DIOPHANTINE APPROXIMATION FOR ALGEBRAIC POINTS ON CURVES By Thomas Plante A DISSERTATION to Michigan State University in partial fulfillment of the requirements Submitted for the degree of Mathematics – Doctor of Philosophy 2020 DIOPHANTINE APPROXIMATION FOR ALGEBRAIC POINTS ON CURVES ABSTRACT By Thomas Plante A foundational result in Diophantine approximation is Roth’s theorem, which asserts that for a given algebraic number (cid:11) and " > 0, the inequality j(cid:11) (cid:0) p(cid:157)qj < 1(cid:157)q2+" has only finitely many solutions in rational numbers p(cid:157)q 2 Q. Ridout and Lang subsequently proved a general form of Roth’s theorem allowing for arbitrary absolute values (including p-adic absolute values) and permitting arbitrary (fixed) number fields k in place of the rational numbers. Wirsing proved a generalization of Roth’s theorem, where the approximating elements are taken from varying number fields of degree d, and the quantity 2 + " in Roth’s theorem is replaced by 2d + ". As a consequence of a deep inequality of Vojta, Wirsing’s theorem, appropriately formulated, may be extended to a Diophantine approximation result for algebraic points of degree d on a nonsingular projective curve. The main theorem of this thesis improves on this general form of Wirsing’s theorem further, but only in the cases where d = 2 and d = 3. More specifically, if we let C be a nonsingular projective curve over a number field k, S a finite set of places of k and P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi, then letting R 2 C„k” and " > 0, we get ∑n mD;S„P” (cid:20) „Nd„D” + "”hR„P” + O„1” for all P 2 C„k” with »k„P” : k… = d 2 f2; 3g, where mD;S„P” is a sum of local heights associated to D and S, hR is a global height associated to R, and {(cid:12)(cid:12)(cid:12)( Nd„D” := max ) (cid:12)(cid:12)(cid:12)} (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ Supp„D” taken over all finite k-morphisms (cid:27) : C ! P1 of degree d. As Nd„D” (cid:20) 2d, the main result gives a refinement of Wirsing’s theorem depending on the divisor D. In much the same way that Roth’s theorem can be used to prove Siegel’s Theorem about integral points on a genus-0 affine curve, the main theorem of this dissertation implies Levin’s generalization of Siegel’s theorem for algebraic points of degree d in the cases d = 2; 3. After a review of some properties of global heights and local heights, we also show that the main theorem is sharp, in the sense that counterexamples exist when Nd„D” + " is replaced with Nd„D” (cid:0) ". To prove the main theorem, we first show that for curves of large enough genus, the number of morphisms (cid:27) : C ! P1 of degree d = 2; 3 defined over k is finite. We then prove that curves for which the number of such morphisms is finite satisfy the main theorem. Finally, we prove the main theorem for the remaining small genus curves on a case by case basis. We transfer the Diophantine approximation problem for points of degree d on C to a Diophan- tine approximation problem for rational points on Symd„C” (the dth symmetric power of C). We exploit the map from Symd„C” to Jac„C”, the Jacobian of C, and its associated geometry. We use Diophantine approximation results for abelian varieties (when we’re on Jac„C”), Diophantine approximation results for projective spaces (fibers of Symd„C” 7! Jac„C”), and Diophantine ap- proximation results on Symd„C” directly, plus algebraic geometry to connect all of these. In the course of the proof we make use of Schmidt’s subspace theorem and its generalizations due to Ru-Wong and Evertse-Ferretti as well as a version of Roth’s theorem for nonreduced divisors and Faltings’ approximation theorem for rational points on abelian varieties. TABLE OF CONTENTS CHAPTER 1 AN INTRODUCTION TO DIOPHANTINE APPROXIMATION . . . . 1.1 The Road to Roth’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Further Approximation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Integral Points on Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 4 7 CHAPTER 2 REVIEW OF HEIGHTS . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.1 Heights on Projective Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2 Heights on Projective Varieties . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.3 Heights on Closed Subschemes . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 CHAPTER 3 PREPARATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 . 22 3.1 Useful Results 3.2 Sharpness of the Main Result . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 3.3 Facts on Trigonal Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 CHAPTER 4 PROOF OF THE MAIN THEOREM . . . . . . . . . . . . . . . . . . . 31 4.1 Algebraic Points on Curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 4.2 Quadratic Points on Elliptic Curves . . . . . . . . . . . . . . . . . . . . . . . . 37 4.3 Cubic Points on Curves of Low Genus . . . . . . . . . . . . . . . . . . . . . . . 45 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . BIBLIOGRAPHY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 iv CHAPTER 1 AN INTRODUCTION TO DIOPHANTINE APPROXIMATION 1.1 The Road to Roth’s Theorem An important question in Diophantine approximation is how closely a given real number (cid:11) 2 R can be approximated by rational numbers p(cid:157)q 2 Q; in particular, that to approximate (cid:11) closely both the numerator, p, and denominator, q, must be large. A typical way of expressing this is to ask, for a given (cid:11) 2 R and e > 0, whether or not the inequality has infinitely many solutions p(cid:157)q 2 Q. In 1842, it was shown by Dirichlet that Theorem 1.1.1. For each (cid:11) 2 R n Q the inequality Theorem 1.1.2 (Liouville). If (cid:11) is an algebraic number of degree d and if e > d, then the inequality In fact, given an algebraic number (cid:11) of degree d, you can find an effective (explicitly described) constant C = C„(cid:11)” > 0 such that the inequality q (cid:12)(cid:12)(cid:12)(cid:12) p (cid:12)(cid:12)(cid:12)(cid:12) p q (cid:12)(cid:12)(cid:12)(cid:12) p (cid:12)(cid:12)(cid:12)(cid:12) p q jqje (cid:12)(cid:12)(cid:12)(cid:12) (cid:20) 1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:20) 1 q2 (cid:0) (cid:11) (cid:0) (cid:11) (cid:12)(cid:12)(cid:12)(cid:12) (cid:20) 1 jqje (cid:0) (cid:11) (cid:12)(cid:12)(cid:12)(cid:12) (cid:21) C jqjd (cid:0) (cid:11) has infinitely many solutions p(cid:157)q 2 Q. Meanwhile, in 1844, Liouville showed that q has at most finitely many solutions p(cid:157)q 2 Q. holds for all p(cid:157)q 2 Q n f(cid:11)g. 1 ∑1 i=1 10(cid:0)i!. Then the partial sums ∑n Example 1.1.3. Let (cid:11) = 1jqje as long as n (cid:21) e. By Liouville’s theorem, (cid:11) is transcendental. i=1 10(cid:0)i! are solutions to (cid:12)(cid:12)(cid:12) (cid:20) (cid:0) (cid:11) (cid:12)(cid:12)(cid:12) p q In order to start getting nice applications to Diophantine equations, however, one needs an exponent of d (cid:0) ". Theorem 1.1.4 (Thue, 1909). Let (cid:11) 2 Q, »Q„(cid:11)” : Q… = d. Let " > 0. Then there are only finitely many rational numbers p q (cid:12)(cid:12)(cid:12)(cid:12) < 2 Q satisfying(cid:12)(cid:12)(cid:12)(cid:12)(cid:11) (cid:0) p ( )3 (cid:0) 2 = ( q x y 1 d 2 +1+" q )(( : )2 (cid:0) 3p 2 x y 3p 2 + 3p 4 x y + x y ) Example 1.1.5. Consider integer solutions to the equation x3 (cid:0) 2y3 = 1. Then (cid:12)(cid:12)(cid:12) x = 1 y3 (cid:12)(cid:12)(cid:12) < C (cid:0) 3p Which implies 2 many solutions x; y 2 Z. y y3 for some constant C > 0. By Thue’s theorem, there are only finitely More generally, Thue proved that if f 2 Z»x; y… is an irreducible binary form with deg„ f ” (cid:21) 3, and r 2 Z, f „x; y” = r has only finitely many integer solutions x and y. p The exponent d p d by Siegel (1921) 2d by Dyson and Gelfond (independently) (1947). In 1955 Roth managed to eliminate the 2 + 1 in Thue’s theorem was subsequently improved to 2 and dependence on d altogether. Theorem 1.1.6 (Roth’s Theorem). For every algebraic number (cid:11) and every " > 0, the inequality q has only finitely many solutions p(cid:157)q 2 Q. C = C„(cid:11); "” > 0 such that for all p(cid:157)q 2 Q n f(cid:11)g (cid:0) (cid:11) (cid:12)(cid:12)(cid:12)(cid:12) p (cid:12)(cid:12)(cid:12)(cid:12) p q (cid:12)(cid:12)(cid:12)(cid:12) (cid:20) (cid:0) (cid:11) 1 jqj2+" (cid:12)(cid:12)(cid:12)(cid:12) (cid:21) C : jqj2+" 2 Equivalently, for every real algebraic number (cid:11) and every " > 0, there exists a constant note here that Roth’s Theorem only asserts the existence of such a constant C and provides no way of calculating it, that is, C is not effective. Another equivalent expression of Roth’s Theorem is that for every real algebraic number (cid:11) and every " > 0, there exists a constant C = C„(cid:11); "” > 0 such that for all p(cid:157)q 2 Q n f(cid:11)g log max (cid:0) (cid:11) (cid:20) „2 + "” log max fjpj; jqjg + C: ; 1 {(cid:12)(cid:12)(cid:12)(cid:12) p q } (cid:12)(cid:12)(cid:12)(cid:12)(cid:0)1 This form is significant, because it is written in terms of functions called heights. The left hand side of the inequality is what’s known as a local height of p(cid:157)q, relative to the algebraic number (cid:11), which for now we denote (cid:21)(cid:11)„p(cid:157)q”. note that the closer p(cid:157)q gets to (cid:11), the larger the local height becomes, so we think of local height like an inverse distance function. The height of a rational number p(cid:157)q, with gcd„p; q” = 1, is h„p(cid:157)q” := log max fjpj; jqjg. Height can be thought of as a measure of the arithmetic complexity of a number. Thus, using big-O notation, we may restate Roth’s theorem: For every algebraic number (cid:11) and every " > 0, the inequality (cid:21)(cid:11)„(cid:12)” (cid:20) „2 + "”h„(cid:12)” + O„1” holds for all (cid:12) 2 Q. Generalizing to an arbitrary absolute value j(cid:1)jv on a number field k (normalized as in Definition 2.1.3), we define the following: Definition 1.1.7. Let k be a number field. Given (cid:11) 2 k and an absolute value j (cid:1) jv on k (with some fixed extension to k), define the local height associated to (cid:11) at v by (cid:21)(cid:11);v„(cid:12)” := log max for all (cid:12) 2 k. } {(cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12) (cid:0) (cid:11) ; 1 v Similarly, we may define the (global) height of an algebraic number (see Definition 2.1.6). 3 1.2 Further Approximation Results In 1958, Ridout generalized Roth’s theorem to p-adic absolute values on Q and soon after Lang generalized the result to an arbitrary number field. Theorem 1.2.1 (Roth’s Theorem, General Form). Let S be a finite set of absolute values on a number field k. For each v 2 S, let (cid:11)v 2 k and fix an extension of j (cid:1) jv to k. Let " > 0. Then for all (cid:12) 2 k n f(cid:11)vjv 2 Sg ∑ v2S (cid:21)(cid:11)v;v„(cid:12)” (cid:20) „2 + "”h„(cid:12)” + O„1”: We can also view Roth’s Theorem as a statement about k-rational points on the projective line by defining local heights (cid:21)D;v with respect to a divisor D on P1„k” (see Definition 2.2.1). Then we get the following: Theorem 1.2.2 (Roth’s Theorem on P1). Let S be a finite set of places of a number field k. Let i=1 Pi, and " > 0. Then for all P 2 P1„k”nfP1; : : : ; Pqg P1; : : : ; Pq 2 P1„k” be distinct points, D = (cid:21)D;v„P” (cid:20) „2 + "”h„P” + O„1”: mD;S„P” := ∑ ∑q v2S We will study throughout inequalities of this form, where a sum of local heights, with respect to some divisor, is bounded by a constant multiple of a global height (where the constant may depend on the divisor). For curves of genus greater than one, such inequalities are uninteresting (for rational points) due to Faltings’ theorem: Theorem 1.2.3 (Faltings’ Theorem). Let C be a smooth projective curve of genus g (cid:21) 2, defined over a number field k. Then the number of k-rational points of C is finite. Proof. See [8, Part E]. □ This is exactly the statement that, given any finite set S of absolute values on a number field k and distinct points P1; : : : ; Pq 2 C„k”, mD;S„P” (cid:20) O„1” for all P 2 C„k” n fP1; : : : ; Pqg, where D = ∑q i=1 Pi. 4 It remains to ask what happens on curves of genus one, i.e. elliptic curves. The answer for this case is due to Siegel. It requires the notion of a global height associated to a divisor on a curve C (see Definition 3.8). In the case of curves, we will typically take the divisor D to consist of a single point R 2 C„k”. By Theorem 2.2.12, if R; R′ 2 C„k”, then for any " > 0, jhR„P” (cid:0) hR′„P”j (cid:20) "hR„P” + O„1”, so that the choice of the point R will not be significant in our inequalities. ∑q Theorem 1.2.4 (Siegel). Let C be a smooth projective curve of genus one, defined over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pq 2 C„k” be distinct points and let D = i=1 Pi. Let R 2 C„k” and let " > 0. Then for all P 2 C„k” n fP1; : : : ; Pqg mD;S„P” (cid:20) "hR„P” + O„1”: Another way to generalize Roth’s theorem is to consider, instead of k-rational points, algebraic points in C„k” of degree d over k. We denote the support of a divisor D by Supp„D”. ∑q Theorem 1.2.5 (Wirsing’s Theorem). Let S be a finite set of places of a number field k. Let P1; : : : ; Pq 2 P1„k” be distinct points and let D = i=1 Pi. Let " > 0 and let d be a positive integer. Then for all points P 2 P1„k” n Supp„D” satisfying »k„P” : k… (cid:20) d, mD;S„P” (cid:20) „2d + "”h„P” + O„1”: Theorem 1.2.6 (Vojta’s Inequality). Let C be a nonsingular projective curve defined over a number field k with canonical divisor K. Let S be a finite set of places of k. Let P1; : : : ; Pq 2 C„k” be distinct points and let D = i=1 Pi. Let A be an ample divisor on C. Let " > 0. If r is a positive integer then ∑q mD;S„P” + hK„P” (cid:20) da„P” + "hA„P” + O„1” for all points P 2 C„k”n Supp„D” for which »k„P” : k… (cid:20) r, where da is the arithmetic discriminant (a value defined by Vojta using Arakelov theory). Proof. See [17]. □ 5 Vojta’s inequality implies many important results in Diophantine Geometry, including Falt- ing’s theorem and Wirsing’s theorem. Vojta also conjectured that the arithmetic discriminant in the inequality could be replaced by the smaller geometric (logarithmic) discriminant d„P” := log jDk„p”j(cid:157)»k : Q…. The relationship between these discriminants is similar to that between the arithmetic and geometric genus. Vojta’s conjecture implies many other famous conjectures, such as the famous ABC conjecture. We can use Vojta’s inequality to generalize Wirsing’s theorem to curves. Theorem 1.2.7. Let C be a nonsingular projective curve over a number field k. Let S be a finite set of places of k and let R 2 C„k”. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let " > 0 and let d be a positive integer. Then ∑n mD;S„P” (cid:20) „2d + "”hR„P” + O„1” for all P 2 C„k” with »k„P” : k… = d. Proof. Follows from Vojta’s inequality. See [14, Equation (2.0.3)]. □ Song and Tucker show that this result is sharp in the following sense. Theorem 1.2.8. Let C be a nonsingular projective curve over a number field k and let φ : C ! P1 be a nonconstant morphism of degree d. Let S be a finite set of places of k and let R 2 C„k”. Let " > 0. Then there exists a choice of D such that there is an infinite set of points P 2 C„k” with »k„P” : k… (cid:20) d satisfying mD;S„P” (cid:21) „2d (cid:0) "”hR„P” + O„1”: Proof. See [14, Theorem 2.3]. □ However, by replacing 2d with a constant depending on the divisor D, we found we could improve on Theorem 1.2.7 as follows: 6 Theorem 1.2.9 (The Main Theorem). Let C be a nonsingular projective curve over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi 2 Div„C”. Let R 2 C„k” and let " > 0. Then if d 2 f2; 3g, ∑n mD;S„Q” (cid:20) „Nd„D” + "”hR„Q” + O„1” for all Q 2 C„k” with »k„Q” : k… = d, where (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” taken over all finite k-morphisms (cid:27) : C ! P1 of degree d. Nd„D” := max {(cid:12)(cid:12)(cid:12)( ) (cid:12)(cid:12)(cid:12)} \ Supp„D” note that the most that Nd„D” could ever be is 2d, which agrees with Theorem 1.2.7. These diophantine approximation inequalities are closely related to qualitative results for integral points on curves. 1.3 Integral Points on Curves Let k be a number field and let S be a finite set of places of k containing all archimedean places. The ring of S-integers, denoted Ok;S, is defined to be the set of all (cid:12) 2 k such that j (cid:12)jv (cid:20) 1 for all v < S. Theorem 1.3.1 (Siegel’s Theorem, 1929). Let C (cid:26) An be a nonsingular affine curve over a number field k and let S be a finite set of places of k containing the archimedean places. An S-integral point of C is a point whose affine coordinates are all in the ring Ok;S. If C has genus g > 0, or if C has at least three distinct points at infinity, then C has only finitely many S-integral points. Proof. See [1, Theorem 7.3.9]. □ Although this qualitative result preceded the above diophantine approximation inequalities on projective curves, they are related in the following way: 7 Let C (cid:26) An be a nonsingular affine curve over a number field k. Let ˜C be the nonsingular projective completion of C and let D be the very ample divisor on ˜C consisting of the points at infinity of C. Let R 2 ˜C„k” and let " > 0. By Lemma 3.1.7, the set of S-integral points on C corresponds (up to a finite number of points) to the set of points P 2 ˜C„k” n Supp„D” satisfying mD;S„P” = hD„P” + O„1”. If C has genus g = 0 and deg„D” (cid:21) 3, then ˜C is birationally equivalent to P1 and deg„D”hR„P” = hdeg„D”R„P” (cid:20) „1 + "”hD„P” + O„1” by Theorem 2.2.12 = „1 + "”mD;S„P” + O„1” (cid:20) „1 + "”„2 + "”hR„P” + O„1” by Theorem 1.2.2, implying hR„P” (cid:20) O„1”. By Theorem 2.2.11, there are only finitely many such points. If C has genus g = 1, then Theorem 1.2.4 says mD;S„P” (cid:20) "hD„P” + O„1” for all P 2 ˜C n Supp„D”. But if mD;S„P” = hD„P” + O„1”, then this implies hD„P” (cid:20) O„1” and once again Theorem 2.2.11 implies there are only finitely many such points. And of course if C has genus g (cid:21) 2 then Theorem 1.2.3 implies mD;S„P” (cid:20) O„1”. So if mD;S„P” = hD„P” + O„1”, then hD„P” (cid:20) O„1” and there are only finitely such points. Thus these three results together imply Siegel’s theorem. In this way diophantine approximation inequalities for rational points on projective varieties can be thought of as more precise statements of qualitative results for integral points on affine varieties. Similar to how Roth’s theorem implies the genus-0 case of Siegel’s theorem, the main theorem can be used to show the d = 2; 3 cases of the following qualitative result on integral points. Theorem 1.3.2 (Levin). Let C (cid:26) An be a nonsingular affine curve over a number field k and let S be a finite set of places of k containing the archimedean places. Let ˜C be a nonsingular projective completion of C and let „ ˜C n C”„k” = fP1; : : : ; Pqg. Let d be a positive integer. Let Ok;S denote the integral closure of Ok;S in k. Then there exists a finite extension L of k such that the set fP 2 C„OL;S” j »L„P” : L… (cid:20) dg 8 is infinite if and only if there exists a morphism φ : φ„fP1; : : : ; Pqg” (cid:26) f0; 1g. ˜C ! P1, over k with deg„φ” (cid:20) d and Proof. See [10, Theorem 1.9] □ To see how this is implied by the main theorem, let C (cid:26) An be a nonsingular affine curve over a number field k and let S be a finite set of places of k containing the archimedean places. Let ˜C be a nonsingular projective completion of C and let „ ˜C n C”„k” = fP1; : : : ; Pqg. Let d 2 f2; 3g. Let D = P1 + : : : + Pq be a divisor on ˜C and let R 2 ˜C„k”. Then for all P 2 ˜C„OL;S” n fP1; : : : ; Pqg with »k„P” : k… = d, deg„D”hR„P” = hdeg„D”R„P” (cid:20) „1 + "”hD„P” + O„1” by Theorem 2.2.12 = „1 + "”mD;S„P” + O„1” (cid:20) „1 + "”„Nd„D” + "”hR„P” + O„1” by Theorem 1.2.2, {(cid:12)(cid:12)(cid:12)( ) (cid:12)(cid:12)(cid:12)} where Nd„D” := max (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ Supp„D” taken over all finite k-morphisms (cid:27) : C ! P1 of degree d. But if there does not exist a morphism φ : ˜C ! P1, over k with deg„φ” (cid:20) d and φ„fP1; : : : ; Pqg” (cid:26) f0; 1g then N < deg„D”, and so the above inequality implies hR„P” (cid:20) O„1” and by Theorem 2.2.11, there are only finitely many such points P. The main idea behind this result and the main theorem is that we expect almost all of the degree d points on a curve are coming from pulling back rational points on P1 by maps of degree d. The following are two other results that are closely related to this heuristic approach. C is called hyperelliptic (respectively bielliptic) if it admits a map φ : C ! X of degree 2 onto a curve X of genus zero (respectively one). 9 Theorem 1.3.3 (Harris-Silverman). Let C be a nonsingular projective curve over a number field k. Suppose C has genus g (cid:21) 2. Assume that C is neither hyperelliptic nor bielliptic. Then the set of points P 2 C„k” with »k„P” : k… (cid:20) 2 is finite. Proof. See [6, Corollary 3] Theorem 1.3.4 (Corvaja-Zannier). Let C (cid:26) An be a nonsingular affine curve over a number field k and let S be a finite set of places of k containing the archimedean places. Let ˜C be a nonsingular projective completion of C and let „ ˜C n C”„k” = fP1; : : : ; Pqg. (a) If q (cid:21) 5, then C contains only finitely many quadratic (over k) S-integral points. (b) If q (cid:21) 4, then there exists a finite set of rational maps φ : □ all but finitely many of the quadratic S-integral points on C are sent to P1 mentioned maps. ˜C ! P1 of degree 2 such that k by some of the Proof. See [2, Corollary 1] □ 10 CHAPTER 2 REVIEW OF HEIGHTS 2.1 Heights on Projective Spaces The height of a rational number p(cid:157)q is defined to be h„p(cid:157)q” = log max fjpj; jqjg. Similarly the height of a point P = „x0; : : : ; xn” 2 Pn„Q”, with coordinates chosen so that x0; : : : ; xn 2 Z and gcd„x0; : : : ; xn” = 1, is defined to be h„P” = log max fjx0j; : : : ; jxnjg. One of the key features of heights is that, given a bound, the number of points of bounded height is finite, that is the set { } P 2 Pn„Q” j h„P” (cid:20) B is finite for any B > 0. In order to generalize the concept of heights to number fields other than Q, we recall some properties of absolute values on fields. An absolute value on a field k is a map j (cid:1) j : k ! R satisfying the following. • j(cid:11)j (cid:21) 0 for all (cid:11) 2 k. • j(cid:11)j = 0 if and only if (cid:11) = 0. • j(cid:11)(cid:12)j = j(cid:11)jj (cid:12)j for all (cid:11); (cid:12) 2 k. • j(cid:11) + (cid:12)j (cid:20) j(cid:11)j + j (cid:12)j for all (cid:11); (cid:12) 2 k. We say two absolute values on a field k are equivalent if they induce the same topology on k. It can be shown that Proposition 2.1.1. Two absolute values j (cid:1) j1; j (cid:1) j2 on a field k are equivalent if and only if there is a positive real number s such that for all x 2 k. jxj1 = jxjs 2 11 Proof. See [11, Proposition 3.3]. □ We denote the set of equivalence classes of non-trivial absolute values on a field k by Mk and call its elements places. If k′ is an extension of k and v is a place of k, we say a place w of k′ extends v, denoted wjv, if the restriction to k of any representative of w is a representative of v. Definition 2.1.2. The completion of k with respect to the place v is an extension field kv, with a place w such that: (a) wjv (b) The topology of kv induced by w is complete. (c) k is a dense subset of kv in the above topology. The completion exists and is unique up to isometric isomorphisms. By abuse of notation, we shall denote the unique place w also by v. The standard absolute values on Q are the (usual) archimedean absolute value j (cid:1) j1 = j (cid:1) j and (cid:0)ordp„(cid:1)” for each prime p. By Ostrowski’s Theorem, this is a the p-adic absolute values j (cid:1) jp = p complete list of all absolute values on Q up to equivalence. Definition 2.1.3. Let p be a place on Q. We consider a number field k and a place v of k extending p. For any x 2 k, we define jxjv = jNkv(cid:157)Qp where Nkv(cid:157)Qp is the norm from kv to Qp. „x”j1(cid:157)»k:Q… p note that j (cid:1) jv is an extension to k of j (cid:1) j»kv:Qp…(cid:157)»k:Q… p 2.1 is equivalent to j (cid:1) jp. That is vjp. on Q, an absolute value that by Proposition These normalized absolute values satisfy the following. 12 Lemma 2.1.4. Let k be a number field with a place v 2 Mk. We consider a finite-dimensional extension field k′ of k. For any x 2 k(cid:3), ∏ Proof. See [1, Lemma 1.3.7]. Proposition 2.1.5 (The Product Formula). Let k be a number field. For any x 2 k(cid:3), □ jxjw = jxjv: jxjv = 1: w2Mk′ wjv ∏ v2Mk ∑ v2Mk ∑ v2Mk Proof. See [1, Proposition 1.4.4]. Definition 2.1.6. The (absolute) height of a point P 2 Pn„Q” with homogeneous coordinates „x0; : : : ; xn” is defined to be □ h„P” = log jxijv max 0(cid:20)i(cid:20)n where k is some number field containing x0; : : : ; xn. We also define the height of an algebraic number (cid:11) 2 Q to be the height of the point „(cid:11); 1” 2 P1„Q” h„(cid:11)” := h„„(cid:11); 1”” = log max fj(cid:11)jv; 1g: Proposition 2.1.7. The height, so defined, is independent of the choice of homogeneous coordinates for P and the choice of number field k containing them. Proof. See [1, Lemmas 1.5.2 and 1.5.3]. □ One reason for studying heights is because they satisfy the following. Theorem 2.1.8 (Northcott’s Theorem). For any numbers B; d (cid:21) 0, the set fP 2 Pn„Q” j h„P” (cid:20) B and »Q„P” : Q… (cid:20) dg 13 is finite. In particular, for any fixed number field k, the set j h„P” (cid:20) Bg fP 2 Pn k is finite. Proof. See [1, Theorem 1.6.8]. □ Another important property of heights is preservation under Galois action. Proposition 2.1.9. Let (cid:27) 2 Gal„Q(cid:157)Q”. Then h„P” = h„(cid:27)„P”” for all P 2 Pn„Q”, where ( ) (cid:27) „xi”n i=1 := „(cid:27)„xi””n i=1 Proof. See [1, Proposition 1.5.17]. □ 2.2 Heights on Projective Varieties Let X be a projective variety over a number field k. We denote the set of divisors on X by Div„X”. We consider a divisor D on X with associated line bundle O„D” and rational section sD. There are base-point-free line bundles L, M on X such that O„D” (cid:27) L (cid:10) M(cid:0)1. Now choose generating global sections s0; : : : ; sm of L and t0; : : : ; tn of M, and call the data D = „sD; L; s; M; t” a presentation of the divisor D. We denote the support of the divisor D by Supp„D”. Definition 2.2.1. Let v 2 Mk. For P 2 X„k” n Supp„D”, we define the local height of P relative to the presentation D at the place v to be (cid:12)(cid:12)(cid:12)(cid:12) sk tlsD (cid:12)(cid:12)(cid:12)(cid:12) „P” : v (cid:21)D;v„P” := log max k min l Definition 2.2.2. If D1 and D2 are divisors with presentations Di = „sDi; Li; si; Mi; ti”, then we define the presentation D1 + D2 of the divisor D1 + D2 to be D1 + D2 := „sD1sD2; L1 (cid:10) L2; s1s2; M1 (cid:10) M2; t1t2”: 14 Similarly, if a divisor D has presentation D = „sD; L; s; M; t”, we define the presentation (cid:0)D of the divisor (cid:0)D to be (cid:0)D := „s (cid:0)1 D ; M; t; L; s”: Furthermore, if (cid:25) : Y ! X is a morphism of projective varieties such that (cid:25)„Y” is not contained in Supp„D”, then we define the presentation (cid:25)(cid:3)D of the divisor (cid:25)(cid:3)D on Y to be (cid:25)(cid:3)D := „(cid:25)(cid:3) sD; (cid:25)(cid:3) L; (cid:25)(cid:3)s; (cid:25)(cid:3) M; (cid:25)(cid:3)t”: With these definitions, it is immediate that the corresponding local heights satisfy (cid:21)D1+D2;v = (cid:21)D1;v + (cid:21)D2;v; (cid:21)(cid:0)D;v = (cid:0)(cid:21)D;v; and (cid:21)(cid:25)(cid:3)D;v = (cid:21)D;v ◦ (cid:25) for all v 2 Mk. Definition 2.2.3. An Mk-constant is a map (cid:13) : Mk ! R with the property that (cid:13)v = 0 for all but finitely many v 2 Mk. Theorem 2.2.4. Let D and D′ be two presentations of the divisor D. Then there is an Mk-constant (cid:13) such that j(cid:21)D;v (cid:0) (cid:21)D′;v j (cid:20) (cid:13)v: Proof. See [1, Theorem 2.2.11]. □ For this reason, whenever an inequality holds only up to the addition of a bounded function, we will omit the presentation D and, by abuse of notation, denote a local height relative to D at a place v by (cid:21)D;v. Proposition 2.2.5. Let D be an effective divisor on X. Then there is a presentation D of D such that for any P < Supp„D” and for any field extension k (cid:26) k′ (cid:26) k such that P 2 X„k′” and any place v 2 Mk′, it holds that (cid:21)D;v„P” (cid:21) 0. Proof. See [1, Proposition 2.3.9]. □ 15 Example 2.2.6. Let (cid:11) 2 k. The point „(cid:11); 1” in P1 A = „x0 (cid:0) (cid:11)x1; O For (cid:12) 2 k n f(cid:11)g and v 2 Mk the corresponding local height is (cid:21)(cid:11);v„(cid:12)” := (cid:21)A;v„„(cid:12); 1”” = log max P1„1”; „x1; x0 (cid:0) (cid:11)x1”; O k has the presentation P1; 1”: } {(cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12) (cid:0) (cid:11) ; 1 v which is the local height we described in Definition 1.1.7. Example 2.2.7. The hyperplane fx0 = 0g in Pn k has the presentation D = „x0; OPn„1”; „x0; x1; : : : ; xn”; OPn; 1”: k with x0„P” , 0 and v 2 Mk the corresponding local height is For P 2 Pn {(cid:12)(cid:12)(cid:12)(cid:12) xi x0 } (cid:12)(cid:12)(cid:12)(cid:12) v (cid:21)D;v„P” = log max „P” and the product formula becomes h„P” = (cid:21)D;v„P”: v2Mk For this reason we make the following definition. i ∑ ∑ v2Mk′ Definition 2.2.8. Let D = „sD; L; s; M; t” be a presentation of a divisor D on X. For P 2 X„k” there are si and tj such that si„P” , 0, tj„P” , 0 by definition of base point free line bundles. Thus we can find a non-zero rational section s of O„D” such that P is not contained in the support of the divisor D„s”. Then D„s” = „s; L; s; M; t” is a presentation of D„s”. If k′ is a finite extension k (cid:26) k′ (cid:26) k such that P 2 X„k′”, then we define the global height of P relative to D by hD„P” = (cid:21)D„s”;v „P”: Proposition 2.2.9. The global height hD is independent of the choices of k′ and of the section s. Proof. See [1, Proposition 2.3.4]. □ 16 It is immediately clear by the definition and the above discussion that, given presentations D1, D2, D of divisors D1, D2, D respectively and a morphism of projective varieties (cid:25) : X ! Y, the corresponding global heights also satisfy hD1+D2 = hD1 + hD2 ; h(cid:0)D = (cid:0)hD; and h(cid:25)(cid:3)D = hD ◦ (cid:25): Furthermore, by Proposition 3.2, if D 2 Div„X” is effective, then there exists a presentation D of D such that hD„P” (cid:21) 0 for all P 2 X„k”. For two divisors D1 and D2, we write D1 (cid:24) D2 if D1 is linearly equivalent to D2, i.e. difference D2 (cid:0) D1 is a principal divisor. Theorem 2.2.10. Let D1, D2 be presentations of divisors D1, D2 2 Div„X” with D1 (cid:24) D2. Then there is a constant (cid:13) > 0 such that the jhD2 (cid:0) hD1 j (cid:20) (cid:13): Proof. See [1, Theorem 2.3.6]. □ Similarly to local heights, whenever an inequality holds only up to the addition of a bounded function, we will omit the presentation D and, by abuse of notation, denote a global height relative to D by hD. Furthermore, in the special case where the divisor D is an effective divisor of degree 1 on a curve and Supp„D” = fRg, we define hR := hD. Global heights also satisfy Northcott’s theorem. Theorem 2.2.11 (Northcott’s Theorem). Let D be an ample divisor on X and let D be a presentation of D. Then the set fP 2 X„k” j hD„P” (cid:20) B; »k„P” : k… (cid:20) dg is finite for any constants B; d 2 R. Proof. See [1, Theorem 2.4.9]. □ Recall that two divisors D1; D2 2 Div„X” are algebraically equivalent if there exists a connected algebraic set T, two points t1; t2 2 T„k” and a line bundle L on X (cid:2) T such that 17 O„D1” (cid:27) LjX(cid:2)ft1g and O„D2” (cid:27) LjX(cid:2)ft2g. In particular if X is a curve, then any two points on X are algebraically equivalent. Theorem 2.2.12. Let A; B 2 Div„X” be algebraically equivalent divisors with A ample. Let " > 0 and d a positive integer. Then for all P 2 X„k” with »k„P” : k… (cid:20) d we have jhA„P” (cid:0) hB„P”j (cid:20) "hA„P” + O„1”: Proof. Follows from [8, Theorem B.5.9] and Theorem 2.2.11. □ Given a presentation D = „sD; L; s; M; t” of a divisor D 2 Div„X”, we have the morphisms and φ : X ! Pm k ; P 7! „s0„P”; s1„P”; : : : ; sm„P”” : X ! Pn k ; P 7! „t0„P”; t1„P”; : : : ; tn„P””: By the product formula we see that hD„P” = h„φ„P”” (cid:0) h„ „P””. Therefore, by Proposition 2.1.9, for any (cid:27) 2 Gal„k(cid:157)k” we have hD„(cid:27)„P”” = hD„P”. 2.3 Heights on Closed Subschemes We identify a closed subscheme X of a nonsingular projective variety V with its ideal sheaf IX. Then for closed subschemes X and Y we define the following: IX+Y = IXIY IX\Y = IX + IY IX[Y = IX \ IY : We say X (cid:26) Y if IY (cid:26) IX. Let : V ! W be a morphism of projective varieties. Then for any closed subscheme X (cid:26) V and Z (cid:26) W we define the scheme-theoretic image f „X” by If „X” = Ker„OW ! f(cid:3)OX” and the scheme theoretic inverse image f (cid:3)Z by If (cid:3)Z = f (cid:0)1IZ (cid:1) OV. f 18 Lemma 2.3.1. Let f : V ! W be a morphism of projective varieties and let Z (cid:26) W be a closed subscheme. The scheme theoretic inverse image f (cid:3)Z is the fibered product f (cid:3)Z Z of Z and V over W, that is f (cid:3)Z = Z (cid:2)W V. V f W Proof. See Tag 01JU on The Stacks Project [15, Lemma 01JU]. Lemma 2.3.2. Let f : V ! W be a morphism of projective varieties. Then for any Zariski-closed subschemes X (cid:26) V and Z (cid:26) W, (a) X (cid:26) f (cid:3)„ f „X”” (b) f „ f (cid:3)Z” (cid:26) Z □ Proof. (a) By definition of f „X” we get a commutative diagram of sheaves, f(cid:3)OX OW(cid:157)If „X” f(cid:3)OV f # OW to which the corresponding commutative diagram of schemes is as follows. X f „X” V f W By the universal property of the fibered product, there exists a unique morphism X ! f „X” (cid:2)W V such that the following diagram commutes. f „X” (cid:2)W V X φ V 19 By Lemma 2.3.1, immersion. Thus we also get the commutative diagram of structure sheaves, f „X” (cid:2)W V is the inverse image f (cid:3)„ f „X”” and the map φ is a closed OV(cid:157)IX OV(cid:157)If (cid:3)„ f „X”” φ# OV which implies that If (cid:3)„ f „X”” (cid:26) IX, that is X (cid:26) f (cid:3)„ f „X””. (b) By Lemma 2.3.1, we know that f (cid:3)Z = Z (cid:2)W V. Thus we get the corresponding commutative diagram of sheaves, f(cid:3)O f (cid:3)Z f(cid:3)OV f # OW OW(cid:157)IZ from which it follows that IZ (cid:26) Ker„OW ! f(cid:3)O f (cid:3)Z” = If „ f (cid:3)Z”, that is f „ f (cid:3)Z” (cid:26) Z. ∩ □ Lemma 2.3.3. Let X (cid:26) V be a closed subscheme. There exist effective divisors D1; : : : ; Dr such that X = Di. Proof. See [13, Lemma 2.2]. □ Given such D1; : : : ; Dr, we define the local heights corresponding to the closed subscheme X to be (cid:21)X;v := minif(cid:21)Di;vg, for all places v 2 Mk. note that on Supp„Di” we may think of (cid:21)Di;v as having value 1 so that (cid:21)X;v is defined (up to a bounded function) outside Supp„Di” = Supp„X”. Theorem 2.3.4. Let V be a projective variety defined over a number field k and let v 2 Mk. Then up to a bounded function we have the following: (a) For all closed subschemes X (cid:26) V and places v 2 Mk, the local height (cid:21)X;v is well-defined, ∩ up to a bounded function, independent of the choice of effective divisors D1; : : : ; Dr. (b) For all closed subschemes X; Y (cid:26) V, (cid:21)X\Y;v = minf(cid:21)X;v; (cid:21)Y;vg: 20 (c) For all closed subschemes X; Y (cid:26) V, (cid:21)X+Y;v = (cid:21)X;v + (cid:21)Y;v: (d) If closed subschemes X; Y (cid:26) V satisfy X (cid:26) Y, then (cid:21)X;v (cid:20) (cid:21)Y;v: (e) For all closed subschemes X; Y (cid:26) V, maxf(cid:21)X;v; (cid:21)Y;vg (cid:20) (cid:21)X[Y;v (cid:20) (cid:21)X;v + (cid:21)Y;v: (f) If closed subschemes X; Y (cid:26) V satisfy Supp„X” (cid:26) Supp„Y”, then there exists a constant c (cid:21) 0 such that (cid:21)X;v (cid:20) c(cid:21)Y;v: (g) Let φ : W ! V be a morphism of varieties, and let X (cid:26) V be a closed subscheme. Then (cid:21)W;φ(cid:3)X;v = (cid:21)V;X;v ◦ φ: Proof. See [13, Theorem 2.1]. □ 21 CHAPTER 3 PREPARATION 3.1 Useful Results ∩r Given a nonsingular variety V defined over a number field k and a closed subscheme X, and i=1 Di, we define the local heights corresponding to the closed given D1; : : : ; Dr such that X = subscheme X to be for all points P 2 V n X and all places v 2 Mk. These are well-defined by Theorem 2.3.4. For a set of places S (cid:26) Mk we similarly define (cid:21)X;v„P” := min 1(cid:20)i(cid:20)r mX;S„P” := f(cid:21)Di;v„P”g ∑ (cid:21)X;v„P”: v2S Faltings generalized Siegel’s approximation result on elliptic curves to abelian varieties as follows. Theorem 3.1.1. Let V be an abelian variety over a number field k. Let S be a finite set of places of k. Let X (cid:26) V be a closed subscheme and A an ample divisor on V. Let " > 0. Then for any P 2 V„k” n Supp„X” mX;S„P” (cid:20) "hA„P” + O„1”: Proof. See [5, Theorem 2]. □ Roth’s theorem was also generalized to higher dimensional projective spaces. Theorem 3.1.2 (Schmidt’s Subspace Theorem). Let k be a number field. Let S be a finite set of places of k. For each v 2 S let fL0v; : : : ; Lnvg be a linearly independent set of linear forms in the variables x0; : : : ; xn, with coefficients in k and let Hiv be the hyperplane in Pn associated with the linear form Liv. Let " > 0. Then there exists a finite set H of hyperplanes of Pn k such that for all P 2 Pn k n∪ H2H H ∑ n∑ (cid:21)Hiv v2S i=0 ;v„P” (cid:20) „Nd„D” + 1 + "”h„P”: 22 Proof. See [1, Chapter 7]. □ We say a set of m +1 hyperplanes in Pn is in j-subgeneral position if n (cid:20) j (cid:20) m and any subset of size j + 1 of the hyperplanes will have empty intersection. In this case we get the following result of Ru and Wong. Theorem 3.1.3. Let k be a number field. Let S be a finite set of places of k. Let fH0; : : : ; Hmg be a set of hyperplanes in Pn in d-subgeneral position. Let " > 0. Then there exists a finite set H of hyperplanes of Pn n∪ k such that for all P 2 Pn k H2H H i=0 Proof. See [12, Theorem 3.5]. mHi;S„P” (cid:20) „2d (cid:0) Nd„D” + 1 + "”h„P”: □ m∑ Generalizations of the subspace theorem to projective varieties have been given, independently, by Corvaja and Zannier [3, Theorem 3] and by Evertse and Ferretti, whose version we state. Theorem 3.1.4 (Evertse-Ferretti). Let X be a projective subvariety of PN of dimension n (cid:21) 1 defined over a number field k. Let S be a finite set of places of k. For all v 2 S, let H0;v; : : : ; Hn;v 2 PN be hypersurfaces over k such that X \ H0;v \ (cid:1) (cid:1) (cid:1) \ Hn;v = ∅ Let " > 0. Then there exists a proper Zariski-closed subscheme Z (cid:26) X such that, for all points P 2 X„k” n Z, ∑ n∑ ;v„P” (cid:21)Hi;v degHi;v < „Nd„D” + 1 + "”h„P”: □ v2S i=0 Proof. See [4, Theorem 1.1] This may be reformulated in terms of divisors as follows. Theorem 3.1.5 (Evertse-Ferretti, reformulated). Let X be a projective variety of dimension n defined over a number field k. Let S be a finite set of places of k. For each v 2 S let D0;v; : : : ; Dn;v be effective divisors on X, defined over k, in general position. Suppose there exists an ample divisor 23 A and positive integers di;v such that Di;v is linearly equivalent to di;v A for all i and for all v 2 S. Let " > 0. Then there exists a proper Zariski-closed subscheme Z (cid:26) X such that for all points P 2 X„k” n Z, ∑ n∑ ;v„P” (cid:21)Di;v di;v (cid:20) „Nd„D” + 1 + "”hA„P” + O„1”: □ i=0 Proof. See [9, Theorem 3.1] v2S Levin proved that the linear equivalence in this reformulation may be replaced by numerical equivalence. Theorem 3.1.6. Let X be a projective variety of dimension n defined over a number field k. Let S be a finite set of places of k. For each v 2 S let D0;v; : : : ; Dn;v be effective divisors on X, defined over k, in general position. Suppose there exists an ample divisor A and positive integers di;v such that Di;v is numerically equivalent to di;v A for all i and for all v 2 S. Let " > 0. Then there exists a proper Zariski-closed subscheme Z (cid:26) X such that for all points P 2 X„k” n Z, i=0 Proof. See [9, Theorem 3.2]. v2S ;v„P” (cid:21)Di;v di;v (cid:20) „Nd„D” + 1 + "”hA„P” + O„1”: □ ∑ n∑ Let k be a number field and let S be a finite set of places of k containing all archimedean places. The ring of S-integers, denoted OS, is defined to be the set of all x 2 k such that jxjv (cid:20) 1 for all v < S. If S consists of only the archimedean places, we call these the ring of integers of k, denoted Ok. If V is a projective variety over k and D a very ample effective divisor on V with x = „x0 = 1; x1; : : : ; xn) a basis for L„D”, then the map P 7! „x1„P”; x2„P”; : : : ; xn„P”” defines an embedding of V n Supp„D” into An. We say a point P 2 V„k” n Supp„D” is „D; x; S”-integral if xi„P” 2 OS for 0 (cid:20) i (cid:20) n. note that for any point P 2 V„k” n Supp„D”, we may choose a b 2 Ok which clears the denominators of xi„P”, 1 (cid:20) i (cid:20) n, so that x′ = „x0 = 1; bx1; : : : ; bxn” is a basis of O„D” such that P is „D; x′; S”-integral. In order to find a more intrinsic definition, we need to look at sets of points. 24 Lemma 3.1.7. Let D be a very ample effective divisor on V. Let R be a subset of V„k” n Supp„D”. Then the following are equivalent. (a) There exists a basis x = „x0 = 1; x1; : : : ; xn” of O„D” such that R is a set of „D; x; S”-integral points. (b) There exists a presentation D of D and an Mk-constant (cid:13) such that for all P 2 R and all v 2 Mk n S, (cid:21)D;v„P” (cid:20) (cid:13)v. Proof. See [16, Lemma 1.4.1]. □ Going back to Roth’s theorem, if the points Pi are not distinct we can still say the following. Theorem 3.1.8 (Roth’s Theorem with Multiplicities). Let S be a finite set of places of a number field k. Let P1; : : : ; Pn 2 P1 k be distinct points and let c1; c2; : : : cn be positive integers with c1 (cid:21) c2 (cid:21) (cid:1) (cid:1) (cid:1) (cid:21) cn. Let D = i=1 ciPi and let " > 0. Then for P 2 P1 mD;S„P” < „c1 + c2 + "”h„P” + O„1”: ∑n k Proof. See [10, Theorem 2.1]. □ 3.2 Sharpness of the Main Result Let C be a nonsingular projective curve defined over a number field k. Fix a point R 2 C„k”. For P 2 C„k” , let k„P” be the field of definition of P (the field extension of k generated by its local coordinates). Theorem 3.2.1. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let " > 0. Denote the set of archimedean places of k by Sk;1. Then there exists a finite extension k′ of k and ∑n 25 taken over all finite k-morphisms (cid:27) : an infinite collection of points P 2 C„k” with »k′„P” : k′… (cid:20) d satisfying {(cid:12)(cid:12)(cid:12)( mD;Sk′;1„P” (cid:21) „N (cid:0) "”hR„P” + O„1” ) (cid:12)(cid:12)(cid:12)} (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” where Nd„D” := max C ! P1 of degree d. Proof. By finitely extending k to k′, we may assume O(cid:3) \ Supp„D” k′ is an infinite set. Since D is effective, we know mD;Sk′;1 (cid:21) O„1”, so we may assume N (cid:21) 1. For each k- morphism (cid:27) : C ! P1 of degree d, let (cid:27)„Supp„D”” = fT1; T2; : : : ; Trg and define ni = ni„(cid:27)” := j(cid:27)(cid:0)1„Ti” \ Supp„D”j for i = 1; 2; : : : ; r. Reorder indices so that n1 (cid:21) n2 (cid:21) : : : (cid:21) nr. Then choose a morphism (cid:27) with T1 = 0 and T2 = 1 such that n1„(cid:27)” + n2„(cid:27)” = N. We consider the infinite set of points P 2 C„k” satisfying (cid:27)„P” = „(cid:11); 1” for some (cid:11) 2 O(cid:3) k′. For each v 2 Mk′ define the local height (cid:21)0+1;v on P1 by ∑m ∑ For any effective divisor E on C and v 2 Mk′, we have (cid:21)E;v„P” (cid:21) 0. Thus we have mE;Sk′;1„P” = i=1 Qi then mQi;Sk′;1„P” (cid:20) v2Sk′;1 (cid:21)E;v„P” (cid:20)∑ (cid:21)E;v„P” = hE„P”. So if (cid:27)(cid:3)„0 + 1” = v2M′ k 26 (cid:21)0+1;v„„a; b”” := log max 0(cid:20)i(cid:20)1 { Since (cid:21)0+1;v„(cid:27)„P”” = log max v 2 Mk′ n Sk′;1 by definition of O(cid:3) (cid:11) j(cid:11)jv; k′. Thus m(cid:27)(cid:3)„0+1”;Sk′;1„P” = = log max v x0x1 for all v 2 Mk′, we see that (cid:21)0+1;v„(cid:27)„P”” = 0 for b a v v {(cid:12)(cid:12)(cid:12)a (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) b (cid:12)(cid:12)(cid:12)(cid:12) ; } : }(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)„a;b” i {(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) x2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) 1 (cid:12)(cid:12)(cid:12) } ∑ ∑ ∑ = v = v2Sk′;1 v2Sk′;1 (cid:21)(cid:27)(cid:3)„0+1”;v „P” + O„1” (cid:21)0+1;v„(cid:27)„P”” + O„1” (cid:21)0+1;v„(cid:27)„P”” + O„1” v2Mk′ = h0+1„(cid:27)„P”” + O„1” = h(cid:27)(cid:3)„0+1”„P” + O„1” „P” and hQi m∑ i=1 mQi;Sk′;1„P” = m(cid:27)(cid:3)„0+1”;Sk′;1„P” + O„1” m∑ = h(cid:27)(cid:3)„0+1”„P” + O„1” „P” + O„1”: = hQi i=1 Therefore in fact mQ;Sk′;1„P” = hQ„P” for all Q 2 Supp„(cid:27)(cid:3)„0 + 1””. So ∑ mD;Sk′;1„P” = (cid:21) Q2Supp„D” ∑ mQ;Sk′;1„P” + O„1” ∑ Q2Supp„D”\Supp„(cid:27)(cid:3)„0+1”” mQ;Sk′;1„P” + O„1” hQ„P” + O„1” = Q2Supp„D”\Supp„(cid:27)(cid:3)„0+1”” (cid:21) „n1„(cid:27)” + n2„(cid:27)” (cid:0) "”hR„P” + O„1” by Theorem 2.2.12 = „N (cid:0) "”hR„P” + O„1”: □ 3.3 Facts on Trigonal Maps From this point forward the symbol (cid:24) will refer to linear equivalence of divisors on a variety. We will make frequent use of the following theorem about divisors on curves. For a divisor D on a nonsingular projective curve C, let l„D” := dimk H0„C; O„D””. Theorem 3.3.1 (Riemann-Roch Theorem). Let D be a divisor on a curve C of genus g. Let K be a canonical divisor on C. Then l„D” (cid:0) l„K (cid:0) D” = deg D + 1 (cid:0) g: Proof. See [7, Theorem IV.1.3]. □ 27 Lemma 3.3.2. Let C be a curve of genus g (cid:21) 3 defined over k. Then C cannot be both hyperelliptic and trigonal. Proof. Suppose C is both hyperelliptic and trigonal. Then there exist finite morphisms φ : C ! P1 and : C ! P1 of degrees 2 and 3 respectively. By the universal property of fibered products there exists a unique morphism φ : C ! P1 (cid:2) P1 such that the following diagram commutes. C φ φ P1 P1 (cid:2) P1 P1 By the commutative diagram we see the degree of φ divides the degrees of φ and , that is 2 and 3, so the degree of φ is 1, meaning φ is a birational morphism. Thus φ„C” is an irreducible curve on P1 (cid:2) P1 of type „2; 3”. It follows that the genus of C is bounded above by the formula „d1 (cid:0) 1”„d2 (cid:0) 2” for the genus of a nonsingular curve of type „d1; d2” on P1 (cid:2) P1. That is to say g (cid:20) „2 (cid:0) 1”„3 (cid:0) 1” = 2. □ Lemma 3.3.3. Let C be a curve of genus g (cid:21) 5 defined over k. Then C can be trigonal in at most one way (up to k-automorphism). Proof. Suppose C is trigonal in more than one way. Then there exist two finite morphisms φ : C ! P1 and : C ! P1 of degree 3, distinct even up to an automorphism of P1. By the universal property of fibered products there exists a unique morphism φ : C ! P1 (cid:2) P1 such that the following diagram commutes. C φ φ P1 (cid:25)1 P1 (cid:2) P1 (cid:25)2 P1 By the commutative diagram we see the degree of φ divides the degrees of φ and , so the degree of φ is either 1 or 3. However if the degree of φ is 3 then the projections (cid:25)1 and (cid:25)2 each have degree 1, so (cid:25)1jφ„C” and (cid:25)2jφ„C” are birational maps. Thus (cid:25)1jφ„C” ◦ (cid:25)2j(cid:0)1 φ„C” is a birational map 28 from P1 to P1, which necessarily extends to a k-automorphism. But then φ = (cid:25)1jφ„C” ◦ (cid:25)2j(cid:0)1 a contradiction since φ and were distinct trigonal maps up to k-automorphism of P1. φ„C” ◦ , Thus we conclude the degree of φ is 1, meaning φ is a birational morphism. Then φ„C” is an irreducible curve on P1 (cid:2) P1 of type „3; 3”. It follows that the genus of C is bounded above by the formula „d1 (cid:0) 1”„d2 (cid:0) 2” for the genus of a nonsingular curve of type „d1; d2” on P1 (cid:2) P1. That is to say g (cid:20) „3 (cid:0) 1”„3 (cid:0) 1” = 4. □ Lemma 3.3.4. Let C be a curve of genus g = 3 defined over k. Then C has only finitely many distinct trigonal k-morphisms up to k-automorphism of P1. Proof. Let φ : C ! P1 be a finite k-morphism of degree 3. Let D be an effective divisor defined over k in the corresponding base-point free linear system. Then deg„D” = 3, l„D” = 2. By Riemann-Roch, l„K (cid:0) D” = 1, which is to say there exists a unique point P 2 C„k” such that K (cid:0) D (cid:24) P. Since K (cid:0) D is defined over k it follows that P is k-rational. If : C ! P1 is another trigonal morphism and E a corresponding divisor such that K (cid:0) E (cid:24) P then D (cid:24) E, implying φ and correspond to the same linear system. Thus φ and are equivalent up to automorphism of P1. By Falting’s Theorem, there exists at most finitely many k-rational points on C. Since each trigonal k-morphism has a corresponding k-rational point and no two distinct trigonal k-morphisms correspond to the same point, it follows that there are finitely many trigonal k-morphisms up to automorphism of P1. □ Lemma 3.3.5. Let C be a curve of genus g = 4 defined over k. Then C can be trigonal in at most two distinct ways up to a k-automorphism of P1. Proof. Suppose C is trigonal in more than one way. Then there exist two finite morphisms φ : C ! P1 and : C ! P1 of degree 3, distinct even up to a k-automorphism of P1. Let D and E be corresponding effective divisors of the respective base-point free linear systems. Let „1; f ” and „1; g” be local equations for φ and . Because D + E + „1”, D + E + „ f ”, D + E + „g”, and 29 D + E +„ f (cid:1) g” are all effective divisors, we can think of the functions 1, f , g, and f (cid:1) g as elements of L„D + E”. Assume 1, f , g, and f (cid:1) g are k-linearly dependent. Consider the map φ : C ! P1 (cid:2) P1 given by „φ; ” = „1; f ” (cid:2) „1; g”. Composing with the Segre embedding gives a map C ! P3 with coordinates „1; f ; g; f (cid:1) g”. Since these are k-linearly dependent, the image lies in a hyperplane in P3. A hyperplane in P3 cuts out a curve of type „1; 1” in P1 (cid:2) P1. Thus the image of φ lies in a curve H of type „1; 1”. Since neither φ nor is constant, H must be a copy of P1 embedded in P1 (cid:2) P1. Thus we get an induced morphism φ′ : C ! H satisfying the following commutative diagram. C φ′ φ (cid:25)1 H P1 P1 (cid:2) P1 (cid:25)2 P1 Since H has type „1; 1” it follows that the projections (cid:25)1jH and (cid:25)2jH restricted to H are morphisms of degree 1 from P1 to itself, that is automorphisms of P1. Thus φ = (cid:25)1jH ◦ (cid:25)2j(cid:0)1 ◦ , a contradiction since φ and were distinct trigonal maps up to k-automorphism of P1. Thus f , g, and f (cid:1) g must in fact be k-linearly independent. Therefore we see the four functions 1, l„D + E” = dimk L„D + E” (cid:21) 4. By Riemann-Roch we have l„K (cid:0) D (cid:0) E” = l„D + E” (cid:0) 3 (cid:21) 1. Since deg„K (cid:0) D (cid:0) E” = 0 we H conclude K (cid:0) D (cid:0) E (cid:24) 0, that is D + E is a canonical divisor. Let (cid:24) : C ! P1 be a finite morphism of degree 3 with corresponding effective divisor F. If (cid:24) is not equivalent to φ (up to a k-automorphism of P1) then the same argument shows that D + F is a canonical divisor. Thus E (cid:24) F, implying and (cid:24) have the same corresponding linear system, □ which is to say (cid:24) is equivalent to . 30 CHAPTER 4 PROOF OF THE MAIN THEOREM 4.1 Algebraic Points on Curves A divisor B on a projective variety X of dimension n is called big if there exists an a > 0 and j0 > 0 such that H0„V; OV„jB”” (cid:21) a jn for all j > j0. Lemma 4.1.1 (Kodaira’s Lemma). Let B be a big divisor and D an arbitrary divisor on a nonsin- gular projective variety V. Then for m ≫ 0, H0„V; OV„mB (cid:0) D”” , 0: Corollary 4.1.2. Let A be an ample divisor and D an arbitrary divisor on a nonsingular projective variety V. Then for m ≫ 0, hD„P” (cid:20) mhA„P” + O„1” for all P 2 V„k”. Proof. Since ample divisors are big, by Kodaira’s Lemma we have H0„V; OV„mA (cid:0) D”” , 0 for m ≫ 0. So there exists an effective divisor E such that mA (cid:24) D + E. Choose an n 2 Z so that E + nA is very ample. „m + n”hA„P” = h„m+n”A„P” + O„1” = hD+E+nA„P” + O„1” = hD„P” + hE+nA„P” + O„1” (cid:21) hD„P” + O„1”: □ Let C be a nonsingular projective curve of genus g (cid:21) 1 defined over a number field k. Fix a point R 2 C„k”. For Q 2 fP 2 C : »k„P” : k… = dg let ˜Q1; : : : ; ˜Qd be the Galois conjugates of Q, in some order, and define ¯Q = „ ˜Q1; : : : ; ˜Qd” 2 Cd and φ : Cd ! Symd„C” by 31 „P1; : : : ; Pd” 7! P1 + : : : + Pd. We identify Symd„C” with the set of effective divisors of degree d on C. Let (cid:25)1 : Cd ! C be the first projection and let (cid:22) : Symd„C” ! Jac„C” be the map given by P1 + : : : + Pd 7! OC„P1 + : : : + Pd (cid:0) dR”. We’ll need the following Lemma, as shown in [9]. Lemma 4.1.3. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Then for all Q 2 C„k” with »k„Q” : k… = d 1 1D;S„φ„ ¯Q”” + O„1” and hR„Q” = d mφ(cid:3)(cid:25)(cid:3) 1 d mD;S„Q” = hφ(cid:3)(cid:25)(cid:3) 1R„φ„ ¯Q”” + O„1”: ∑n Proof. Since local heights are invariant under Galois maps, i=1 i=1 d∑ d∑ d∑ m∑d i=1 (cid:25)(cid:3) mφ(cid:3)φ(cid:3)(cid:25)(cid:3) mφ(cid:3)(cid:25)(cid:3) i=1 mD;S„ ˜Qi” + O„1” mD;S„(cid:25)i„ ¯Q”” + O„1” i D;S„ ¯Q” + O„1” m(cid:25)(cid:3) „ ¯Q” + O„1” i D;S 1D;S„ ¯Q” + O„1” 1D;S„φ„ ¯Q”” + O„1”: mD;S„Q” = = = = = = 1 d 1 d 1 d 1 d 1 d 1 d □ The proof for global heights is identical. Theorem 4.1.4. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and i=1 Pi. Let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = d and define the divisor D := (cid:22)„φ„ ¯Q”” < (cid:22)„φ(cid:3)(cid:25)(cid:3) „D””, 1 ∑n mD;S„Q” (cid:20) "hR„Q” + O„1”: Proof. note that φ(cid:3)(cid:25)(cid:3) 1 sion d (cid:0) 1. By Lemma 2.3.2(a), φ(cid:3)(cid:25)(cid:3) 1 „D” is an effective divisor on Symd„C” and so a closed subscheme of dimen- „D”””. Let A be an ample divisor on Jac„C”. „D” (cid:26) (cid:22)(cid:3)„(cid:22)„φ(cid:3)(cid:25)(cid:3) 1 32 note that φ(cid:3)(cid:25)(cid:3) 1R is also ample. By Corollary 4.1.2 there exists an integer m > 0 such that h(cid:22)(cid:3) A„E” (cid:20) mhφ(cid:3)(cid:25)(cid:3) 1R„E” + O„1” for all E 2 Symd„C”. Furthermore, for all Q 2 C„k” with »k„Q” : k… = d and (cid:22)„φ„ ¯Q”” < (cid:22)„φ(cid:3)(cid:25)(cid:3) 1 „D””, mφ(cid:3)(cid:25)(cid:3) 1D;S„φ„ ¯Q”” (cid:20) m(cid:22)(cid:3)„(cid:22)„φ(cid:3)(cid:25)(cid:3) = m(cid:22)„φ(cid:3)(cid:25)(cid:3) (cid:20) " m " m (cid:20) "hφ(cid:3)(cid:25)(cid:3) = 1D””;S„φ„ ¯Q”” + O„1” by Theorem 2.3.4(d) 1D”;S„(cid:22) ◦ φ„ ¯Q”” + O„1” hA„(cid:22) ◦ φ„ ¯Q”” + O„1” by Theorem 3.1.1 h(cid:22)(cid:3) A„φ„ ¯Q”” + O„1” 1R„φ„ ¯Q”” + O„1” by Corollary 4.1.2. By Lemma 4.1.3, the desired result holds on C. □ Conversely, we see the following lemma. Lemma 4.1.5. Let C be a nonsingular projective curve of genus g defined over a number field k. Let P 2 C„k” and let Q 2 C„k” with »k„Q” : k… = d. Suppose that (cid:22)„φ„ ¯Q”” 2 (cid:22)„φ(cid:3)(cid:25)(cid:3) „D””, that is 1 for some points Rj 2 C„k”. If g (cid:21) ˜Q1 + (cid:1) (cid:1) (cid:1) + ˜Qd (cid:24) P + R1 + (cid:1) (cid:1) (cid:1) + Rd(cid:0)1 8>>>><>>>>:1 when d = 2 2 when d = 3 then Rj , ˜Qk for any j and k. In particular, there exists a k-morphism (cid:27) : C ! P1 of degree d such that (cid:27)(cid:3)„(cid:27)„Q”” = ˜Q1 + (cid:1) (cid:1) (cid:1) + ˜Qd. Furthermore if d = 4 and g (cid:21) 5, then we also get such a k-morphism, provided ˜Q1+ ˜Q2+ ˜Q3+ ˜Q4 is not a sum of two hyperelliptic divisors on C. 33 Proof. note that because »k„P” : k… = 1 and »k„ ˜Q j” : k… = d > 1, we know ˜Q j , P for all 1 (cid:20) j (cid:20) d. Without loss of generality, assume Rd(cid:0)1 = ˜Qd, so that cancelling like terms would leave us with ˜Q1 + (cid:1) (cid:1) (cid:1) + ˜Qd(cid:0)1 (cid:24) P + R1 + (cid:1) (cid:1) (cid:1) + Rd(cid:0)2: ˜Q1 = P or g = 0. But ˜Q1 , P and by If d = 2, then this says ˜Q1 (cid:24) P, implying either hypothesis g (cid:21) 1. Thus we reach a contradiction. If d = 3, then this says ˜Q1 + ˜Q2 (cid:24) P + R1. If one of ˜Q2, this would imply R1 is equal to either ˜Q1 and ˜Q2 is equal to one of P ˜Q1 or and R1, then since P is not equal to ˜Q1 or ˜Q2, so P is linearly equivalent to one of ˜Q1 and ˜Q2 further implying g = 0, a contradiction. On the other hand, if the points on the left side of the linear equivalence are distinct from the points on the right side then there exists a morphism (cid:27) : C ! P1 of degree two such that (cid:27)„P” = (cid:27)„R1”. Since g (cid:21) 2, this says C is hyperelliptic. Since P is a k-rational point and the k(cid:157)k unique hyperelliptic map (cid:27) is a k-morphism, R1 must be k-rational as well. Let (cid:24) 2 Gal . Then (cid:24)„R2” (cid:24) (cid:24)„ ˜Q1 + ˜Q2 + ˜Q3”(cid:0) P (cid:0) R1 = ˜Q1 + ˜Q2 + ˜Q3 (cid:0) P (cid:0) R1 (cid:24) R2. Since g > 0, (cid:24)„R2” = R2. Since (cid:24) was arbitrary, R2 is k-rational, contradicting the assumption R2 = ˜Q3 since ˜Q3 is a cubic point over k. ( ) If d = 4, then this says ˜Q1 + ˜Q2 + ˜Q3 (cid:24) P +R1 +R2. Since g (cid:21) 5, there exists at most one trigonal map (up to automorphism of P1), at most one hyperelliptic map (up to automorphism of P1), and not both. If C has a unique trigonal map (cid:27) with (cid:27)„P” = (cid:27)„R1” = (cid:27)„R2”, by uniqueness, (cid:27) is k-rational, else we could use Galois conjugation to get a new hyperelliptic morphism. Since P is a k-rational point, we see that R1 and R2 are at worst quadratic conjugates. Thus for any (cid:24) 2 Gal , we have (cid:24)„R3” (cid:24) (cid:24)„ ˜Q1 + ˜Q2 + ˜Q3 + ˜Q4”(cid:0) P (cid:0) (cid:24)„R1 + R2” = ˜Q1 + ˜Q2 + ˜Q3 + ˜Q4 (cid:0) P (cid:0) R1 (cid:0) R2 (cid:24) R3. Since g > 0, (cid:24)„R3” = R3. Since (cid:24) was arbitrary, R3 is a k-rational point, contradicting the assumption R3 = ˜Q4, since ˜Q4 is a quartic point over k. k(cid:157)k ( ) If this does not give a trigonal map, say R2 = ˜Q3 as well, then C has a unique hyperelliptic map (cid:27) with (cid:27)„ ˜Q1” = (cid:27)„ ˜Q2” and (cid:27)„P” = (cid:27)„R1” and by the same argument in the d = 3 case, R1 is 34 ( ) k(cid:157)k k-rational. Thus for any (cid:24) 2 Gal , we have (cid:24)„R2 + R3” (cid:24) (cid:24)„ ˜Q1 + ˜Q2 + ˜Q3 + ˜Q4” (cid:0) P (cid:0) R1 = ˜Q1 + ˜Q2 + ˜Q3 + ˜Q4 (cid:0) P (cid:0) R1 (cid:24) R2 + R3. So (cid:27)„R2” = (cid:27)„R3”. But R2 = ˜Q3 and R3 = ˜Q4, so (cid:27)„ ˜Q3” = (cid:27)„ ˜Q4”, meaning ˜Q1 + ˜Q2 + ˜Q3 + ˜Q4 is a sum of two hyperelliptic divisors on C, □ contradicting the hypothesis. Theorem 4.1.6. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define i=1 Pi. Let (cid:27) : C ! P1 be a k-morphism of degree d. Let (cid:27)„Supp„D”” = the divisor D := fT1; T2; : : : ; Trg and define ni = ni„(cid:27)” := j(cid:27)(cid:0)1„Ti” \ Supp„D”j for i = 1; 2; : : : ; r. Reorder indices so that n1 (cid:21) n2 (cid:21) : : : (cid:21) nr. Let " > 0. Then for all Q 2 C n Supp„D” with »k„Q” : k… = d and (cid:27)„Q” 2 P1„k” ∑n mD;S„Q” (cid:20) „n1 + n2 + "”hR„Q” + O„1” Proof. Let (cid:19) : P1 ! Symd„C” (cid:26) Div+„C” be given by (cid:19)„T” = (cid:27)(cid:3)„T”. Then (cid:19) is an embedding with (cid:19)(cid:3)φ(cid:3)(cid:25)(cid:3) 1 = (cid:27)(cid:3). Since (cid:27)„Q” 2 P1„k”, any Galois map over k will only permute the set (cid:27)(cid:0)1„(cid:27)„Q”” implying (cid:19) ◦ (cid:27)„Q” = (cid:27)(cid:3)„(cid:27)„Q”” = ˜Q1 + (cid:1) (cid:1) (cid:1) + ˜Qd. Thus up to O„1” „D”;S„φ„ ¯Q”” by Lemma 4.1.3 „D”;S„i ◦ (cid:27)„Q”” „D”;S„(cid:27)„Q”” mD;S„Q” = = = mφ(cid:3)(cid:25)(cid:3) 1 mφ(cid:3)(cid:25)(cid:3) 1 mi(cid:3)φ(cid:3)(cid:25)(cid:3) 1 m(cid:27)(cid:3)„D”;S„(cid:27)„Q”” „n1 + n2 + "”h(cid:27)„Q”„(cid:27)„Q”” by Theorem 3.1.8 „n1 + n2 + "”h(cid:27)(cid:3)(cid:27)„Q”„Q” „n1 + n2 + "” 1 d 1 d 1 d 1 = d (cid:20) 1 d 1 d 1 d (cid:20) „n1 + n2 + "”„1 + "”hR„Q” = „n1 + n2 + "„n1 + n2 + 2””hR„Q”: „Q” + (cid:1) (cid:1) (cid:1) + h ˜Qd h ˜Q1 „Q” ] [ = = And replacing " with "(cid:157)„2d + 2” gives the desired result. □ 35 ∑n Lemma 4.1.7. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the i=1 Pi. Let " > 0. Suppose there exist only a finite number of k-morphisms C ! P1 divisor D := of degree d (modulo automorphisms of P1). Then for all Q 2 C„k” with »k„Q” : k… = d such that (cid:27)„Q” 2 P1„k” for some k-morphism (cid:27) : C ! P1 of degree d, we have mD;S„Q” (cid:20) „Nd„D” + "”hR„Q” + O„1” \ Supp„D” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” where Nd„D” := max C ! P1 of degree d. Proof. Let f(cid:27)jg be the unique set of representatives of the equivalence classes of k-morphisms C ! P1 of degree d (modulo automorphisms of P1) such that taken over all finite k-morphisms (cid:27) : n1„(cid:27)j” = „0” \ Supp„D” „1” \ Supp„D” (cid:12)(cid:12)(cid:12) and n2„(cid:27)j” = (cid:12)(cid:12)(cid:12)(cid:27)(cid:0)1 j (cid:12)(cid:12)(cid:12) {(cid:12)(cid:12)(cid:12)( (cid:12)(cid:12)(cid:12)(cid:27)(cid:0)1 j for all j, where ni is as defined in Theorem 4.1.6. By definition it is clear that N = n1„(cid:27)j” + n2„(cid:27)j” for some j. Let Q 2 C„k” with »k„Q” : k… = d such that (cid:27)„Q” 2 P1„k” for some k-morphism (cid:27) : C ! P1 of degree d. After composing with an automorphism of P1 we may assume (cid:27) = (cid:27)j for some j. Thus by Theorem 4.1.6, for all Q 2 C„k” with »k„Q” : k… = d such that (cid:27)„Q” 2 P1„k” for some k-morphism (cid:27) : C ! P1 of degree d, we have mD;S„Q” (cid:20) max j f„n1„(cid:27)j” + n2„(cid:27)j” + "”hR„Q” + Oj„1”g (cid:20) „Nd„D” + "”hR„Q” + O„1”: □ Corollary 4.1.8. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let " > 0. If g (cid:21) 2, then for all Q 2 C„k” with »k„Q” : k… = 2, ∑n mD;S„Q” (cid:20) „Nd„D” + "”hR„Q” + O„1” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ Supp„D” taken over all finite k-morphisms (cid:27) : {(cid:12)(cid:12)(cid:12)( where Nd„D” := max C ! P1 of degree 2. (cid:12)(cid:12)(cid:12)} (cid:12)(cid:12)(cid:12)} ) ) 36 Proof. For those Q 2 C„k” with (cid:22)„φ„ ¯Q”” < (cid:22)„φ(cid:3)(cid:25)(cid:3) „D”” the result follows from Theorem 4.1.4, 1 regardless of the value of Nd„D”. For those Q 2 C„k” with »k„Q” : k… = 2 and (cid:22)„φ„ ¯Q”” 2 (cid:22)„φ(cid:3)(cid:25)(cid:3) „D””, Lemma 4.1.5 implies there exists a morphism (cid:27) : C ! P1 of degree 2 such that 1 (cid:27)(cid:3)„(cid:27)„Q”” = φ„ ¯Q”, that is (cid:27)„Q” 2 P1„k”. Since g (cid:21) 2, such a hyperelliptic map is unique up to an automorphism of P1, so the result follows from Lemma 4.1.7. □ Corollary 4.1.9. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let " > 0. If g (cid:21) 3, then for all Q 2 C„k” with »k„Q” : k… = 3, ∑n {(cid:12)(cid:12)(cid:12)( (cid:12)(cid:12)(cid:12)} \ Supp„D” mD;S„Q” (cid:20) „Nd„D” + "”hR„Q” + O„1” ) (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” taken over all finite k-morphisms (cid:27) : where Nd„D” := max C ! P1 of degree 3. Proof. For those Q 2 C„k” with (cid:22)„φ„ ¯Q”” < (cid:22)„φ(cid:3)(cid:25)(cid:3) „D”” the result follows from Theorem 4.1.4, 1 regardless of the value of Nd„D”. For those Q 2 C„k” with »k„Q” : k… = 3 and (cid:22)„φ„ ¯Q”” 2 (cid:22)„φ(cid:3)(cid:25)(cid:3) „D””, Lemma 4.1.5 implies there exists a morphism (cid:27) : C ! P1 of degree 3 such that 1 (cid:27)(cid:3)„(cid:27)„Q”” = φ„ ¯Q”, that is (cid:27)„Q” 2 P1„k”. If g (cid:21) 5, by Theorem 3.3.2 such a trigonal morphism is unique up to an automorphism of P1. If g = 4, then by Theorem 3.3.5 there exist at most two such trigonal morphisms. If g = 3, then by Theorem 3.3.4 there are only finitely many trigonal □ k-morphisms. So the result follows from Lemma 4.1.7. 4.2 Quadratic Points on Elliptic Curves First, a general lemma. ∑n Lemma 4.2.1. Let C be a projective nonsingular curve of genus g defined over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor i=1 Pi. Let R 2 C„k” and let " > 0. Let d be a positive integer and let Z be a finite union of D := irreducible curves in Symd„C”. 37 (a) If g = 1 and d = 2 and there does not exist a subset fi1; : : : ; i4g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 (cid:24) Pi3 + Pi4 then for all Q 2 C„k” with »k„Q” : k… = 2 and φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „3 + "”hR„Q” + O„1”: (b) If g = 1 and d = 3 and there does not exist a subset fi1; : : : ; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 (cid:24) Pi4 + Pi5 + Pi6 then for all Q 2 C„k” with »k„Q” : k… = 3 such that φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1”: (c) If g = 2 and d = 3 and there does not exist a subset fi1; : : : ; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 (cid:24) Pi4 + Pi5 + Pi6 then for all Q 2 C„k” with »k„Q” : k… = 2 such that φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1”: Furthermore if there does not exist a subset fi1; : : : ; i5g (cid:26) f1; : : : ; ng and a point T 2 C„k” distinct from Pi1 ; Pi3 such that ; Pi2 Pi1 + Pi2 + Pi3 (cid:24) Pi4 + Pi5 + T then for all Q 2 C„k” with »k„Q” : k… = 3 such that φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1”: ∑r Proof. Let Z = Supp„φ(cid:3)(cid:25)(cid:3) j=1 Z j be the decomposition of Z into irreducible curves. Since φ„ ¯Q” < 1D” is nonempty, that is Z j is not a subset of 1D”, we may assume Z j n Supp„φ(cid:3)(cid:25)(cid:3) 38 Supp„φ(cid:3)(cid:25)(cid:3) 1Pi” for i = 1; : : : ; n. We think of Z j with the reduced induced closed subscheme structure and let (cid:26)j : Z j ! X be the closed immersion. If there are only finitely many points Q 2 C„k” such that »k„Q” : k… = d and φ„ ¯Q” 2 Z j, then we are done. Else Z j has infinitely many k-rational points, thus by Falting’s Theorem Z j has genus 0 or 1. If Z j has genus 1 then for those Q 2 C„k” with »k„Q” : k… = d and φ„ ¯Q” 2 (cid:26)j„Z j”, mD;S„Q” = = „φ„ ¯Q””” + O„1” j „φ„ ¯Q””” + O„1” by Lemma 3.1.1 1D;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 mφ(cid:3)(cid:25)(cid:3) 1D;S„(cid:26)(cid:0)1 φ(cid:3)(cid:25)(cid:3) m(cid:26)(cid:3) j 1R„(cid:26)(cid:0)1 h(cid:26)(cid:3) φ(cid:3)(cid:25)(cid:3) j 1R„φ„ ¯Q”” + O„1” hφ(cid:3)(cid:25)(cid:3) 1 d 1 d (cid:20) " d " d = "hR„Q” + O„1” by Lemma 4.1.3. = j Alternatively if Z j has genus 0 then, since abelian varieties admit no rational subvarieties, the image (cid:22)„Z j” 2 Jac„C” = C is a point. Thus Z j is contained in a fiber Y of (cid:22). Let E 2 Y. Then the subvariety Y (cid:26) Symd„C” may be thought of as the complete linear system jEj of the degree d divisor E on C. Furthermore the divisor φ(cid:3)(cid:25)(cid:3) 1PijY may be thought of as the linear system jE (cid:0) Pij + Pi, and thus as a linear subspace of the projective space Y. If d = g + 1, where g = 1 or 2, then by Riemann-Roch the complete linear systems of divisors of degree d all have dimension 1, so Y = P1. Since φ(cid:3)(cid:25)(cid:3) 1PijY is a linear subspace, it is either a single point, or all of Y. Furthermore, because Z j has dimension 1, it follows that Z j = Y. So either Z j (cid:26) Supp„φ(cid:3)(cid:25)(cid:3) 1Pi”, a contradiction, or (cid:26)(cid:3) 1PijY is a degree 1 divisor. 1Pi = φ(cid:3)(cid:25)(cid:3) φ(cid:3)(cid:25)(cid:3) j If g = 1 and d = 3, then by Riemann-Roch, the complete linear systems of divisors of degree 1PijY 1Pi are in 3-subgeneral position, thus so 1PijY. By Theorem 3:1:3, there exists a finite set H of lines in Y such that for all d all have dimension 2. Riemann-Roch also tells us that jE (cid:0) Pij has dimension 1 and so φ(cid:3)(cid:25)(cid:3) is a line in the projective plane Y. note the divisors φ(cid:3)(cid:25)(cid:3) are the lines φ(cid:3)(cid:25)(cid:3) 39 P 2 Y n∪ H2H H n∑ mφ(cid:3)(cid:25)(cid:3) 1DjY;S„P” = mφ(cid:3)(cid:25)(cid:3) 1PijY;S„P” + O„1” i=1 (cid:20) „5 + "”h„P” + O„1” by Theorem 3.1.3, where h is the canonical height on Y (cid:27) P2. Thus if Z j < H, then for all Q 2 C„k” with »k„Q” : k… = d and φ„ ¯Q” 2 (cid:26)j„Z j” 1 mD;S„Q” = 1D;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 3mφ(cid:3)(cid:25)(cid:3) 1D;S„(cid:26)(cid:0)1 1 „φ„ ¯Q””” + O„1” 3m(cid:26)(cid:3) φ(cid:3)(cid:25)(cid:3) = j 1R„(cid:26)(cid:0)1 (cid:20) 1 „5 + "”h(cid:26)(cid:3) φ(cid:3)(cid:25)(cid:3) 3 1 „5 + "”hφ(cid:3)(cid:25)(cid:3) 1R„φ„ ¯Q”” + O„1” 3 = „5 + "”hR„Q” + O„1” by Lemma 4.1.3. „φ„ ¯Q””” + O„1” = j j j Since this would satisfy every inequality in part (b), we need not consider this case any further. If 1Pi”, a contradiction, or (cid:26)(cid:3) Z j 2 H, then either Z j (cid:26) Supp„φ(cid:3)(cid:25)(cid:3) φ(cid:3)(cid:25)(cid:3) „Pi” can coincide, and if c divisors 1 φ(cid:3)(cid:25)(cid:3) (cid:26)(cid:3) j At most d of the divisors (cid:26)(cid:3) 1Pi has degree 1. ”; : : : ; (cid:26)(cid:3) j 1„Pic” φ(cid:3)(cid:25)(cid:3) φ(cid:3)(cid:25)(cid:3) 1„Pi1 j j coincide, then Pi1 + : : : + Pic + Rc+1 + : : : + Rd (cid:24) E ∑p for some Rc+1; : : : ; Rd 2 C„k”. Let (cid:26)(cid:3) c1; c2; (cid:1) (cid:1) (cid:1) ; cp with c1 (cid:21) c2 (cid:21) (cid:1) (cid:1) (cid:1) (cid:21) cp. Then for Q 2 C„k” with »k„Q” : k… = d and φ„ ¯Q” 2 Z j m=1 cmEm for some positive integers „D” = φ(cid:3)(cid:25)(cid:3) 1 j mD;S„Q” = = „D”;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 „D”jYjk mφ(cid:3)(cid:25)(cid:3) 1 ;S„φ„ ¯Q”” + O„1” mφ(cid:3)(cid:25)(cid:3) 1 „c1 + c2 + "”hφ(cid:3)(cid:25)1(cid:3)„R”jYjk „c1 + c2 + "”hφ(cid:3)(cid:25)1(cid:3)„R”„φ„ ¯Q”” + O„1” 1 d 1 d (cid:20) 1 d 1 d = „c1 + c2 + "”hR„Q” + O„1” by Lemma 4.1.3: = „φ„ ¯Q”” + O„1” by Lemma 3.1.8 40 If there does not exist a subset fi1; : : : ; i2dg (cid:26) f1; : : : ; ng such that (cid:24) Pid+1 + : : : + Pi2d Pi1 + : : : + Pid then for all Q 2 C„k” with »k„Q” : k… = d such that φ„ ¯Q” 2 Z j we have mD;S„Q” (cid:20) „2d (cid:0) 1 + "”hR„Q” + O„1”: Finally, in the special case where g = 2 and d = 3, if there does not exist a subset fi1; : : : ; i5g (cid:26) f1; : : : ; ng and a point T 2 C„k” distinct from Pi1 Pi1 + Pi2 + Pi3 ; Pi3 such that ; Pi2 (cid:24) Pi4 + Pi5 + T then either such a linear equivalence does not exist for any T 2 C„k” , in which case, for all Q 2 C„k” with »k„Q” : k… = 3 such that φ„ ¯Q” 2 Z, we have mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1”: Else, there does exist a subset fi1; : : : ; i5g (cid:26) f1; : : : ; ng and a point T 2 C„k” such that Pi1 + Pi2 + Pi3 (cid:24) Pi4 + Pi5 + T ; Pi2 ; Pi3 ; Pi2 ; Pi3 1PijY has degree 1 as shown above. g. However since every other point in the linear with either T < C„k” or T 2 fPi1 equivalence is k-rational and thus fixed under Galois maps over k, it follows that T is also fixed, g, say T = Pi3. But that would imply and so T 2 C„k” automatically. Thus T 2 fPi1 φ(cid:3)(cid:25)(cid:3) 1Pi3 intersects the projective line Y = Z j in more than one point, contradicting the fact that φ(cid:3)(cid:25)(cid:3) 1Pi = φ(cid:3)(cid:25)(cid:3) (cid:26)(cid:3) j Since there are only finitely many irreducible components Z j in Z, we may take the maximum of the constants O„1” in each of the above inequalities to prove the same inequalities holds for all Q 2 C„k” with »k„Q” : k… = d such that φ„ ¯Q” 2 Z. □ Proposition 4.2.2. Suppose g = 1. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be i=1 Pi. Let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = 2 distinct and define the divisor D := ∑n mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1” 41 and if there does not exist a subset fi1; : : : ; i4g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 then for all Q 2 C„k” with »k„Q” : k… = 2 we have (cid:24) Pi3 + Pi4 mD;S„Q” (cid:20) „3 + "”hR„Q” + O„1”: „R” + (cid:25)(cid:3) 2 „R” + (cid:25)(cid:3) 2 „R”” = 2φ(cid:3)(cid:25)(cid:3) 1 „R” is an ample divisor on Sym2„C” and therefore so is φ(cid:3)(cid:25)(cid:3) 1 Proof. We may assume n > 2. Since (cid:25)(cid:3) 1 φ(cid:3)„(cid:25)(cid:3) 1 Since any two points on a curve are numerically equivalent, we also have φ(cid:3)(cid:25)(cid:3) 1 equivalent to φ(cid:3)(cid:25)(cid:3) 1 of φ(cid:3)(cid:25)(cid:3) 1 „R” is an ample divisor on C2, we see that „R”. „Pi” is numerically „Pi” are in general position since the intersection ” in Sym2„C” is the point Pi1 + Pi2. „R” for all i. The divisors φ(cid:3)(cid:25)(cid:3) 1 For each v 2 S, choose two of the divisors φ(cid:3)(cid:25)(cid:3) 1 „Pi” to be D1;v and D2;v. Then applying Theorem 3.1.6 with X = Sym2„C”, A = φ(cid:3)(cid:25)(cid:3) „R” and di = 1, for 1 (cid:20) i (cid:20) 2, we see there exists a proper 1 Zariski-closed subscheme Z (cid:26) Sym2„C” such that for all points Q 2 C„k” with »k„Q” : k… = 2 and „Pi1 ” and φ(cid:3)(cid:25)(cid:3) 1 „Pi2 φ„ ¯Q” 2 Sym2„C”„k” n Z,∑ 2∑ ;v„φ„ ¯Q”” < „3 + "”hφ(cid:3)(cid:25)(cid:3) 1 „P1”„φ„ ¯Q””: i=1 v2S (cid:21)Di;v Since at most two of the divisors φ(cid:3)(cid:25)(cid:3) „Pi” intersect at a point, we see that a point in Sym2„C”„k”n Z 1 can be v-adically close to at most two of these divisors for each v 2 S. Points v-adically closest to D1;v and D2;v cannot be (arbitrarily) v-adically close to any other of the divisors φ(cid:3)(cid:25)(cid:3) „Pi”, that 1 „Pi” are bounded for such points. Thus for all points Q 2 C„k” with is those local heights (cid:21)φ(cid:3)(cid:25)(cid:3) 1 »k„Q” : k… = 2 and φ„ ¯Q” 2 Sym2„C”„k” n Z with φ„ ¯Q” v-adically closest to D1;v and D2;v for each 42 v 2 S, mD;S„Q” = = „φ„ ¯Q”” + O„1” ∑ n∑ 1 „D”;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 2mφ(cid:3)(cid:25)(cid:3) 2∑ ∑ 1 (cid:21)φ(cid:3)(cid:25)(cid:3) „Pi”;v 2 1 ;v„φ„ ¯Q”” + O„1” „R”„φ„ ¯Q”” + O„1” 1 2 < 1 2 = „3 + "”hR„Q” + O„1” by Lemma 4.1.3. (cid:21)Di;v v2S „3 + "”hφ(cid:3)(cid:25)(cid:3) 1 i=1 1 v2S i=1 = Since this inequality does not depend upon the choices of D1;v and D2;v for each place v 2 S, the result follows for all Q 2 C„k” with »k„Q” : k… = 2 and φ„ ¯Q” 2 Sym2„C”„k” n Z. Since φ„ ¯Q” 2 Sym2„C”„k” it only remains to consider the case where φ„ ¯Q” 2 Z. By Lemma 4.2.1, then for all Q 2 C„k” with »k„Q” : k… = 2 such that φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1” and if there does not exist a subset fi1; i2; i3; i4g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 (cid:24) Pi3 + Pi4 then for all Q 2 C„k” with »k„Q” : k… = 2 such that φ„ ¯Q” 2 Z we have mD;S„Q” (cid:20) „3 + "”hR„Q” + O„1”: Combining the above inequalities, we get that the proposition holds for all Q 2 C„k” with »k„Q” : k… = 2. □ Corollary 4.2.3. Let C be a nonsingular projective curve of genus-1 defined over a number field k and let R 2 C„k”. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = 2, ∑n mD;S„Q” (cid:20) „N2„D” + "”hR„Q” + O„1” 43 {(cid:12)(cid:12)(cid:12)( ) (cid:12)(cid:12)(cid:12)} \ Supp„D” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” taken over all k-morphisms (cid:27) : C ! P1 where N2„D” := max of degree 2. If no such k-morphism exists, we say N2„D” = 0. Proof. First note that since degree 2 morphisms are at most 2-to-1, we have N2„D” (cid:20) 4. By Riemann-Roch, the dimension of the complete linear system jEj of any divisor E on C of positive degree is given by dimjEj = deg„E” (cid:0) 1. If there are only finitely many points Q 2 C„k” with »k„Q” : k… = 2, then mD;S„Q” (cid:20) O„1”, so we assume there are infinitely many points Q 2 C„k” with »k„Q” : k… = 2. In particular, there exists a pair of conjugate k-quadratic points Q1; Q2 2 C„k”. Since jQ1 + Q2j=1 and jQ1 + Q2 (cid:0) Tj = 0 for all T 2 C„k”, it follows that Q1 + Q2 is a degree 2 divisor defined over k, whose complete linear system, jQ1 + Q2j, is base-point-free of dimension 1. Thus it induces a degree 2 morphism ! : C ! P1 defined over k. If deg„D” = 1, then after composing with a k-automorphism of P1, we may assume !„P1” = 0, so N2„D” = 1. If deg„D” = 2, then after composing with a k-automorphism of P1, we may assume !„P1” = 0 and either !„P2” = 0 or !„P2” = 1, so N2„D” = 2. In both these cases N = deg„D”, so we have mD;S„Q” (cid:20) hD„Q” by Definition 2.2.8 (cid:20) „deg„D” + "”hR„Q” + O„1” by Theorem 2.2.12 = „Nd„D” + "”hR„Q” + O„1”: Suppose deg„D” (cid:21) 3. Then since jP1 + P2j=1 and jP1 + P2 (cid:0) Tj = 0 for all T 2 C„k”, it follows that P1 + P2 is a degree 2 divisor defined over k, whose complete linear system, jP1 + P2j, is base-point-free of dimension 1. Thus it induces a degree 2 morphism (cid:27) : C ! P1 defined over k, such that (cid:27)„P1” = (cid:27)„P2”. By composing with an automorphism of P1, we may assume (cid:27)„P1” = (cid:27)„P2” = 0 and (cid:27)„P3” = 1. Thus we see N2„D” (cid:21) 3. If N2„D” = 4, then the inequality follows from Proposition 4.2.2. Conversely, if there exists a subset fi1; i2; i3; i4g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 ” = (cid:27)„Pi2 then there exists a k-morphism (cid:27) : C ! P1 of degree 2 such that (cid:27)„Pi1 (cid:27)„Pi3 ” = 1, so that N2„D” = 4. ” = (cid:27)„Pi4 (cid:24) Pi3 + Pi4, ” = 0 and 44 Thus if N2„D” = 3, then such a subset of f1; : : : ; ng does not exist and the inequality again □ follows from Proposition 4.2.2. 4.3 Cubic Points on Curves of Low Genus It remains to show the main theorem holds for points of degree 3 on curves of genus 1 and 2. Theorem 4.3.1. Let X be an irreducible projective surface defined over a number field k. Let S be a finite set of places of k and let n 2 N. For each v 2 S, let D1;v; : : : ; Dn;v be effective divisors on X, defined over k having no irreducible components in common in their supports. Suppose there exists an ample divisor A and positive integers di;v such that Di;v is numerically equivalent to di;v A for all i and for all v 2 S. Suppose there exists a nonconstant k-morphism (cid:27) : X ! E for some elliptic curve E defined over k. Let " > 0. Then there exists a proper closed subscheme Z (cid:26) X such that for all points P 2 X„k” n Z,∑ n∑ (cid:20) „3 + "”hA„P” + O„1”: Proof. Let R 2 E„k”. By Theorem 3.1.1, for any Q 2 E„k” we have v2S ;v„P” (cid:21)Di;v di;v i=1 mQ;S„P” (cid:20) "hR„P” + O„1” for all P 2 E„k” n fQg. Thus for any Q 2 E„k” m(cid:27)(cid:3)Q;S„P” = mQ;S„(cid:27)„P”” + O„1” (cid:20) "hR„(cid:27)„P”” + O„1” = "h(cid:27)(cid:3)R„P” + O„1” for all P 2 X„k” n (cid:27)(cid:3)„Q”. By Corollary 4.1.2, there exists an M 2 N, depending only on R and A, such that for all P 2 X„k” "h(cid:27)(cid:3)R„P” (cid:20) "M (cid:1) hA„P” + O„1”: By replacing " with "(cid:157)M we can say that for any Q 2 E„k” m(cid:27)(cid:3)Q;S„P” (cid:20) "hA„P” + O„1” 45 for all P 2 X„k” n (cid:27)(cid:3)„Q”. Therefore, by Lemma 2.3.2(a) and Theorem 2.3.4(d), for any Q 2 X„k” we have mQ;S„P” (cid:20) m(cid:27)(cid:3)„(cid:27)„Q””;S„P” + O„1” (cid:20) "hA„P” + O„1” for all P 2 X„k” n (cid:27)(cid:3)„(cid:27)„Q””. For each v 2 S, let ∪ Yv = 1(cid:20)i< j(cid:20)n Supp„Di;v” \ Supp„Dj;v”: Since no two of the divisors share irreducible components in common, it follows that Yv is a finite set of points. Partition the set X by which of the points of Yv each point is v-adically closest to. note that if a point is v-adically closest to two points of Yv, then it cannot be v-adically close to any of the divisors Di;v, that is all of the local heights (cid:21)Di;v ;v will be bounded. Consider those points P that are v-adically closest to Q 2 Yv. For each pair of indices i and j we have, by definition, (cid:21)Di;v \Dj;v ;v„P” = minf(cid:21)Di;v ;v„P”; (cid:21)Dj;v ;v„P”g for all P 2 X„k” n Supp„Di;v” \ Supp„Dj;v”. But if Q 2 Supp„Di;v” \ Supp„Dj;v”, we have (cid:21)Q;v„P” = minf(cid:21)Di;v ;v„P”; (cid:21)Dj;v ;v„P”g + O„1” for those points P 2 X„k” n fQg such that Q is the v-adically closest point to P out of all of the points in Yv. Thus we have minf(cid:21)Di;v ;v„P”; (cid:21)Dj;v ;v„P”g (cid:20) "hA„P” + O„1” for all such P 2 X„k”n (cid:27)(cid:3)„(cid:27)„Q””. Since we can repeat this for each pair of indices i and j, it follows ;v„P”, can be greater than " (cid:1) hA„P” for that at most one of the local heights (cid:21)Di;v infinitely many of the points P 2 X„k” n (cid:27)(cid:3)„(cid:27)„Q”” that are v-adically closest to Q. As such ;v„P”, say (cid:21)DiQ;v ;v„P” (cid:21)Di;v di;v (cid:20) ;v„P” (cid:21)DiQ;v di;v + "„n (cid:0) 1”hA„P” + O„1” ∑ i 46 (∪n ) that are v-adically closest to Q. for all P 2 X„k” n i=1 Supp„Di;v” [ (cid:27)(cid:3)„(cid:27)„Q”” Repeating this process for each v 2 S gives a collection of partitions, which together make a finer partition of X into those Qv 2 Yv that the points are closest to for each v 2 S together with a proper (since (cid:27) is nonconstant) closed subscheme of X, Then we have ∑ n∑ for all P 2 X„k” n∪ v2S v2S i=0 ∪ ∪ v2S Q2Yv (cid:27)(cid:3)„(cid:27)„Q””: Z1 = ∑ ) ;v„P” (cid:20) v2S (∪n (cid:21)DiQv ;v di;v ;v„P” (cid:21)Di;v di;v i=1 Supp„Di;v” [ (cid:27)(cid:3)„(cid:27)„Qv”” ∑ ∪ ;v„P” + "„n (cid:0) 1”jSjhA„P” + O„1” v 2 S. But by Theorem 3.1.6, there exists a proper closed subscheme Z2 (cid:26) X such that that are v-adically closest to Qv for each v2S for all P 2 X„k” n Z2. Let Z = v2S ;v„P” (cid:21)Di;v di;v ∑ n∑ v2S i=0 ∪n (cid:20) „3 + "”hA„P” (cid:21)DiQv ;v di;v i=1 Supp„Di;v” [ Z1 [ Z2. Then for all P 2 X„K” n Z (cid:20) 3hA„P” + " „1 + „n (cid:0) 1”jSj” hA„P” + O„1”: Thus by replacing " with "(cid:157)„1 + „n (cid:0) 1”jSj”, we obtain the desired result. □ Proposition 4.3.2. Let C be an elliptic curve defined over k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let R 2 C„k” and let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = 3 ∑n mD;S„Q” (cid:20) „6 + "”hR„Q” + O„1” and if there does not exist a subset fi1; : : : ; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 then for all Q 2 C„k” with »k„Q” : k… = 3 we have (cid:24) Pi4 + Pi5 + Pi6 mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1”: 47 Proof. By Theorem 3.1.6, there exists a proper closed subscheme Z (cid:26) Sym3„C” such that for all Q 2 C„k” such that »k„Q” : k… = 3 and φ„ ¯Q” < Z we have 1 mD;S„Q” = 3mφ(cid:3)(cid:25)(cid:3) (cid:20) 1 „4 + "”hφ(cid:3)(cid:25)(cid:3) 3 = „4 + "”hR„Q” + O„1” by Lemma 4.1.3. 1D;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 1R„Q” + O„1” by Theorem 3.1.6 Let Z j be an irreducible component of Z. Then (cid:22)jZ j is a k-morphism from the irreducible surface Z j to an elliptic curve Jac„C” = C. If (cid:22)jZ j is a constant map then Z j is contained in a fiber of (cid:22). But the fibers of (cid:22) are 2-dimensional by Riemann-Roch, so it follows that Z j is a fiber of (cid:22), so there exists an E 2 Div„C” such that Z j = jEj (cid:27) P2. Since φ(cid:3)(cid:25)(cid:3) 1Pi is the set of effective divisors on C with Pi in their support, we see that it cuts out the 1-dimensional linear system jE (cid:0) Pij + Pi in jEj. Furthermore the divisors „φ(cid:3)(cid:25)(cid:3) 1Pi”jZ j on Z j are in 3-subgeneral position. Thus we may apply Theorem 3.1.3 with r = 2 and Li = „φ(cid:3)(cid:25)(cid:3) 1Pi”jZ j to get mD;S„Q” = = 1D;S„φ„ ¯Q”” + O„1” by Lemma 4.1.3 mLi;S„φ„ ¯Q”” + O„1” n∑ 1 3mφ(cid:3)(cid:25)(cid:3) 1 3 (cid:20) 1 3 = „5 + "”hR„Q” + O„1” by Lemma 4.1.3. 1R„Q” + O„1” by Theorem 3.1.3 i=1 „5 + "”hφ(cid:3)(cid:25)(cid:3) for all Q 2 C„k” such that »k„Q” : k… = 3 and φ„ ¯Q” 2 Z j n Yj, where Yj is a union of hyperplanes Yjk (cid:26) Z j (cid:27) P2. By Lemma 4.2.1, if there does not exist a subset fi1; i2; i3; i4; i5; i6g (cid:26) f1; : : : ; ng (cid:24) Pi4 + Pi5 + Pi6, then for all Q 2 C„k” with »k„Q” : k… = 3 and φ„ ¯Q” 2 Yj such that Pi1 + Pi2 + Pi3 otherwise mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1” mD;S„Q” (cid:20) „6 + "”hR„Q” + O„1”: 48 Combining the above inequalities we get the result on all of Z j. Since there are only finitely many irreducible components Z j in Z, we may take the maximum of all of the above constants in O„1” to get the result for all Q 2 C„k” with »k„Q” : k… = 3. □ Lemma 4.3.3. Let C be a nonsingular projective curve of genus 1 over an algebraic number field k. Let d (cid:21) 3 be an integer and let P1; : : : ; P2d(cid:0)1 2 C„k” be distinct. Let P2d(cid:0)1 be the identity element of the elliptic curve C. Suppose that for every morphism (cid:27) : C ! P1 over k of degree d we have (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ fP1; : : : ; P2d(cid:0)1g (cid:12)(cid:12)(cid:12) (cid:20) 2d (cid:0) 2: Then for every i 2 f1; : : : ; 2d (cid:0) 2g, the inverse of Pi in the elliptic curve C is Pj for some j 2 f1; : : : ; 2d (cid:0) 2g n fig. Proof. We proceed by induction on the number of pairs of inverse points identified. Trivially, we have already identified zero such pairs. Suppose we have identified m pairs already and that m < d (cid:0) 1. Let M (cid:26) f1; : : : ; 2d (cid:0) 2g be a set of representative indices, each pair of inverse points being represented by exactly one index, so that jMj = m. For each i 2 M, define i(cid:3) to be the index of the inverse of Pi, that is to say Pi + Pi(cid:3) (cid:24) 2P2d(cid:0)1. Let I (cid:26) f1; : : : ; 2d (cid:0) 2g be a subset disjoint from M with jIj = d (cid:0) 2(cid:0) m and let J = f1; : : : ; 2d (cid:0) 2g n „M [ I”. By Riemann-Roch, there exists a unique R 2 C„k” such that ∑ ∑ R + P2d(cid:0)1 + Pi (cid:24) Pj : j2J i2M[I j j 2 Jg, then this linear equivalence defines a Furthermore R 2 C„k” by uniqueness. If R < fPj base-point-free 1-dimensional linear system over k of degree d. Thus there exists a corresponding morphism (cid:27) : C ! P1 over k with (cid:27)(cid:0)1„0” = fR; P2d(cid:0)1g [ fPi j i 2 M [ Ig and (cid:27)(cid:0)1„1” = fPj j j 2 Jg, so that (cid:12)(cid:12)(cid:12) = 2d (cid:0) 1: (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” (cid:12)(cid:12)(cid:12)( (cid:12)(cid:12)(cid:12)( ) ) \ fP1; : : : ; P2d(cid:0)1g 49 As this would contradict the hypothesis, we conclude there exists an r 2 J such that R = Pr. That is to say ∑ ∑ Pj : Once again by Riemann-Roch, there exists a T 2 C„k” such that j2Jnfrg i2M[I P2d(cid:0)1 + Pi (cid:24) (4.1) Pi: ∑ i2M[I j j 2 J n frgg. But if T = P2d(cid:0)1, P2d(cid:0)1 + j2Jnfrg Pj (cid:24) T + Pr + By the same reasoning as above, either T = P2d(cid:0)1 or T 2 fPj then ∑ Pj (cid:24) Pr + Pi: i2M[I ∑ ∑ j2Jnfrg Comparing this to equivalence (4.1) reveals that Pr (cid:24) P2d(cid:0)1. But according to Riemann-Roch, if two points on C are linearly equivalent, then they are equal, that is Pr = P2d(cid:0)1, a contradiction since r , 2d (cid:0) 1 and the Pi are distinct. Therefore there exists a t 2 J n frg such that T = Pt. That is to say ∑ ∑ ∑ ∑ i2M[I P2d(cid:0)1 + j2Jnfr;tg Pj (cid:24) Pr + Pi: 2P2d(cid:0)1 + Pi (cid:24) i2f1;:::;2d(cid:0)2gnfr;tg i2f1;:::;2d(cid:0)2g (4.2) Pi: Adding equivalences (4.1) and (4.2) together we get And after cancelling, we get Pr + Pt (cid:24) 2P2d(cid:0)1, which is exactly the condition that Pr and Pt are inverse points in the elliptic curve C with identity P2d(cid:0)1. Thus we have determined m + 1 distinct pairs of inverse points on C. Therefore, by induction we conclude that we can find d (cid:0) 1 distinct pairs of inverse points, which must consist of P1; : : : ; P2d(cid:0)2 as desired. □ Proposition 4.3.4. Let C be a nonsingular projective curve of genus 1 over an algebraic number field k. Let P1; : : : ; P5 2 C„k” be distinct. Suppose that for every morphism (cid:27) : C ! P1 over k of degree 3 we have (cid:12)(cid:12)(cid:12)( (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ fP1; : : : ; P5g (cid:12)(cid:12)(cid:12) (cid:20) 4: ) 50 Then if one of the five given points is chosen to be the identity element of the elliptic curve C, the five given points then form a subgroup of C„k” of order 5. Proof. Without loss of generality, let P5 be the identity element of the elliptic curve C. By Lemma 4.3.3, for every k 2 f1; : : : ; 4g there exists a k(cid:3) 2 f1; : : : ; 4g n fkg such that Pk + Pk(cid:3) (cid:24) 2P5. Let these four non-identity points be called Pi; Pi(cid:3); Pj; Pj(cid:3). By Riemann-Roch, there exists an R 2 C„k” such that R + P5 + Pi(cid:3) (cid:24) Pi + Pj + Pj(cid:3) : (4.3) Since we assume the above morphism (cid:27) does not exist, it must be the case that either R = Pi, R = Pj, or R = Pj(cid:3). But if R = Pi, then equivalence (4.3) simplifies to P1(cid:3) + P5 + Pi(cid:3) (cid:24) Pi + 2P5: So Pi(cid:3) (cid:24) P5. By Riemann-Roch, this implies Pi(cid:3) = P5, a contradiction. Therefore R = Pr with r 2 f j; j(cid:3)g. In this case, equivalence (4.3) simplifies to Pr + P5 + Pi(cid:3) (cid:24) Pi + 2P5: Adding Pi to both sides and cancelling like terms gives Pr + P5 (cid:24) 2Pi. As elements of the elliptic curve C, this says 2Pi = Pr. Taking inverses, we also get 2Pi(cid:3) = Pr(cid:3). Repeating this process for the linear equivalence T + P5 + Pj(cid:3) (cid:24) Pj + Pi + Pi(cid:3) gives T = Pt for some t 2 fi; i(cid:3)g and so Pt + P5 (cid:24) 2Pj. As elements of the elliptic curve C, this says 2Pj = Pt and 2Pj(cid:3) = Pt(cid:3). Since 2P5 = P5, we see that these five points are closed under doubling. Once again by Riemann-Roch, there exists a point A 2 C„k” such that A + Pi + Pj (cid:24) Pi(cid:3) + Pj(cid:3) + P5: Since we assume the above morphism (cid:27) does not exist, it must be the case that either A = Pi(cid:3), A = Pj(cid:3), or A = P5. If A = P5 however, then Pi + Pj = Pi(cid:3) + Pj(cid:3) and adding Pi + Pj to both sides 51 gives Pr + Pt = 0, which is not possible, given r 2 f j; j(cid:3)g and t 2 fi; i(cid:3)g. Thus A 2 fPi(cid:3); Pj(cid:3)g and Pi + Pj = Pu for some u 2 fi(cid:3); j(cid:3)g. Taking inverses gives Pi(cid:3) + Pj(cid:3) = Pu(cid:3). Similarly, Pi + Pj(cid:3) = Pv for some v 2 fi(cid:3); jg and so Pi(cid:3) + Pj = Pv (cid:3). Thus the set fP1; : : : ; P5g is closed under addition, and □ therefore forms a subgroup of order 5 in the elliptic curve C. It is worth noting that this obstruction does not appear in higher degree examples, as demon- strated in the following. Proposition 4.3.5. Let C be a nonsingular projective curve of genus 1 over an algebraic number field k. Let d (cid:21) 4 be an integer and let P1; : : : ; P2d(cid:0)1 2 C„k” be distinct. Then there exists a morphism (cid:27) : C ! P1 over k of degree d such that (cid:12)(cid:12)(cid:12)( ) (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ fP1; : : : ; P2d(cid:0)1g (cid:12)(cid:12)(cid:12) = 2d (cid:0) 1: Proof. By way of contradiction, assume such a morphism does not exist. Let P2d(cid:0)1 be the identity element of the elliptic curve C. Then by Lemma 4.3.3, for every i 2 f1; : : : ; 2d (cid:0) 2g there exists an i(cid:3) 2 f1; : : : ; 2d (cid:0) 2g n fig such that Pi + Pi(cid:3) (cid:24) 2P2d(cid:0)1. Let M be a set of representative indices of these pairs, that is for each i 2 f1; : : : ; 2d (cid:0) 2g, either i 2 M or i(cid:3) 2 M and not both, so that jMj = d (cid:0) 1. Let M1 = M n f1; 1(cid:3)g and note jM1j = d (cid:0) 2. The proof is broken into two cases, but those cases are treated very similarly. First suppose d ∑ is even. Let K (cid:26) M1 with jKj = d(cid:0)2 2 . By Riemann-Roch, there exists an R 2 C„k” such that „Pi + Pi(cid:3)” (cid:24) P2d(cid:0)1 + P1(cid:3) + „Pi + Pi(cid:3)”: R + P1 + (4.4) ∑ i2K i2M1nK Since we assume the desired morphism does not exist, it must be the case that either R = P2d(cid:0)1, R = P1(cid:3), or R 2 fPi j i 2 K or i(cid:3) 2 K}. But if R = P2d(cid:0)1, then equivalence (4.4) simplifies to P2d(cid:0)1 + P1 + „d (cid:0) 2”P2d(cid:0)1 (cid:24) P2d(cid:0)1 + P1(cid:3) + „d (cid:0) 2”P2d(cid:0)1: So P1 (cid:24) P1(cid:3). By Riemann-Roch, this implies P1 = P1(cid:3), a contradiction to Lemma 4.3.3. And if R = P1(cid:3), then equivalence (4.4) simplifies to P1(cid:3) + P1 + „d (cid:0) 2”P2d(cid:0)1 (cid:24) P2d(cid:0)1 + P1(cid:3) + „d (cid:0) 2”P2d(cid:0)1: 52 So P1 (cid:24) P2d(cid:0)1. By Riemann-Roch, this implies P1 = P2d(cid:0)1, a contradiction to the hypothesis. Therefore R = Pr for some r 2 f1; : : : ; 2d (cid:0) 2g with either r 2 K or r(cid:3) 2 K. In this case equivalence (4.4) simplifies to Pr + P1 + „d (cid:0) 2”P2d(cid:0)1 (cid:24) P2d(cid:0)1 + P1(cid:3) + „d (cid:0) 2”P2d(cid:0)1: Adding P1(cid:3) to both sides and cancelling like terms gives Pr + P2d(cid:0)1 (cid:24) 2P1(cid:3) : Now choose K′ (cid:26) M1 n fr; r(cid:3)g with jK′j = d(cid:0)2 (4.5) 2 . note this subset exists because jM1 n fr; r(cid:3)gj = d (cid:0) 3 and d(cid:0)2 2 (cid:20) d (cid:0) 3 () d (cid:21) 4. By Riemann-Roch, there exists an T 2 C„k” such that „Pi + Pi(cid:3)” (cid:24) P2d(cid:0)1 + P1(cid:3) + „Pi + Pi(cid:3)”: (4.6) ∑ T + P1 + i2M1nK′ ∑ i2K′ Once again this implies that T = Pt for some t 2 f1; : : : ; 2d (cid:0) 2g with either t 2 K′ or t(cid:3) 2 K′. Therefore equivalence (4.6) simplifies to Pt + P1 + „d (cid:0) 2”P2d(cid:0)1 (cid:24) P2d(cid:0)1 + P1(cid:3) + „d (cid:0) 2”P2d(cid:0)1: Adding P1(cid:3) to both sides and cancelling like terms gives Pt + P2d(cid:0)1 (cid:24) 2P1(cid:3) : (4.7) Comparing equivalences (4.5) and (4.7) reveals that Pr (cid:24) Pt, which implies that Pr = Pt, but that implies r = t, a contradiction to the construction of K′. Therefore our initial assumption must have been false and there exists a morphism (cid:27) : C ! P1 over k of degree d such that (cid:12)(cid:12)(cid:12)( (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ fP1; : : : ; P2d(cid:0)1g Now suppose d is odd. Let K (cid:26) M1 with jKj = d(cid:0)1 2 . By Riemann-Roch, there exists an R 2 C„k” such that R + P2d(cid:0)1 + P1 + „Pi + Pi(cid:3)” (cid:24) P1(cid:3) + „Pi + Pi(cid:3)”: (4.8) ) ∑ i2M1nK 53 (cid:12)(cid:12)(cid:12) = 2d (cid:0) 1: ∑ i2K Since we assume the desired morphism does not exist, it must be the case that either R = P1(cid:3) or R 2 fPi j i 2 K or i(cid:3) 2 K}. But if R = P1(cid:3), then equivalence (4.8) simplifies to P1(cid:3) + P2d(cid:0)1 + P1 + „d (cid:0) 3”P2d(cid:0)1 (cid:24) P1(cid:3) + „d (cid:0) 1”P2d(cid:0)1: So P1 (cid:24) P2d(cid:0)1. By Riemann-Roch, this implies P1 = P2d(cid:0)1, a contradiction. Therefore R = Pr for some r 2 f1; : : : ; 2d (cid:0) 2g with either r 2 K or r(cid:3) 2 K. In this case equivalence (4.8) simplifies to Pr + P2d(cid:0)1 + P1 + „d (cid:0) 3”P2d(cid:0)1 (cid:24) P1(cid:3) + „d (cid:0) 1”P2d(cid:0)1: Adding P1(cid:3) to both sides and cancelling like terms gives Pr + P2d(cid:0)1 (cid:24) 2P1(cid:3) : Now choose K′ (cid:26) M1nfr; r(cid:3)g with jK′j = d(cid:0)1 ∑ and d(cid:0)1 2 (4.9) 2 . note this subset exists because jM1nfr; r(cid:3)gj = d(cid:0)3 (cid:20) d (cid:0) 3 () d (cid:21) 5. By Riemann-Roch, there exists an T 2 C„k” such that T + P2d(cid:0)1 + P1 + „Pi + Pi(cid:3)” (cid:24) P1(cid:3) + „Pi + Pi(cid:3)”: (4.10) i2M1nK′ Once again this implies that T = Pt for some t 2 f1; : : : ; 2d (cid:0) 2g with either t 2 K′ or t(cid:3) 2 K′. Therefore equivalence (4.10) simplifies to ∑ i2K′ Pt + P2d(cid:0)1 + P1 + „d (cid:0) 3”P2d(cid:0)1 (cid:24) P1(cid:3) + „d (cid:0) 1”P2d(cid:0)1: Adding P1(cid:3) to both sides and cancelling like terms gives Pt + P2d(cid:0)1 (cid:24) 2P1(cid:3) : (4.11) Comparing equivalences (4.9) and (4.11) reveals that Pr (cid:24) Pt, which implies that Pr = Pt, but that implies r = t, a contradiction to the construction of K′. Therefore our initial assumption must have been false and there exists a morphism (cid:27) : C ! P1 over k of degree d such that (cid:12)(cid:12)(cid:12)( ) (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” (cid:12)(cid:12)(cid:12) = 2d (cid:0) 1: □ \ fP1; : : : ; P2d(cid:0)1g 54 Now we prove the desired inequality holds for cubic points on elliptic curves, excluding the special case when the support of the divisor consists of exactly five points forming a subgroup of the elliptic curve. Corollary 4.3.6. Let C be a nonsingular projective curve of genus 1 defined over a number field k and let R 2 C„k”. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct, such that the set fP1; : : : ; Png is not a subgroup of order five of the elliptic curve „C„k”; Pn” (which is i=1 Pi. Let " > 0. Then for all Q 2 C„k” automatically true if n , 5). Define the divisor D := with »k„Q” : k… = 3, ∑n (cid:12)(cid:12)(cid:12)} {(cid:12)(cid:12)(cid:12)( ) mD;S„Q” (cid:20) „N3„D” + "”hR„Q” + O„1” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” \ Supp„D” taken over all k-morphisms (cid:27) : C ! P1 where N3„D” := max of degree 3. If no such k-morphism exists, we say N3„D” = 0. Proof. First note that since degree 3 morphisms are at most 3-to-1, we have N3„D” (cid:20) 6. We also clearly have N3„D” (cid:20) deg„D”. By Riemann-Roch, the dimension of the complete linear system jEj of any divisor E on C of positive degree is given by dimjEj = deg„E” (cid:0) 1. If there are only finitely many points Q 2 C„k” with »k„Q” : k… = 3, then mD;S„Q” (cid:20) O„1”, so we assume there are infinitely many points Q 2 C„k” with »k„Q” : k… = 3. In particular, there exists a triple of conjugate k-cubic points Q1; Q2; Q3 2 C„k”. Since jQ1 + Q2 + Q3j = 2 and jQ1 + Q2 + Q3 (cid:0) T (cid:0) Uj = 0 for all T; U 2 C„k”, it follows that Q1 + Q2 + Q3 is a degree 3, very ample divisor defined over k, whose complete linear system jQ1 +Q2 +Q3j has dimension 2. Thus it induces an embedding F : C ! P2 defined over k, whose image is a curve over k of degree 3 such that the points F„Q1”; F„Q2”; F„Q3” lie on a line H over k. Projecting from any k-rational point A not on the curve F„C” in P2 produces a degree 3 morphism fA : C ! P1 over k. If deg„D” = 1, then choose any line L over k in P2 containing F„P1”. As the intersection of two lines over k, the point A := L \ H is k-rational, thus different from F„Q1”; F„Q2”; F„Q3”, thus not a point on F„C”. So fA is a degree 3 morphism over k and after composing with a k-automorphism of P1, we may assume fA„P1” = 0, so that N3„D” = 1. 55 If deg„D” = 2, then let L be the line in P2 containing F„P1” and F„P2”, thus a line over k. As the intersection of two lines over k, the point B := L \ H is k-rational, thus different from F„Q1”; F„Q2”; F„Q3”, thus not a point on F„C”. So fB is a morphism of degree 3 over k and after composing with a k-automorphism of P1, we may assume fB„P1” = fB„P2” = 0, so that N3„D” = 2. If deg„D” = 3, then after composing with a k-automorphism of P1, we may assume fB„P1” = fB„P2” = 0 and fB„P3” = 1, so that N3„D” = 3. If deg„D” = 4, then let H2 be the line in P2 containing F„P1”; F„P2”; F„P3”, and let gA : C ! P1 be the morphism of degree 3 over k induced by the projection from a point A 2 H2„k”. Then gA„P1” = gA„P2” = gA„P3”, so composing with a k-automorphism of P1, we may assume gA„P1” = gA„P2” = gA„P3” = 0 and gA„P4” = 1, so that N3„D” = 4. If deg„D” = 5, then by the hypothesis and by Proposition 4.3.4, we have N3„D” = 5. In these cases N3„D” = deg„D”, so we have mD;S„Q” (cid:20) hD„Q” by Definition 2.2.8 (cid:20) „deg„D” + "”hR„Q” + O„1” by Theorem 2.2.12 = „N3„D” + "”hR„Q” + O„1”: Now suppose deg„D” (cid:21) 6. If we assume that N3„D” (cid:20) 4, then by Proposition 4.3.4, we know that the points fP1; : : : ; P5g form a subgroup of order 5 of the elliptic curve C with identity element P5. In particular there exist i; j 2 f2; 3; 4g such that Pi + Pj = P1. Replacing P1 with P6 and applying Proposition 4.3.4 again to show that the points fP2; : : : ; P6g form a subgroup of order 5 in the elliptic curve C with identity element P5. But these points obey the same group law as before, and Pi + Pj = P1, meaning the set is not closed under addition, a contradiction. Therefore we conclude that N3„D” (cid:21) 5. If N3„D” = 6, then the inequality follows from Proposition 4.3.2. (cid:24) Conversely, if there exists a subset fi1; i2; i3; i4; i5; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 ” = Pi4 + Pi5 + Pi6, then there exists a k-morphism (cid:27) : C ! P1 of degree 3 such that (cid:27)„Pi1 ”. After composing with a k-automorphism of (cid:27)„Pi2 ” and (cid:27)„Pi4 ” = (cid:27)„Pi3 ” = (cid:27)„Pi5 ” = (cid:27)„Pi6 56 P1, we may assume (cid:27)„Pi1 N3„D” = 6. ” = (cid:27)„Pi2 ” = (cid:27)„Pi3 ” = 0 and (cid:27)„Pi4 ” = (cid:27)„Pi5 ” = (cid:27)„Pi6 ” = 1, so that Thus if N3„D” = 5, then such a subset of f1; : : : ; ng does not exist and the inequality again □ follows from Proposition 4.3.2. Proposition 4.3.7. Let C be a nonsingular projective curve of genus 2 defined over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := i=1 Pi. Let R 2 C„k” and let " > 0. ∑n If there exists a subset fi1; i2; i3; i4; i5; i6g (cid:26) f1; : : : ; ng such that Pi1 +Pi2 +Pi3 (cid:24) Pi4 +Pi5 +Pi6, then for all Q 2 C„k” with »k„Q” : k… = 3 mD;S„Q” (cid:20) „6 + "”hR„Q” + O„1”: If such a linear equivalence does not exist, but there exists a subset fi1; i2; i3; i4; i5g (cid:26) f1; : : : ; ng (cid:24) Pi4 + Pi5 +T, ; Pi3 such that Pi1 + Pi2 + Pi3 and a point T 2 C„k” distinct from the points Pi1 then for all Q 2 C„k” with »k„Q” : k… = 3 ; Pi2 mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1”: And if neither of these hold, then for all Q 2 C„k” with »k„Q” : k… = 3 mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1” Proof. By Theorem 3.1.6, there exists a proper closed subscheme Z (cid:26) Sym3„C” such that for all P 2 Sym3„C”„k” n Z, mD;S„P” (cid:20) „4 + "”hφ(cid:3)(cid:25)(cid:3) 1R„P”: Let Z j be an irreducible component of Z. Then (cid:22)jZ j is a k-morphism from the irreducible surface Z j to the abelian variety Jac„C”. If (cid:22)jZ j were a constant map then Z j would be contained in a fiber of (cid:22). But the fibers of (cid:22) are 1-dimensional by Riemann-Roch, a contradiction, so (cid:22)jZ j is nonconstant. 57 If the image of (cid:22)jZ j is a curve, then Theorem 4.3.1 says there exists a proper closed subscheme Yj of Z j such that for all P 2 Z j„k” n Yj mDjZi ;S„P” (cid:20) „3 + "”h„φ(cid:3)(cid:25)(cid:3) 1R”jZi „P” + O„1”: Since Yj is a proper closed subscheme we may assume it is 1-dimensional, and is thus a union of curves. By Lemma 4.2.1, if there does not exist a subset fi1; i2; i3; i4; i5g (cid:26) f1; : : : ; ng and a point T 2 C„k” distinct from the points Pi1 ; Pi2 ; Pi3 such that (cid:24) Pi4 + Pi5 + T; Pi1 + Pi2 + Pi3 then for all Q 2 C„k” with »k„Q” : k… = 3 and φ„ ¯Q” 2 Yj mD;S„Q” (cid:20) „4 + "”hR„Q” + O„1”: And if there does not exist a subset fi1; i2; i3; i4; i5; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 Pi4 + Pi5 + Pi6, then for all Q 2 C„k” with »k„Q” : k… = 3 and φ„ ¯Q” 2 Yj (cid:24) mD;S„Q” (cid:20) „5 + "”hR„Q” + O„1”; else ( ) mD;S„Q” (cid:20) „6 + "”hR„Q” + O„1”: Finally, if the image of (cid:22)jZ j is all of Jac„C”, then Yj = Supp is a finite union of irreducible curves Yjk on Z j. By Theorem 4.1.4, it is enough to show the inequality holds on these curves, which of course follows once again from Lemma 4.2.1. (cid:22)(cid:3) (cid:22)(cid:3)φ(cid:3)(cid:25)(cid:3) 1DjZ j Combining the above inequalities we get the result on all of Z j. Since there are only finitely many irreducible components Z j in Z, we make take the maximum of all of the above constants O„1” to get the result for all Q 2 C„k” with »k„Q” : k… = 3. □ Lemma 4.3.8. Let C be a nonsingular projective curve of genus 2. Let D be an effective divisor on C of degree 3. Then the complete linear system jDj has a base point if and only if K (cid:20) D for a canonical divisor K on C. 58 Proof. By Riemann-Roch, dimjDj = 1. Suppose D has a base point P 2 C„k”. Then there exists an effective divisor E of degree 2 such that D = E + P. Furthermore dimjEj = dimjD (cid:0) Pj = dimjDj = 1. By Riemann-Roch, dimjK (cid:0) Ej = 0, and since deg„K (cid:0) E” = 0, this implies E (cid:24) K. That is to say, E is a canonical divisor with E (cid:20) D, as was to be shown. Conversely, suppose K (cid:20) D for a canonical divisor K. Then D = K + P for some P 2 C„k”. □ Thus dimjD (cid:0) Pj = dimjKj = 1 = dimjDj, so P is a base point of D. {(cid:12)(cid:12)(cid:12)( ) (cid:12)(cid:12)(cid:12)} ∑n Corollary 4.3.9. Let C be a nonsingular projective curve of genus 2 defined over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor i=1 Pi. Suppose that either n (cid:20) 2 or else that there exist i1; i2; i3 2 f1; : : : ; ng such that the D := set fPi1 g does not contain a pair of hyperelliptic conjugates (which is automatically true ; Pi2 when n (cid:21) 5). Let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = 3, mD;S„Q” (cid:20) „N3„D” + "”hR„Q” + O„1” ; Pi3 \ Supp„D” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” taken over all finite k-morphisms (cid:27) : where N3„D” := max C ! P1 of degree 3. If no such k-morphism exists, we say N3„D” = 0. Proof. First note that since degree 3 morphisms are at most 3-to-1, we have N3„D” (cid:20) 6. We also clearly have N3„D” (cid:20) deg„D”. If there are only finitely many points Q 2 C„k” with »k„Q” : k… = 3, then mD;S„Q” (cid:20) O„1”, so we assume there are infinitely many points Q 2 C„k” with »k„Q” : k… = 3. In particular, there exists a triple of conjugate k-cubic points Q1; Q2; Q3 2 C„k”. By Riemann-Roch, dimjQ1 + Q2 + Q3j = 1. If the linear system jQ1 + Q2 + Q3j has a base point, then by Lemma 4.3.8, K (cid:20) Q1 + Q2 + Q3 for a canonical divisor K on C. Without loss of generality, say K = Q1 + Q2. By uniqueness, jKj is fixed under Galois action. Thus if (cid:24) is any Galois action that takes Q2 to Q3, then (cid:24)„Q1” = Qi for some i 2 f1; 2g, so Q1 + Q2 (cid:24) (cid:24)„Q1 + Q2” = Qi + Q3 and subtracting Qi from both sides leaves Q j (cid:24) Q3 for some j 2 f1; 2g, implying Q j = Q3, a contradiction. Therefore jQ1 + Q2 + Q3j is a 1-dimensional base point free linear system of degree 3 defined over k. It follows that the 59 corresponding k-morphism ! : C ! P1 has degree 3 with !„Q1” = !„Q2” = !„Q3”. Therefore if deg„D” = 1 or 2, then N3„D” = deg„D”. Alternatively, if there exist i1; i2; i3 2 f1; : : : ; ng such that fPi1 of hyperelliptic conjugates, then the complete linear system jPi1 + Pi2 + Pi3 by Lemma 4.3.8. So there exists a k-morphism (cid:13) : C ! P1 of degree 3 with (cid:13)„Pi1 (cid:13)„Pi3 ” = 0. If deg„D” = 3, then this shows N3„D” = 3. In these cases, where N3„D” = deg„D”, we have ; Pi2 ; Pi3 g does not contain a pair j is base point free ” = ” = (cid:13)„Pi2 mD;S„Q” (cid:20) hD„Q” by Definition 2.2.8 (cid:20) „deg„D” + "”hR„Q” + O„1” by Theorem 2.2.12 = „N3„D” + "”hR„Q” + O„1”: From here on, suppose deg„D” (cid:21) 4. Then there exists a fourth point in the support of D and the morphism (cid:13) must map it somewhere. After a k-automorphism of P1, we may assume it is sent to infinity, and so N3„D” (cid:21) 4. If N3„D” = 6, then the inequality follows from Proposition 4.3.7. (cid:24) Conversely, if there exists a subset fi1; i2; i3; i4; i5; i6g (cid:26) f1; : : : ; ng such that Pi1 + Pi2 + Pi3 ” = Pi4 + Pi5 + Pi6, then there exists a k-morphism (cid:27) : C ! P1 of degree 3 such that (cid:27)„Pi1 (cid:27)„Pi2 ” = (cid:27)„Pi3 Thus if N3„D” = 5, then such a subset of f1; : : : ; ng does not exist and the inequality again ” = 1, so that N3„D” = 6. ” = 0 and (cid:27)„Pi4 ” = (cid:27)„Pi6 ” = (cid:27)„Pi5 follows from Proposition 4.3.7. ” = (cid:14)„Pi2 ” = (cid:14)„Pi3 If N3„D” (cid:21) 5 then there exists a subset fi1; i2; i3; i4; i5g (cid:26) f1; : : : ; ng and a k-morphism (cid:14) : C ! P1 of degree 3 such that (cid:14)„Pi1 ” = 1. Since 0 (cid:24) 1 in P1, it follows that (cid:14)(cid:3)0 (cid:24) (cid:14)(cid:3)1. That is Pi1 + Pi2 + Pi3 (cid:24) Pi4 + Pi5 + T for some T 2 C„k”. Since (cid:14)„T” = 1, it is clear that T is distinct from the points Pi1 ; Pi3. Since (cid:14) is defined over k, for any Galois action (cid:24) 2 G„k(cid:157)k” we have (cid:24)„(cid:14)(cid:3)„1”” = (cid:14)(cid:3)„(cid:24)„1”” = (cid:14)(cid:3)„1” so Pi4 + Pi5 + (cid:24)„T” = (cid:24)„Pi4 + Pi5 + T” = Pi4 + Pi5 + T, implying (cid:24)„T” = T. Since (cid:24) was arbitrary, we have T 2 C„k”. Ergo if such a subset and point do not exist, then N3„D” < 5, so N3„D” = 4 and ” = 0 and (cid:14)„Pi4 ” = (cid:14)„Pi5 ; Pi2 60 the result follows from Proposition 4.3.7. □ Corollary 4.3.10. Let C be a nonsingular projective curve of genus 2 defined over a number field k. Let S be a finite set of places of k. Let P1; : : : ; Pn 2 C„k” be distinct and define the divisor D := ∑n i=1 Pi. Let " > 0. Then for all Q 2 C„k” with »k„Q” : k… = 3, mD;S„Q” (cid:20) „ ˜N3„D” + "”hR„Q” + O„1” \ Supp„D” (cid:27)(cid:0)1„0” [ (cid:27)(cid:0)1„1” {(cid:12)(cid:12)(cid:12)( ) (cid:12)(cid:12)(cid:12)} where ˜N3„D” = N3„C; k; D” := max phisms (cid:27) : C ! P1 of degree 3. If no such morphism exists, we say ˜N3„D” = 0. Proof. note that N3„D” (cid:20) ˜N3„D”, so the inequality is implied by Corollary 4.3.9 except in cases where deg„D” = 3 or 4. taken over all finite mor- Consider the morphism »(cid:1); (cid:1)… : C2 ! Jac„C” defined by »T; U… = O„U (cid:0) T”. If »T; U… = »T′; U′…, then T + U′ (cid:24) T′ + U, and since the hyperelliptic morphism is unique, it follows that »(cid:1); (cid:1)… is 2-to-1, and comparing dimensions, we see it is surjective. Thus if deg„D” = 4, there exist T; U 2 C„k” such that Pi1 + Pi2 + T (cid:24) Pi3 + Pi4 + U: We can always arrange the points so that Pi1 and Pi2 are not hyperelliptic conjugates and Pi3 and Pi4 are not hyperelliptic conjugates, so this defines a base point free linear system of dimension 1 and degree 3. Thus there is a morphism (cid:27) : C ! P1 of degree 3 with (cid:27)„Pi1 ” = 0 and (cid:27)„Pi3 ” = (cid:27)„Pi4 Similarly if deg„D” = 3, then for any V 2 C„k” there exist T; U 2 C„k” such that ” = 1, that is ˜N3„D” = 4. ” = (cid:27)„Pi2 Pi1 + Pi2 + T (cid:24) Pi3 + V + U: We can always arrange the points so that Pi1 and Pi2 are not hyperelliptic conjugates and choose V so that V is not the hyperelliptic conjugate of Pi3 and V is not equal to Pi1 or Pi2, so that this defines a base point free linear system of dimension 1 and degree 3. Thus there is a morphism (cid:27) : C ! P1 of degree 3 with (cid:27)„Pi1 ” = 1, that is ˜N3„D” = 3. ” = 0 and (cid:27)„Pi3 ” = (cid:27)„Pi2 61 In these cases, where ˜N3„D” = deg„D”, we have mD;S„Q” (cid:20) hD„Q” by Definition 2.2.8 (cid:20) „deg„D” + "”hR„Q” + O„1” by Theorem 2.2.12 = „ ˜N3„D” + "”hR„Q” + O„1”: □ 62 BIBLIOGRAPHY 63 BIBLIOGRAPHY [1] Enrico Bombieri and Walter Gubler. Heights in Diophantine geometry, volume 4 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2006. [2] P. Corvaja and U. Zannier. On integral points on surfaces. Ann. of Math. (2), 160(2):705–726, 2004. [3] Pietro Corvaja and Umberto Zannier. On a general Thue’s equation. Amer. J. Math., 126(5):1033–1055, 2004. [4] Jan-Hendrik Evertse and Roberto G. Ferretti. A generalization of the Subspace Theorem with polynomials of higher degree. In Diophantine approximation, volume 16 of Dev. Math., pages 175–198. SpringerWienNewYork, Vienna, 2008. [5] Gerd Faltings. Diophantine approximation on abelian varieties. Ann. of Math. (2), 133(3):549– 576, 1991. Joe Harris and Joe Silverman. Bielliptic curves and symmetric products. Proc. Amer. Math. Soc., 112(2):347–356, 1991. [6] [7] Robin Hartshorne. Algebraic geometry. Springer-Verlag, New York-Heidelberg, 1977. Grad- uate Texts in Mathematics, No. 52. [8] Marc Hindry and Joseph H. Silverman. Diophantine geometry, volume 201 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000. An introduction. [9] Aaron Levin. On the Schmidt subspace theorem for algebraic points. Duke Math. J., 163(15):2841–2885, 2014. [10] Aaron Levin. Wirsing-type inequalities. Bull. Inst. Math. Acad. Sin. (N.S.), 9(4):685–710, 2014. [11] Jürgen Neukirch. Algebraic number theory, volume 322 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1999. Translated from the 1992 German original and with a note by Norbert Schappacher, With a foreword by G. Harder. [12] Min Ru and Pit-Mann Wong. planes in general positiong. Invent. Math., 106(1):195–216, 1991. Integral points of Pn (cid:0) f2n + 1 hyper- [13] Joseph H. Silverman. Arithmetic distance functions and height functions in Diophantine geometry. Math. Ann., 279(2):193–216, 1987. [14] Xiangjun Song and Thomas J. Tucker. Dirichlet’s theorem, Vojta’s inequality, and Vojta’s conjecture. Compositio Math., 116(2):219–238, 1999. [15] The Stacks Project Authors. Stacks Project. http://stacks.math.columbia.edu, 2018. 64 [16] Paul Vojta. Diophantine approximations and value distribution theory, volume 1239 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 1987. [17] Paul Vojta. A generalization of theorems of Faltings and Thue-Siegel-Roth-Wirsing. J. Amer. Math. Soc., 5(4):763–804, 1992. 65