be the maximum norm of x Z m prim.

S-PARTS OF VALUES OF POLYNOMIALS 7 Let S = {p 1,..., p s } be a finite set of primes, and F Z[X 1,..., X m ] a decomposable form. For x Z m prim with F (x) 0, we can write (2.4) F (x) = p a 1 1 p as s b, where a 1,..., a s are non-negative integers and b is an integer coprime with p 1 p s. Then the S-part [F (x)] S is p a 1 1 p as s. We may view (2.4) as a Diophantine equation in x Z m prim and a 1,..., a s Z 0, a socalled decomposable form equation. Schlickewei [23] considered (2.4) in the case that F is a norm form (i.e., a decomposable form that is irreducible over Q) and formulated a criterion in terms of F implying that (2.4) has only finitely many solutions. Evertse and Győry [7] gave another finiteness criterion in terms of F, valid for arbitrary decomposable forms. Recently [8, Chap. 9, Thm. 9.1.1], they refined this as follows. Call an integer S-free if it is non-zero, and coprime with the primes in S. Theorem B. Let F Z[X 1,..., X m ] be a decomposable form with splitting field K, given in the form (2.3), and let L be a finite set of linear forms in K[X 1,..., X m ], containing L F. Then the following two assertions are equivalent: (i) rank L F = m, and for every Gal(K/Q)-proper subset M of L F with M L F, we have ( ) (2.5) L [σ(m)] [L F \ σ(m)] ; σ Gal(K/Q) (ii) for every finite set of primes S = {p 1,..., p s } and every S- free integer b, there are only finitely many x Z m prim and nonnegative integers a 1,..., a s such that (2.6) F (x) = p a 1 1 p as s b, l(x) 0 for l L. This theorem was deduced from a finiteness theorem of Evertse [5] and van der Poorten and Schlickewei [20, 21] on S-unit equations over number fields. The following result gives an improvement of (ii). We denote by the standard archimedean absolute value on Q, and for a prime p by p the standard p-adic absolute value, with p p = p 1. Further, x denotes the maximum norm of x Z m prim.

8 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY Theorem 2.7. Let F Z[X 1,..., X m ] be a decomposable form in m 2 variables with splitting field K and L L F a finite set of linear forms in K[X 1,..., X m ], satisfying condition (i) of Theorem B. Further, let S be a finite set of primes and let ɛ > 0. Then there are only finitely many x Z m prim with F (x) p x (1/(m 1)) ɛ, (2.7) p S { } l(x) 0 for l L. Chen and Ru [4] proved a similar result with L = L F the set of linear factors of F and with a stronger condition instead of (i), on the other hand they considered decomposable forms with coefficients in an arbitrary number field. From Theorem 2.7 and Theorem B we deduce the following corollary. Corollary 2.8. Let F Z[X 1,..., X m ] be a decomposable form in m 2 variables with splitting field K and L L F a finite set of linear forms in K[X 1,..., X m ]. (i) Assume that F and L satisfy condition (i) of Theorem B. Suppose F has degree n. Let S be a finite set of primes and let ɛ > 0. Then for every x Z m prim with l(x) 0 for l L we have (2.8) [F (x)] S F,L,S,ɛ F (x) 1 (1/n(m 1))+ɛ. (ii) Assume that F and L do not satisfy condition (i) of Theorem B. Then there are a finite set of primes S and a constant γ > 0 such that [F (x)] S γ F (x) holds for infinitely many x Z m prim with l(x) 0 for all l L. Indeed, if F, L satisfy condition (i) of Theorem B, S is a finite set of primes and ɛ > 0, then F (x) = F (x) p x (1/(m 1)) ɛ F (x) (1/n(m 1)) ɛ/n [F (x)] S p S { } holds for all x Z m prim with l(x) 0 for all l L, where the implied constants depend on F, S and ɛ. This implies part (i) of Corollary 2.8. If on the other hand F and L do not satisfy condition (i) of Theorem

S-PARTS OF VALUES OF POLYNOMIALS 9 B then there are a finite set of primes S and an S-free integer b such that (2.6) has infinitely many solutions. This yields infinitely many x Z m prim such that l(x) 0 for all l L and [F (x)] S = F (x) / b. Thus, part (ii) of Corollary 2.8 follows. We can improve on Corollary 2.8 if we assume condition (i) of Theorem B with L = L F, i.e., (2.9) rank L F = m, and for every Gal(K/Q)-proper subset M of L F with M L F we have ( ) L F [σ(m)] [L F \ σ(m)] and in addition to this, σ Gal(K/Q) (2.10) F (x) 0 for every non-zero x Q m. Let D be a non-zero Q-linear subspace of Q m. We say that a nonempty subset M of L F is linearly dependent on D if there is a nontrivial K-linear combination of the forms in M that vanishes identically on D; otherwise, M is said to be linearly independent on D. Further, for a non-empty subset M of L F we define rank D M to be the cardinality of a maximal subset of M that is linearly independent on D, and then l M q D (M) := e(l) rank D M. For instance, rank D L F = dim D, so q D (L F ) = deg F/ dim D. Then put q D (F ) := max{q D (M) : M L F, rank D M < dim D}. Finally, put (2.11) c(f ) := max D q D (F ) q D (L F ) = max D q D(F ) dim D deg F, where the maximum is taken over all Q-linear subspaces D of Q m with dim D 2. Lemma 5.2, which is stated and proved in Section 5 below, implies that if F satisfies both (2.9) and (2.10), then c(f ) < 1. We will not consider the problem how to compute c(f ), that is, how to

10 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY determine a subspace D for which q D (F )/q D (L F ) is maximal; this may involve some linear algebra that is beyond the scope of this paper. Given a decomposable form F Z[X 1,..., X m ], a finite set of primes S, and reals ɛ > 0, B > 0, we define N(F, S, ɛ, B) to be the set of x Z m prim with [F (x)] S F (x) ɛ and x B. Theorem 2.9. Let m 2 and let F Z[X 1,..., X m ] be a decomposable form as in (2.3) satisfying (2.9) and (2.10). Let c(f ) be defined as in (2.11). Then c(f ) < 1 and (i) for every finite set of primes S, every ɛ > 0 and every x Z m prim we have [F (x)] S F,S,ɛ F (x) c(f )+ɛ ; (ii) there are infinitely many primes p, and for each of these primes p infinitely many x Z m prim, such that [F (x)] {p} F,p F (x) c(f ) ; (iii) for every finite set of primes S and every ɛ with 0 < ɛ < 1 we have N(F, S, ɛ, B) F,S,ɛ B m(1 ɛ) as B. Assertions (i) and (iii) follow without too much effort from work in Liu s thesis [16], while (ii) is an application of Minkowski s Convex Body Theorem. The constants implied by the Vinogradov symbols in Theorems 2.7 and part (i) of Theorem 2.9 cannot be computed effectively from our method of proof. In fact, these constants can be expressed in terms of the heights of the subspaces occurring in certain instances of the p-adic Subspace Theorem, but for these we can as yet not compute an upper bound. The constant in (ii) can be computed once one knows a subspace D for which the quotient q D (F )/q D (L F ) is equal to c(f ). The work of Liu from which part (iii) is derived uses a quantitative version of the p-adic Subspace Theorem, giving an explicit upper bound for the number of subspaces. This enables one to compute effectively the constant in part (iii). Likely, a result of the same type as part (iii) of Theorem 2.9 can be proved in a similar way as Theorem 2.6 using the lattice point counting result of Barroero and Widmer, thereby avoiding Liu s work and the

S-PARTS OF VALUES OF POLYNOMIALS 11 quantitative Subspace Theorem. But such an approach would be less straightforward. 2.3. Effective results for decomposable forms. We consider again S-parts of values F (x), where F is a decomposable form in Z[X 1,..., X m ] and x Z m prim. Under certain stronger conditions on F, we shall give an estimate of the form [F (x)] S κ 6 F (x) 1 κ 5, with effectively computable positive κ 5, κ 6 that depend only on F and S. For applications, we make the dependence of κ 5 and κ 6 explicit in terms of S. The decomposable forms with the said stronger conditions include binary forms, and discriminant forms of an arbitrary number of variables. Let again S = {p 1,..., p s } be a finite set of primes and b an integer coprime with p 1 p s, and consider equation (2.4) in x Z m prim and non-negative integers a 1,..., a s. Under the stronger conditions for the decomposable form F mentioned above, explicit upper bounds were given in Győry [11, 12] for the solutions of (2.4), from which upper bounds can be deduced for [F (x)] S. Later, more general and stronger explicit results were obtained by Győry and Yu [15] on another version of (2.4). These explicit results provided some information on the arithmetical properties of F (x) at points x Z m prim. In this paper, we deduce from the results of Győry and Yu [15] a better bound for [F (x)] S ; see Theorem 2.10. This will give more precise information on the arithmetical structure of those non-zero integers h 0 that can be represented by F (x) at integral points x; see Corollary 7.1. To state our results, we introduce some notation and assumptions. Let F Z[X 1,..., X m ] be a non-zero decomposable form. Denote by K its splitting field. We choose a factorization of F into linear forms with coefficients in K as in (2.3), with L F a Gal(K/Q)-symmetric set of pairwise non-propertional linear forms. Denote by G(L F ) the graph with vertex set L F in which distinct l, l in L F are connected by an edge if λl + λ l + λ l = 0 for some l L F and some non-zero λ, λ, λ in K. Let L 1,..., L k be the vertex sets of the connected components of G(L F ). When k = 1 and L F has at least three elements, L F is said to be triangularly connected; see Győry and Papp [14].

12 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY In what follows, we assume that F in (2.4) satisfies the following conditions: (2.12) (2.13) L F has rank m; either k = 1; or k > 1 and X m can be expressed as a K-linear combination of the forms from L i, for i = 1,..., k. We note that these conditions are satisfied by binary forms with at least three pairwise non-proportional linear factors, and also discriminant forms, index forms and a restricted class of norm forms in an arbitrary number of variables. As has been explained in [8, Chap. 9], conditions (2.12), (2.13) imply condition (i) of Theorem B. As before, let S = {p 1,..., p s } be a finite set of primes, and put P := max p i. Further, let K denote the splitting field of F, and put 1 i s d := [K : Q]. Then we have Theorem 2.10. Under assumptions (2.12), (2.13), we have (2.14) [F (x)] S κ 6 F (x) 1 κ 5 for every x = (x 1,..., x m ) Z m prim with F (x) 0, and with x m 0 if k > 1, where κ 5 = ( ( c s 3 (P (log p1 ) (log p s ) ) d) 1 (c s 3 (2P (log P ) s ) d ) 1 and κ 6, c 3 are effectively computable positive numbers, depending only on F. It is easy to check that if F Z[X, Y ] is a binary form with at least three pairwise non-proportional linear factors over its splitting field, then it satisfies (2.12), (2.13) with m = 2 and k = 1. Thus, Theorem 2.5 follows at once from Theorem 2.10. We shall deduce Theorem 2.10 from a special case of Theorem 3 of Győry and Yu [15]. The constants κ 5, κ 6, c 3 could have been made explicit by using the explicit version of this theorem of Győry and Yu [15]. Further, Theorem 2.10 could be proved more generally, over number fields and for a larger class of decomposable forms. Weaker versions of Theorem 2.10 can be deduced from the results of Győry [11, 12].

S-PARTS OF VALUES OF POLYNOMIALS 13 3. Proofs of Theorems 2.1, 2.3, 2.4, 2.6 Let again S = {p 1,..., p s } be a finite, non-empty set of primes. We denote by the ordinary absolute value, and by p the p-adic absolute value with p p = p 1 for a prime number p. Further, we set Q := R, Q := C. The following result is a very well-known consequence of the p-adic Thue-Siegel-Roth Theorem. The only reference we could find for it is [18, Chap.IX, Thm.3]. For convenience of the reader we recall the proof. Proposition 3.1. Let F (X, Y ) Z[X, Y ] be a binary form of degree n 2 and of non-zero discriminant. Then F (x, y) [F (x, y)] S F,S,ɛ max( x, y ) n 2 ɛ for all ɛ > 0 and all (x, y) Z 2 prim with F (x, y) 0. Proof. We assume that F (1, 0) 0. This is no loss of generality. For if this is not the case, there is an integer b of absolute value at most n with F (1, b) 0 and we may proceed with the binary form F (X, bx + Y ). Our assumption implies that for each p S { } we have a factorization F (X, Y ) = a n i=1 (X β ipy ) with a Z and β ip Q p algebraic over Q for i = 1,..., n. For every (x, y) Z 2 prim with F (x, y) 0 we have F (x, y) ( [F (x, y)] S (max( x, y ) = n F,S F,S min 1 i n p S p S { } x β ip y p max( x p, y p ) p S { } min (1, x y β 1p p,..., x ) y β np p. F (x, y) p ) / max( x, y ) n The latter is F,S,ɛ max( x, y ) 2 ɛ for every ɛ > 0 by the p-adic Thue-Siegel-Roth Theorem. Proposition 3.1 follows. Proof of Theorem 2.1. Let f(x) Z[X] be the polynomial from Theorem 2.1. (i). The binary form F (X, Y ) := Y n+1 f(x/y ) has degree n + 1 and non-zero discriminant. Now by Proposition 3.1, we have for every ɛ > 0

14 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY and every sufficiently large integer x, f(x) [f(x)] S f,s,ɛ x n 1 nɛ f,s,ɛ f(x) (n 1 nɛ)/n, implying [f(x)] S f,s,ɛ f(x) (1/n)+ɛ. (ii). There are infinitely many primes p such that f(x) 0 (mod p) is solvable. Excluding the finitely many primes dividing the leading coefficient or the discriminant of f(x), there remain infinitely many primes. Take such a prime p. By Hensel s Lemma, there is for every positive integer k an integer x k such that f(x k ) 0 (mod p k ). We may choose such an integer with p k x k < 2p k. Then clearly, x 1 < x 2 < and for k sufficiently large, f(x k ) 0 and f(x k ) 0 (mod p k ). Consequently, [f(x k )] {p} p k 1 2 x k f f(x k ) 1/n. This proves Theorem 2.1. Proof of Theorem 2.4. Let F (X, Y ) Z[X, Y ] be the binary form from Theorem 2.4. (i) By Proposition 3.1, we have for every ɛ > 0 and every pair (x, y) Z 2 prim with F (x, y) 0 and max( x, y ) sufficiently large, F (x, y) [F (x, y)] S F,S,ɛ max( x, y ) n 2 nɛ F,S,ɛ F (x, y) 1 (2/n) ɛ. (ii) We assume that F (1, 0) 0 which, similarly as in the proof of Proposition 3.1, is no loss of generality. By Chebotarev s Density Theorem, there are infinitely many primes p such that F splits into linear factors over Q p. From these, we exclude the finitely many primes that divide D(F ) or F (1, 0). Let P be the infinite set of remaining primes. Then for every p P, we can express F (X, Y ) as F (X, Y ) = a n (X β ip Y ) i=1 with a Z with a p = 1, β ip Z p for i = 1,..., n and β ip β jp p = 1 for i, j = 1,..., n with i j. We distinguish two cases. First assume that F does not split into linear factors over Q. Take p P. Then without loss of generality,

S-PARTS OF VALUES OF POLYNOMIALS 15 β 1p Q. Let k be a positive integer. By Minkowski s Convex Body Theorem, there is a non-zero pair (x, y) Z 2 such that x β 1p y p p k, max( x, y ) p k/2. We may assume without loss of generality that gcd(x, y) is not divisible by any prime other than p. Assume that gcd(x, y) = p u with u 0, and let x k := p u x, y k := p u y. Then (x k, y k ) Z 2 prim and x k β 1p y k p p u k, max( x k, y k ) p (k/2) u. This clearly implies u k/2. We observe that if we let k then (x k, y k ) runs through an infinite subset of Z 2 prim. Indeed, otherwise we would have a pair (x 0, y 0 ) Z 2 prim with x 0 β 1p y 0 p p k/2 for infinitely many k which is impossible since β 1p Q. Next we have F (x k, y k ) 0 for all k. Indeed, suppose that F (x k, y k ) = 0 for some k. Then x k /y k = β ip for some i 2. Since β ip Z p we necessarily have y k p = 1. But then x k β 1p y k p = β ip β 1p p = 1, which is again impossible. Finally, since clearly x k β ip y k p 1 for i = 2,..., n, we derive that for each positive integer k, [F (x k, y k )] {p} = F (x k, y k ) 1 p p k u max( x k, y k ) 2 F,p F (x k, y k ) 2/n. Next, we assume that F (X, Y ) splits into linear factors over Q. Then F (X, Y ) = a n i=1 (X β iy ) with a Z, a p = 1 for p P, β i Q and β i p 1 for p P, i = 1,..., n, and β i β j p = 1 for p P, i, j = 1,..., n, i j. Pick distinct p, q P and let S = {p, q}. Then there is an integer u, coprime with pq, such that uβ 1, uβ 2 and u/(β 2 β 1 ) are all integers. Choose positive integers k, l. Then x := u(β 2p k β 1 q l ) β 2 β 1, y := u(pk q l ) β 1 β 2 are integers satisfying x β 1 y = up k, x β 2 y = uq l. By our choice of p, q P and by direct substitution, it follows that the numbers x β i y (i = 3,..., n) have p-adic and q-adic absolute values equal to 1. Thus, F (x, y) p = p k, F (x, y) q = q l and so [F (x, y)] S = p k q l. Clearly, g := gcd(x, y) is coprime with pq. Let x k,l := x/g, y k,l := y/g so that (x k,l, y k,l ) Z 2 prim. Then clearly, [F (x k,l, y k,l )] S = p k q l. We now

16 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY choose k, l such that p k, q l are approximately equal, say p k < q l < q p k. Then max( x k,l, y k,l ) max( x, y ) F,S (p k q l ) 1/2 and thus, [F (x k,l, y k,l )] S F,S max( x k,l, y k,l ) 2 F,S F (x k,l, y k,l ) 2/n. In the proofs of Theorems 2.3 and 2.6 we need a few auxiliary results. Lemma 3.2. Let f(x) Z[X] be a polynomial of non-zero discriminant and a an integer and p a prime. Denote by g p the largest nonnegative integer g such that p g divides the discriminant D(f) of f. For k > 0 denote by r(f, a, p k ) the number of congruence classes x modulo p k with f(x) 0 (mod p k ), x a (mod p). Then r(f, a, p k ) = r(f, a, p gp+1 ) for k g p + 1. Proof. This is a consequence of [25, Thm. 2]. Given a positive integer h, we say that two pairs (x 1, y 1 ), (x 2, y 2 ) Z 2 prim are congruent modulo h if x 1 y 2 x 2 y 1 (mod h). With this notion, for a given binary form F (X, Y ) Z[X, Y ] we can divide the solutions (x, y) Z 2 prim of F (x, y) 0 (mod h) into congruence classes modulo h. Lemma 3.3. Let F (X, Y ) Z[X, Y ] be a binary form of degree n 2 and of non-zero discriminant and p a prime. Denote by g p the largest non-negative integer g such that p g divides the discriminant D(F ) of F. For k > 0 denote by r(f, p k ) the number of congruence classes modulo p k of (x, y) Z 2 prim with F (x, y)) 0 (mod p k ). Then r(f, p k ) = r(f, p gp+1 ) for k g p + 1. Proof. Neither the number of congruence classes under consideration, nor the discriminant of F, changes if we replace F (X, Y ) by F (ax + by, cx + dy ) for some matrix ( ) a b c d GL2 (Z). After such a replacement, we can achieve that F (1, 0)F (0, 1) 0, so we assume this henceforth. Let f(x) := F (X, 1) and f (X) := F (1, X). The map (x, y) x y 1 (mod p k ) gives a bijection between the congruence classes modulo p k of pairs (x, y) Z 2 prim with F (x, y) 0 (mod p k ) and y 0 (mod p) and the congruence classes modulo p k of integers z with f(z) 0 (mod p). Likewise, the map (x, y) y x 1 (mod p k ) establishes a bijection between the congruence classes modulo p k of (x, y) Z 2 prim with F (x, y) 0 (mod p k ) and y 0 (mod p) and the congruence classes modulo p k of integers z with f (z) 0 (mod p k ) and

S-PARTS OF VALUES OF POLYNOMIALS 17 z 0 (mod p). Further, our assumption F (1, 0)F (0, 1) 0 implies that D(F ) = D(f) = D(f ). Now an application of Lemma 3.2 yields that r(f, p k ) = p 1 a=0 r(f, a, pk ) + r(f, 0, p k ) is constant for k g p + 1. For a binary form F (X, Y ) R[X, Y ] and for positive reals B, M, we denote by V F (B, M) the set of pairs (x, y) R 2 with max( x, y ) B and F (x, y) M, and by µ F (B, M) the area (two-dimensional Lebesgue measure) of this set. Our next lemma is a consequence of a general lattice point counting result of Barroero and Widmer [1, Thm. 1.3]. Lemma 3.4. let n be an integer 2. Then there is a constant c(n) > 0 such that for every non-zero binary form F (X, Y ) R[X, Y ] of degree n, every lattice Λ Z 2 and all positive reals B, M, #(V F (B, M) Λ) µ F (B, M) det Λ c(n) max(1, B/m(Λ)), where m(λ) is the length of the shortest non-zero vector of Λ. Proof. We write points in R n+3 R 2 as (z 0,..., z n, u, v, x, y). The set Z R n+3 R 2 given by the inequalities z 0 x n + z 1 x n 1 y + + z n y n v, x u, y u is a definable family in the sense of [1], parametrized by the tuple T = (z 0,..., z n, u, v). By substituting for this tuple the coefficients of F, respectively B and M, we obtain the set V F (B, M) as defined above. The sum of the one-dimensional volumes of the orthogonal projections of V F (B, M) on the x-axis and y-axis is at most 4B, and the first minimum of Λ is m(λ). Now Lemma 3.4 follows directly from [1, Thm. 3.1]. A lattice Λ Z 2 is called primitive if it contains points (x, y) Z 2 prim. Lemma 3.5. Let again n be an integer 2. Then there is a constant c (n) > 0 such that for every binary form F Z[X, Y ] of degree n, every primitive lattice Λ Z 2, and all reals B, M > 1, #( ) ( 6 ) V F (B, M) Λ Z 2 prim (1 + p 1 ) 1 µf (B, M) π 2 det Λ c (n)b log 3B. p det Λ

18 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY Proof. In the proof below, p, p i denote primes. Let F (X, Y ) Z[X, Y ] be a binary form, Λ Z 2 a primitive lattice, and B, M reals > 1. Put d := det Λ. For a positive integer h, define the lattice Λ h := Λ hz 2. Since Λ is primitive, there is a basis {a, b} of Z 2 such that {a, db} is a basis of Λ. Hence {ha, lcm(h, d)b} is a basis of Λ h, and so (3.1) det Λ h = h lcm(h, d) = d h 2 gcd(h, d). Further, the shortest non-zero vector of Λ h has length (3.2) m(λ h ) h. We define ρ(h) := #(V F (B, M) Λ h ). Then by the rule of inclusion and exclusion, ( ) # V F (B, M) Λ Z 2 prim = ρ(1) ρ(p) + ρ(p 1 p 2 ) p B = h B µ(h)ρ(h), p 1 <p 2 : p 1 p 2 B where µ(h) denotes the Möbius function. The previous lemma together with (3.1), (3.2) implies ( ) # V F (B, M) Λ Z 2 prim µ F (B, M) gcd(d, h) µ(h) d h 2 c(n)b h B µ(h) h, h B hence ( ) # V F (B, M) Λ Z 2 prim µ F (B, M) µ(h) d µ F (B, M) d µ(h) h>b c (n)b log 3B, gcd(d, h) h 2 h=1 + c(n)b gcd(d, h) h 2 h B µ(h) h

S-PARTS OF VALUES OF POLYNOMIALS 19 where we have used h>b µ(h) gcd(d,h) h 2d/B, µ 2 F (B, M) 4B 2, and µ(h) h B h log 3B. Now the proof is finished by observing that gcd(d, h) µ(h) = (1 p 1 ) (1 p 2 ) = 6 h 2 π (1 + p 1 ) 1. 2 h=1 p d p d p d Lemma 3.6. Let α 1,..., α t be positive reals. Denote by N(A) the number of tuples of non-negative integers (u 1,..., u t ) with (3.3) A α 1 u 1 + + α t u t A + 2(α 1 + + α t ). Then N(A) t,α1,...,α t A t 1 as A. Proof. Constants implied by the Vinogradov symbols, will depend on t, α 1,..., α t. For u = (u 1,..., u t ) Z t, denote by C u the cube in R t consisting of the points y = (y 1,..., y t ) with u i y i < u i + 1 for i = 1,..., t. Let C be the union of the cubes C u over all points u with non-negative integer coordinates satisfying (3.3). Put α := α 1 + + α t. Then C 1 C C 2, where C 1, C 2 are the subsets of R s given by A + α α 1 y 1 + + α t y t A + 2α, y 1 0,..., y t 0, A α 1 y 1 + + α t y t A + 3α, y 1 0,..., y t 0, respectively. Clearly N(A) is estimated from below and above by the measures of C 1 and C 2, the first being (A + 2α) t (A + α) t A t 1, the second being A t 1. The lemma follows. We first give the complete proof of Theorem 2.6. The proof of Theorem 2.3 is then obtained by making a few modifications. Proof of Theorem 2.6. Let F (X, Y ) Z[X, Y ] be a binary form of degree n 3 with non-zero discriminant, ɛ a real with 0 < ɛ < 1 n and S = {p 1,..., p s } a finite set of primes. Let S = {p 1,..., p s } be the set of p S such that F (x, y) 0 (mod p gp+1 ) has a solution in Z 2 prim, and let S = {p s +1,..., p s } be the set of remaining primes. In what follows, constants implied by Vinogradov symbols, and by the Landau O-symbol will depend only on F, S and ɛ.

20 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY We first prove that N(F, S, ɛ, B) F,S,ɛ B 2 nɛ (log B) s 1 as B. The set of pairs (x, y) under consideration can be partitioned into sets N h, where h runs through the set of positive integers composed of primes from S, and N h is the set of pairs (x, y) Z 2 prim with max( x, y ) B, [F (x, y)] S = h, F (x, y) h 1/ɛ. We first estimate #N h from above by means of Lemma 3.5 where h is any positive integer composed of primes from S. Notice that for (x, y) N h we have F (x, y) 0 (mod h). By Lemma 3.3 and the Chinese Remainder Theorem, the set of these (x, y) lies in 1 congruence classes modulo h. Each of these congruence classes is precisely the set of primitive points in a set of the shape {(x, y) Z 2 : y 0 x x 0 y (mod h)} with (x 0, y 0 ) Z 2 prim, which is a primitive lattice of determinant h. So N h is contained in 1 primitive lattices of determinant h. We next estimate the area µ F (B, h 1/ɛ ) of V (B, h 1/ɛ ). There is a constant c F > 0 such that (3.4) F (x, y) c F (max( x, y ) n for (x, y) R 2. If h (c F B n ) ɛ then the condition F (x, y) h 1/ɛ is already implied by max( x, y ) B, and so µ F (B, h 1/ɛ ) = 4B 2. On the other hand, if h < (c F B n ) ɛ, we have, denoting by µ the area, µ F (B, h 1/ɛ ) µ ( {(x, y) R 2 : F (x, y) h 1/ɛ } ) = h 2/nɛ µ ( {(x, y) R 2 : F (x, y) 1} ) h 2/nɛ, since the set of (x, y) R 2 with F (x, y) 1 has finite area (see for instance [17]). Now invoking Lemma 3.5, we infer { B (3.5) #N h 2 /h + O(B log B) if h (c F B n ) ɛ, h (2/nɛ) 1 + O(B log B) if h < (c F B n ) ɛ. Finally, from (3.4) it is clear that N h = if h > c F B n. Let α := log(p 1 p s ). For j Z, let M j be the union of the sets N h with (3.6) e 2jα (c F B n ) ɛ h < e (2j+2)α (c F B n ) ɛ.

S-PARTS OF VALUES OF POLYNOMIALS 21 We restrict ourselves to j with (3.7) e 2jα (c F B n ) ɛ c F B n, e (2j+2)α (c F B n ) ɛ > 1, since for the remaining j the set M j is empty. Thus, (3.8) N(F, S, ɛ, B) j #M j, where the summation is over j with (3.7). We estimate the number of h with (3.6). Write h = h p u 1 1 p u s s where h is composed of primes from S. Then h divides p S p gp, so we have 1 possibilities for h. By applying Lemma 3.6 with t = s, A = e 2jα (c F B n ) ɛ )/h, α i = log p i for i = 1,..., s, we infer from Lemma 3.6 that for given h the number of possibilities for (u 1,..., u s ) is (log B) s 1. Hence the number of h with (3.6) is (log B) s 1. Now from (3.5) it follows that for j with (3.7), { e #M j 2jα B 2 nɛ (log B) s 1 + O(B(log B) s ) if j 0, e 2 j α((2/nɛ) 1) B 2 nɛ (log B) s 1 + O(B(log B) s ) if j < 0. Finally, from these estimates and (3.8) we deduce, taking into consideration that the number of j with (3.7) is log B, and also our assumption 0 < ɛ < 1 n, ( N(F, S, ɛ, B) e 2jα + ) e 2 j α((2/nɛ) 1) B 2 nɛ (log B) s 1 j<0 We next prove that j 0 B 2 nɛ (log B) s 1. N(F, S, ɛ, B) B 2 nɛ (log B) s 1 as B. +O(B(log B) s +1 ) For i = s +1,..., s, let a i be the largest integer u such that F (x, y) 0 (mod p u i ) is solvable in (x, y) Z 2 prim. Let for the moment h be any integer of the shape h = p u 1 1 p us s where u i g pi + 1 for i = 1,..., s and u i = a i for i = s + 1,..., s, and where h (c F B n ) ɛ. By Lemma 3.5 and the Chinese Remainder Theorem, the number of congruence classes modulo h of (x, y) Z 2 prim with F (x, y) 0 (mod h) is s r := r(f, p gp i +1 i ) i=1 s i=s +1 r(f, p a i i ),

22 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY which is independent of h. As mentioned above, each of these congruence classes is just the set of primitive points in a primitive lattice of determinant h. Furthermore, since these lattices arise from different residue classes modulo h of points in Z 2 prim, the intersection of any two of these lattices does not contain points from Z 2 prim anymore. Since moreover by our assumption h (c F B n ) ɛ the set V (B, h 1/ɛ ) has area (4B) 2, an application of Lemma 3.5 yields that the set of (x, y) Z 2 prim with max( x, y ) B, F (x, y) h 1/ɛ and F (x, y) 0 (mod h) has cardinality cr (4B)2 + O(B log B), h where c = (6/π 2 ) p S 0 (1 + p 1 ) 1, with S 0 the set obtained from S by removing those primes p i from S for which a i = 0. By the rule of inclusion and exclusion, the set N h, i.e., the set of (x, y) Z 2 prim as above with F (x, y) divisible by h but not by hp for p S, has cardinality (3.9) cr (4B)2 h p S cr (4B)2 ph + = cr p S (1 p 1 ) (4B)2 h p,q S,p<q + O(B log B) B2 h cr (4B)2 pqh +O(B log B) + O(B log B). We now consider the set of integers h of the shape p u 1 1 p us s with u i g pi + 1 for i = 1,..., s and u i = a i for i = s + 1,..., s, and with (c F B n ) ɛ h e 2α (c F B n ) ɛ, where again α = log(p 1 p s ). By Lemma 3.6, there are (log B) s 1 such integers. Using again 0 < ɛ < 1 n, it follows that N(F, S, ɛ, B) h #N h B 2 nɛ (log B) s 1. This completes the proof of Theorem 2.6. Proof of Theorem 2.3. Let f Z[X] be a polynomial of degree n 2 with non-zero discriminant, ɛ a real with 0 < ɛ < 1 n and S = {p 1,..., p s } a finite set of primes. Similarly as above S = {p 1,..., p s } is the set of p S such that f(x) 0 (mod p gp+1 ) is solvable in Z and S = {p s +1,..., p s }.

S-PARTS OF VALUES OF POLYNOMIALS 23 The proof is the same as that of Theorem 2.6 except from a few small modifications. The main difference is that instead of Lemma 3.5 we use the simple observation that if V f (B, M) is the set of x R with x B and f(x) M and µ f (B, M) is the one-dimensional measure of this set, then for all a, h Z with h > 0, the number of integers x V f (B, M) with x a (mod h) is (3.10) µ f (B, M)/h + error term, with error term c(n) for some quantity c(n) depending only on n = deg f. We first prove that (3.11) N(f, S, ɛ, B) f,s,ɛ B 1 nɛ (log B) s 1 as B. Let c f be a constant such that f(x) c f x n for x R. Consider the set N h of integers x with x B, [f(x)] S = h and f(x) h 1/ɛ. Then if h (c f B n ) ɛ we have µ f (B, h 1/ɛ ) = 2B, while otherwise, µ f (B, h 1/ɛ ) h 1/nɛ, since f(x) x n if x 1. Now a similar computation as in the proof of Theorem 2.6, using Lemma 3.2 instead of Lemma 3.3, gives instead of (3.5), { B/h + O(1) if h (cf B #N h n ) ɛ, h (1/nɛ) 1 + O(1) if h < (c f B n ) ɛ, and then the proof of (3.11) is completed in exactly the same way as in the proof of Theorem 2.6. The proof of (3.12) N(f, S, ɛ, B) f,s,ɛ B 1 nɛ (log B) s 1 as B follows the same lines as that of Theorem 2.6. For i = s + 1,..., s let a i be the largest integer u such that f(x) 0 (mod p a i i ) is solvable. Let h = p u 1 1 p us s with u i g pi + 1 for i = 1,..., s and u i = a i for i = s + 1,..., s, and with h (c f B n ) ɛ. Then by combining (3.10) with Lemma 3.2 one obtains that the set of integers x with x B, f(x) 0 (mod h) and f(x) h 1/ɛ has cardinality rb/h + O(1) with r > 0 depending only on f, and then an inclusion and exclusion argument gives #N h B/h + O(1).

24 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY Again, an argument completely similar to that in the proof of Theorem 2.6 gives (3.12). 4. Proof of Theorem 2.7 The theorem can be proved by modifying the arguments from [4]. We prefer to follow [6, 8], which already contains the basic ideas. Let F Z[X 1,..., X m ] be a decomposable form of degree n with splitting field K. We take a factorization of F as in (2.3). Assume that F satisfies condition (i) of Theorem B. Let D be a linear subspace of Q m of dimension 2. Denote by D the K-vector space of linear forms in K[X 1,..., X m ] that vanish identically on D. Then a set of linear forms in K[X 1,..., X m ] is linearly dependent on D if some non-trivial K-linear combination of these forms belongs to D and linearly independent on D if no such linear combination exists. The D-rank rank D M of a set of linear forms M K[X 1,..., X m ], is the maximal number of linear forms in M that are linearly independent on D. We have rank D L F = dim D. We call a subset I of L F minimally linearly dependent on D, if I itself is linearly dependent on D, but every proper, non-empty subset of I is linearly independent on D. We define a(n undirected) graph G D as follows. The set of vertices of G D is L F ; and {l, l } is an edge of G D if there is a subset of L F that is minimally linearly dependent on D and contains both l and l. Clearly, if {l, l} is an edge of G D, then so is {σ(l), σ(l )} for each σ Gal(K/Q), i.e., each σ acts on G D as an automorphism. Lemma 4.1. Let D be a linear subspace of Q m of dimension 2 such that none of the linear forms in L vanishes identically on D. Then G D is connected. Proof. Assume that G D is not connected. Let M be the vertex set of a connected component of G D. Then M L F. Clearly, for each σ Gal(K/Q), σ(m) is also the vertex set of a connected component of G D, hence either σ(m) = M, or σ(m) M =. That is, M is Gal(K/Q)-proper. By assumption (i) of Theorem B, the K-vector space [σ(m)] [L F \ σ(m)] σ Gal(K/Q)

S-PARTS OF VALUES OF POLYNOMIALS 25 contains a linear form from L, which, by assumption, does not lie in D. Hence there is σ Gal(K/Q) such that [σ(m)] [L F \ σ(m)] contains a linear form outside D. But since D is defined over Q, we have σ(d ) = D and so [M] [L F \ M] contains a linear form not in D, say l 0. Take maximal subsets M 1, M 2 of M and L F \ M, respectively, that are both linearly independent on D. Then there are λ l K for l M 1 M 2 such that λ l l λ l l l 0 (mod D ). l M 1 l M 2 This implies that M 1 M 2 is linearly dependent on D. We can take a subset of M 1 M 2 that is minimally linearly dependent on D. This set necessarily must have elements with both M 1 and M 2 in common. But then there would be an edge connecting an element of M with one of L F \ M, which contradicts that M is the vertex set of a connected component of G D. Lemma 4.2. Let D be a linear subspace of Q m of dimension d 2 and M a non-empty subset of L F with rank D M < d. Then there is a subset I of L F that is minimally linearly dependent on D, such that M I and rank D M I > rank D M. Proof. Let M consist of all linear forms in L F that are linear combinations of the linear forms in M and of the linear forms in D. Then rank D M = rank D M < d, hence M L F. Take a maximal subset M 1 of M that is linearly independent on D; then it is also a maximal subset of M that is linearly independent on D. Let M 2 be a maximal subset of L F \ M that is linearly independent on D. By Lemma 4.1 there is a set J L F that is minimally linearly dependent on D and contains elements of both M and L F \ M. This gives a linear combination l J λ ll D, with l 0 := l J M λ l l D. Writing the linear forms in J M as linear combinations modulo D of the linear forms in M 1, and the linear forms in J (L F \ M ) as linear combinations modulo D of the linear forms in M 2, we obtain a relation l M 1 M 2 µ l l D, with l M 1 µ l l l 0 0 (mod D ). Hence M 1 M 2 is linearly dependent on D. Take a subset I of M 1 M 2 that is minimally linearly dependent on D. We have I M 1 and I M 2 since M 1 and M 2 are linearly independent on D. This implies I M. Further, M 2 M =, therefore each of the linear

26 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY forms in M 2 is linearly independent on D of the linear forms in M. Hence rank D M I > rank D M. Denote by M K the set of places of K. We choose normalized absolute values v (v M K ) in such a way that if v lies above p { } {primes}, then x v = x p [Kv:Qp]/[K:Q] for x Q. These absolute values satisfy the product formula v M K x v = 1 for x K. For a vector y = (y 1,..., y r ) K r, we define y v := max 1 i r y i v (v M K ), H(y) := v M K y v. By the product formula, H(λy) = H(y) for y K r, λ K. For x Z m prim and a subset I of L F, we define H I (x) := max l(x) v. l I v M K Lemma 4.3. Let x Z m prim with l(x) 0 for l L F and let I, J be subsets of L F with I J. Then H I J (x) H I (x) H J (x). Proof. Let l 0 I J. Then by the product formula, H I J (x) = max l(x)/l 0(x) v l I J v M K ( ) ( max l(x)/l ) 0(x) v max l(x)/l 0(x) v l I l J v M K v M K = H I (x) H J (x). Lemma 4.4. Let D be a linear subspace of Q m of dimension 2 on which none of the linear forms in L vanishes identically. Then for every x Z m prim D with l(x) 0 for l L F, there is a subset I of L F that is minimally linearly dependent on D such that H I (x) F,D x 1/(m 1). Proof. Let x Z m prim D with l(x) 0 for l L F. Start with a linear form l 0 L F. By Lemma 4.2, there is a subset I 1 of L F that is minimally linearly dependent on D, that contains l 0 and for which rank D I 1 2. Using Lemma 4.2, we choose inductively subsets I 2, I 3,... of L F that are minimally linearly dependent on D as follows: if

S-PARTS OF VALUES OF POLYNOMIALS 27 rank D I 1 I t < dim D, we choose I t+1 such that I t+1 (I 1 I t ) and rank D I 1 I t+1 > rank D I 1 I t. It is clear that for some s dim D 1 m 1 we get rank D I 1 I s = dim D. Then X 1,..., X m can be expressed as linear combinations modulo D of the linear forms in I 1 I s, implying x = H(x) F,D H I1 I s (x). Now from Lemma 4.3 we infer x F,D s i=1 H Ii (x) max 1 i s H I i (x) s max 1 i s H I i (x) m 1. Proof of Theorem 2.7. Without loss of generality, we assume that the linear forms in (2.3) have their coefficients in the ring of integers O K of K. We prove by induction on d that if D is a linear subspace of Q m of dimension d on which none of the linear forms in L vanishes identically, then (2.7) has only finitely many solutions in Z m prim D. For d = 1 this is clear. Assume that d 2 and that our assertion holds true for all linear subspaces of Q m of dimension smaller than d. Let D be a linear subspace of Q m of dimension d on which none of the linear forms in L vanishes identically and let 0 < ɛ < 1 m 1. Take x Zm prim D satisfying (2.7). Choose a subset I of L F that is minimally linearly dependent on D such that (4.1) H I (x) F,D x 1/(m 1). Let T be the set of places of K lying above the places in S { }. For v T, choose l v I such that l v (x) v = max l I l(x) v, and let I v := I \ {l v }. We have l v (x) v F 1 for l L F, v M K \ T since x Z m. So by the product formula, v T l(x) v F 1 for l L F. Together with (4.1) this implies l(x) v F F (x) v x (1/(m 1)) ɛ v T l I v T F,D H I (x) 1 (m 1)ɛ, and subsequently, dividing both sides by v T l v(x) v, (4.2) l(x) v F,D H I (x) (m 1)ɛ. v T l I v

28 Y. BUGEAUD, J.-H. EVERTSE, AND K. GYŐRY Write I = {l 0,..., l u }. Then l 0 β 1 l 1 + + β u l u (mod D ) with β i 0 for i = 1,..., u Put y i := l i (x) for i = 1,..., u, and y = (y 1,..., y u ). Then y OK u. We can express l(x) (l I v) as u linearly independent linear forms in y, say l 1,v (y),..., l u,v (y), taken from the set {Y 1,..., Y u, β 1 Y 1 + + β u Y u }. Now (4.2) translates into u l i,v (y) v F,D H(y) (m 1)ɛ, y OK. u v T i=1 Thus, we can apply the p-adic Subspace Theorem [22], and conclude that the vectors y lie in finitely many proper linear subspaces of K u. It follows that the solutions x Z m prim D of (2.7), corresponding to the same sets I v (v T ) in (4.2), lie in finitely many proper linear subspaces of D. Since there are only finitely many possibilities for the sets I v, it follows that the solutions x Z m prim D altogether lie in only finitely many proper linear subspaces of D. By applying the induction hypothesis to each of these spaces, it follows that (2.7) has only finitely many solutions in Z m prim D. This completes our proof. 5. Proof of Theorem 2.9 Let F Z[X 1,..., X m ] be a decomposable form in m 2 variables with a factorization as in (2.3), satisfying (2.9) and (2.10). Our first goal is to prove that c(f ) < 1. We have used some arguments from [16, 3.3]. We start with some preparations. For a subset M of L F we put M := l M e(l). Let D be a linear subspace of Q m of dimension d 2. A subset M of L F is called D- critical if q D (M) is maximal among all non-empty subsets of L F. A D-critical subset is called minimal if none of its proper subsets is D- critical. Lemma 5.1. Let M 1, M 2 be two D-critical subsets of L F. (i) Assume that M 1, M 2 are minimal and M 1 M 2. Then M 1 M 2 =. (ii) Assume that M 1 M 2 =. Then M 1 M 2 is D-critical. Proof. We use that for any two subsets N 1, N 2 of L F we have (5.1) { rankd N 1 N 2 + rank D N 1 N 2 rank D N 1 + rank D N 2, N 1 N 2 + N 1 N 2 = N 1 + N 2.