Download as pdf
Download as pdf
You are on page 1of 224
ELEMENTS OF THE THEORY OF ALGEBRAIC CURVES A. SEIDENBERG University of California, Berkeley, California ADDISON-WESLEY PUBLISHING COMPANY Reading, Massachusetts - Menlo Park, California - London - Don Mills, Ontario This book is in the ADDISON-WESLEY SERIES IN MATHEMATICS Consulting Editor: Lynn H. Loomis Copyright © 1968 by Addison-Wesley Publishing Company, Inc. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or trans- mitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of the publisher. Printed in the United States of America. Published simultaneously in Canada, Library of Congress Catalog Card No. 68-19345. PREFACE A large part of the present book has been offered as a graduate course several times, and, moreover, with a preliminary presentation of the ring and field theory necessary for complete proofs. Several reasons, however, have led us to abandon, as part of the course, this preliminary material, to enter directly into the geometry, and to bring the subject within the scope of an undergraduate audience. First, there are separate courses for the algebra, and, although we have little confidence in “prerequisites,” the curriculum can ill bear the duplica- tion. Moreover, we find that a real understanding of the ring and field theory is not necessary for the geometry, but that a clear statement of the basic facts is sufficient. The student accepts that the complex number field is algebraically closed; in the same way, he may accept that every field has an algebraic closure. Except, then, for some of the subtleties of the algebra, the geometry becomes quite accessible. Second, there is a tendency to reserve the basic graduate course in Algebraic Geometry for the general theory of algebraic varieties. This, together with the other basic algebra, commutative and otherwise, consti- tutes a heavy load for the first years of graduate study; and at the same time there is a deficiency in Algebra in the undergraduate program (relative to Analysis, say). We think that Algebraic Curves is no harder than the much more widely offered undergraduate course in Differential Geometry, and can compete with it in usefulness and interest. The typical student we have in mind, then, at least for the first half of the book, is a senior mathematics major who has taken a course in algebra. We would hope that he has taken a course in Projective Geometry and knows what homogeneous coordinates are, but we have supplied a chapter on this. Probably he knows what a ring is, but we have tried to put down the definitions of all terms, except those he has encountered in the basic calculus sequence. Sometimes this is done in the notes, but often in the text with the euphemism “Recall that.” (The notes are signaled by a superior iii iv PREFACE dagger.) If a term is not known, the index should be first consulted. We refer the student to G. Birkhoff and S. MacLane, A Survey of Modern Algebra, for definitions we may have inadvertently omitted. ‘As for the organization of the book, in the first eighteen chapters, except for preliminary remarks, we work over the field C of complex numbers. The arguments hold word for word for arbitrary algebraically closed fields K of characteristic zero; but we do not have to make constant reference to the base field, or bother the student with the essentially spurious generality. The restriction to characteristic zero is made initially to avoid secondary issues; but the student will also want the results for arbitrary characteristic. The modifications needed for characteristic not equal to 0 are given in Chapter 19. In this way, the student will not have to unlearn, or relearn, the subject. Two equivalent definitions, i(I', A; P) and j(T, A; P), are given for the intersection multiplicity of two curves I’, A at a point P. The definition of i is given by means of resultants. It is an awkward definition for several reasons, but has the advantage that it can be explained easily and gives some of the basic projective properties of intersection multiplicity without excessive difficulty. It is, however, definitely unsuited for a birational study. One goes over to a definition, (I, A; P), by means of branches. To avoid making the equivalence i = j central in our exposition, we show in Chapters 12 and 14 that i and j share certain basic properties. The proof of the equivalence is postponed to Chapter 21, but logically is not really needed even then. The branches of a curve at a point can be obtained quite rapidly, in the case of characteristic zero, by using Hensel’s Lemma, as van der Waerden does in his Einfiihrung in die Algebraische Geometrie. Our method in Chapter 12 is not so short; but one gets the branches as a side result of an analysis of a point by means of locally quadratic transformations—an analysis that in no way involves power-series (as Hensel’s Lemma does). It also, inci- dentally, in no way uses the characteristic. This method is not really new, but it does not occur in the textbooks. Chapters 9 and 10 deal with a special topic, namely, cubies. They introduce several general ideas and provide illustrative material, but are not strictly necessary for the later developments. Originally we learned the subject from lectures by Professor Oscar Zariski. Since then we have studied S. Lefschetz’s Algebraic Geometry, B. L. van der Waerden’s Hinfiihrung, F. Severi’s Vorlesungen ther Algebraische Geometrie (not to mention works on algebraic functions of one variable, a closely related subject); and, more recently, have consulted R. J. Walker’s Algebraic Curves, and J. G. Semple and G. T. Kneebone’s Introduction to Algebraic Curves. There is no reason why the student cannot learn the subject, PREFACE Vv and very well indeed, from these books. We can only hope that here and there there may be an especially attractive element in our presenta- tion. An invitation by the University of Mexico in 1962 to deliver some lectures on a not too advanced level was the occasion for my deciding to write the present book. Professor E. Lluis wrote up my lectures and my wife trans- lated them from Spanish into English for me. I thank them both for helping me to get started. The visit to the University of Mexico was sponsored by the State Depart- ment upon recommendation by the Conference Board of Associated Research Councils, Committee on International Exchange of Persons. Professor Jack Ohm read the manuscript, made helpful comments, and pointed out slips and unfelicitous passages. I wish to express my gratitude to him. Berkeley, California AS. May 1968 CONTENTS List of Symbols Introductory Remarks . Homogeneous Coordinates Projective planes The line at infinity Projective planes over an extension field of K Algebraic Curves The equations of a curve Intersections Affine algebraic curves Resultants . Linear Transformations Coordinate systems ‘Transformations in the affine plane Simple Points and Singular Points . The polar . Some computational rules Bezout’s Theorem The Basic Inequality Cubies Rational functions on an irreducible curve Abelian varieties vi 21 27 29 30 31 38 43 48 55 60 10 n 12 iu 15, 16 17 18 19 21 22 23 CONTENTS Cubics (Continued) Points of inflection Cubics through eight points ‘The modulus of a nonsingular cubic Expansion at a Simple Point Formal power series Branches Imprimitive branch representations Generic Points Rational mappings Places Zeros and Poles . Differentials ‘The genus Infinitely near points . Noetherian Conditions Linear Series Linear Systems The Theorem of Riemann-Roch Extension and reformulation of the theorem of Riemann-Roch The Question of the Characteristic Linear Series and Rational Mappings; Space Curves Analytic Branches Noetherian Condi ns and Intersection Multiplicity Specialization; Space Curves . Specializations and residue maps (“places”) . Valuations and morphisms (“places”) Infinitely Near Points Notes Index vii 67 67 13 15 82 83 89 101 103 105 ll 113 115 118 127 129 135 143 147 149 152 160 173 186 193 196 199 2038 209 213 LIST OF SYMBOLS Page 104 110 113 113 116 118 129 129 133 135 135 140 141 Symbol AX, Y]: ring of polynomials in two letters C: complex number field ¢: is not an element of ANK: affine number plane over K K —0: set of elements of K with the exception of 0 (a + 2 tay): the class {(tirg, try, taq)} PNK: projective number plane over K U: set-theoretic union €: is not contained in = 0: is divisible by G.C.D.: greatest common divisor C(Xo, X,): field of quotients f/g with f, g € OLX, X,] R(f, g): resultant of f and g |A|: determinant of the matrix A L4: linear transformation of matrix A iT, A; P): intersection multiplicity of P and A at P min {a, 6}: minimum of the numbers a, 6 S x T: Cartesian product of § and 7 K[[X, Y]]: ring of formal power series in two letters K((X, Y)): field of quotients f/g with f, g ¢ K[[X, Y]] j(A, T; P): j-multiplicity of A and I at P K(a,.. .,a,): field adjunction of a,. . ., 0, to K (P; (& 9): model of curve I’ with generic point (£, 7) = — C: set of elements in E not in [2 : O(€)]: degree of © over C(é) dO: differential of 0 B-dl,+2Z-dl, +--+: set of linear combinations of dl, df... . with coefficients in © TA A: intersection of T and A (F, @): ideal generated by F, G A-T: divisor cut out on T by A (E)os (Elo Zeros, poles of & A = B: equivalence of divisors A, B gy: a linear series of effective divisors |C|: linear series of effective divisors equivalent to C viii Chapter 1 INTRODUCTORY REMARKS Algebraic Geometry is a study in which algebraic and geometric consider- ations come together. It has two aspects: on the one hand one studies geometrical problems by means of algebra, and on the other, algebraic prob- lems by means of geometry. Thus let us consider the assertion that in the plane a circle is cut by a straight line in at most two points. This is a purely geometrical assertion, but it can be established algebraically: The circle is given by an equation (X —a)?+(¥Y —6)?= 7%, and the line, by an equation uX +v¥ + w=0, with «4 0orv 40; supposing that v 4 0, we may rewrite the equation as ¥Y = mX + c. Now the problem becomes to solve simultaneously (X —aP + (Y¥ — bP =r, Y=mX +e. Eliminating Y, one obtains an equation in X. This equation may have no (real) roots and can have at most two. Each root will yield exactly one intersection of the circle and line. Thus the two curves intersect in at most two points. Here we have studied a geometrical problem algebraically. On the other hand, let us consider the circle X* + Y? —1= 0. Associated with the circle is the polynomial X? + Y®— 1 in thering &{X, Y] of polynomials in two indeterminates, or letters, over (i.e., with coefficients in) the field & of real numbers.t Here we may ask a purely algebraic question: Is the poly- nomial X? + Y?— 1 irreducible in AX, Y]? To answer this question, let us suppose that X?-+ Y2—1= F(X, Y)-G(X, Y), F, @ in &[X, Y]. Since the degree of FG is the sum of the degrees of F and of G, we obtain deg F + deg @ = 2. If deg F = 0, then F is in k, F has a reciprocal in k{X, Y], and we do not consider that F - @ is really a factorization in KX, Y] (since such factorizations can be produced at will). Similarly, we exclude the possibility that deg @ = 0. This leads to deg F = 1, deg @ = 1. 1 2 INTRODUCTORY REMARKS But a polynomial of degree one has a line as locus. The conclusion is that if X?4 Y?— 1 factors in kX, Y], then the circle X? + Y2?—1=0 is made up of two lines. Since every line has at least three points on it, and since a line and a circle have at most two points in common, this cannot happen, and hence X? + Y¥? — 1 is irreducible. These two examples are typical. In the first we answered a geometrical question algebraically; in the second, an algebraic question geometrically. By an algebraic curve (in the plane) we mean the locus determined by an equation F(X, Y) = 0. Our initial point of view is that F has coefficients in k, the field of real numbers, and by a point (z, y) we mean a pair of numbers in k. However, here we at once meet a difficulty. Let us suppose that we wish to study X? + Y?+4 1 in the same way that previously we studied X24 Y2—1. Presumably we wish to do this by considering the locus of X?4 Y241=0. However, this locus has no points. Naturally, we can obtain little information about X? + Y? + 1 if we start from the empty set. Hence we will change our point of view: we will allow the coefficients of F to be complex numbers, and by a point we will mean an ordered pair of numbers (x, y), x, y in the field C of complex numbers. By a Line (or complex line) we will mean the locus determined by an equation uX + vY + w = 0; here u, v, w may be any complex numbers with u # 0 or v # 0. This change in point of view has many repercussions. On one hand, we cannot be guided so easily by our geometrical intuition, since this has been formed relative to k, rather than to C. However, with some precaution, our intuition retains its usefulness. Another change is that the questions we ask are different. Thus, previously we asked whether X? + Y?—1 was irreducible in kX, Y];now we will ask whether it is irreducible in C[X, Y]. To answer this question, we can make use of our previous result, but first we need a lemma. Lemma 1.1. The real points of a complex line are on a real line. Proof. Let wX + v¥ +w=0 be the given line; let w= a + ity, v= 01 + ivy, and w = w, + ivy, with 4, 04, Wy, Up, 2, and w, in k and i = V—1. Weask for the real numbers (2, y) such that ux + vy + w = 0. We obtain the conditions uz + ny + w, = 0, Ugt + gy + Wz = 0. As u #0 or v £0, at least one of ty, Uy, %, ¥2 is # 0; and (a, y) is on uyX + ¥ + w, = 0 and on u,X + v,¥ + w, = 0, of which equations at least one is that of a real line. Now we have the following theorem. Theorem 1.2. X? + Y* — 1 is irreducible in C[X, Y]. INTRODUCTORY REMARKS 3 Proof. If X* + Y* — 1 were reducible, then reasoning as before, one would have a decomposition into two factors X24 Y2—1 = (wX + oF + wywX +o ¥ tw). Hence the (complex) locus of X* + Y2 — 1 = 0 would be made up of two (complex) lines. But then the real points of X* + Y? — 1 = 0 would lie on two real lines, and this is not: so. Theorem 1.3. X® + Y* + 1 is reducible in C[X, Y]. Proof. Let us consider the transformation t: (x, y) — (ia, ty). This transformation is univalent (that is, distinct points are transformed into distinct points) and is onto the plane (that is, every point of the plane is an image of some point under the transformation ¢); moreover it transforms lines into lines and X?+ Y¥2+4+1=0 into X?4+ ¥2—1=0 [the equations of ¢ are X’ = iX, Y’ = iY, from which the transformed locus is obtained as (X’/i)? + (Y’/i)® + 1 = 0 and we omit the primes]. Thus if X? 4 Y?+ 1 factors, then also X? + Y¥? — 1 factors, and we know that this is not so. We have changed our point of view, but someone else may say that he is still interested in the field & of real numbers: he wants answers to his original questions, for example, is X? + Y?+ 1 irreducible in A[X, Y]? This is easy to answer: X? + Y2 + 1 is irreducible in C[X, Y] and hence, a fortiori, is irreducible in M{X, Y]. Let us observe, however, that factorization in C[X, Y] is not always the same as in k[X, Y]. For example, X* + Y? is irreducible in kX, Y] but is reducible in C[X, Y]; in fact, X* + Y2 = (X + i¥\(X —iY). This example shows that we ought to be careful in using our geometric intuition, that is, technically speaking, in passing from k[X, Y] to C[X, Y]. Thus X? + Y2 and X?— Y® are essentially different in &[X, Y] (one factors but the other does not), but are essentially the same in C[X, Y] (both factor into distinct linear factors; moreover, the transformation X’ = X, Y’ = i¥ takes one into the other). Initially we passed from & to C' so that the locus of X? + Y2+1=0 would not be empty. This leads us to our first general theorem. Theorem 1.4. On every (plane, algebraic) curve there are infinitely many points. [Let us repeat, but with more precision: by a plane algebraic curve we mean the set of points (#, y) with x, y in C that satisfy an equation F(X, Y) = 0, where F is in C[X, Y] but not in C (in symbols: F ¢ C). We insist that F¢C. If F = ce, the locus of F = 0 contains no point 4 INTRODUCTORY REMARKS if c 4 0 and contains every point if c = 0; we do not wish to call either of these loci a curve.] Proof. Let the curve be given by F(X, Y) = 0. As F¢0, either X or ¥ (or both) occur in F, let us say Y occurs. We write F as a polynomial in Y with coefficients in OLX]: F = a(X)¥™ 4+ a(X)¥"™1 4 +++ + an(X), where a(X)EO[X], i=0,...,m, and aX) #0. Let deg a,(X) = n. Then a,(X) has at most n roots. Let a € C be such that a,(a) # 0; there are infinitely many ways to choose a. Now consider the equation ag(a)¥™ + ay(a)¥" + +++ + a,(a) = 0. Here m > 0. Now we use the basic fact about C, namely, that every poly- nomial in one letter with coefficients in C of positive degree has at least one zero in C. Hence there exists a number 6 € C such that a(a)b" + a(a)b"-? + +++ + ag(a) = 0. Hence the point (a, 6) is on F(X, Y) = 0. Since there are infinitely many possibilities for a, one obtains infinitely many points (a, b) on F(X, Y) = 0. The same argument shows that there are infinitely many points not on F(X, Y)=0. Let us take a€C as above. Now we use the fact that a,(a)¥™ +-+-+++ a,,(a) = 0 has at most m roots to obtain a 6 in C such that a,(a)b™ + +++ + a,(a) 4 0. Our next question is: Is it true that every pair of curves has a point in common? Clearly no, as parallel lines (for example, X = 1 and X = 2) do not meet. We know how to remedy this difficulty: we pass to the pro- jective plane. In it we define an algebraic curve as the locus determined by an equation F(X, Y, Z)=0 with F a homogeneous polynomial in CLX, Y, Z], F¢C. The question now is: Is it true that every curve in the projective plane over the field C of complex numbers has with every other such curve at least one point in common? Here we have still another change in point of view: we pass from the affine plane to the projective plane. It is not necessary to do this; one can study curves in the affine plane, and one will find a criterion (more or less complicated) for two curves to meet. However, it is very convenient to pass to the projective plane, where we will find simpler formulations for many of our theorems. Remark. We said above that we could learn little about X? + Y? +41 from the locus of real points of X? + Y® + 1 = 0, since this locus is empty. INTRODUCTORY REMARKS 5. Actually, we can prove that X® + Y# + 1 is irreducible in H{X, Y] already from this fact: if X? + Y% + 1 factored, it would factor into two linear factors, the locus X? + ¥2 + 1 = 0 would consist of two straight lines, and such a locus cannot be empty. A better example is X* + Y4 +1 = 0. A priori, this could factor in k[X, Y] into a linear factor and a factor of degree three, or into two factors of degree two. The first possibility is excluded by the argument just given, but the second cannot be similarly excluded. Later, however, we shall define the notion of a singular point of a curve, and we shall see that X* + Y4 + 1 is irreducible in C[X, Y] be- cause the curve X* + ¥4 + 1 = 0 has no singular points (in the projective plane over C), while any reducible curve does have singularities. A fortiori X44 Y4 + 1 does not factor in k[X, Y]. Exercises* 1. Show that any line meets the curve Y? — X? — X? = 0 in at most three points. (Consider separately lines of the form ¥ — mX — b = 0 and lines of the form X — a = 0.) 2. Does ¥2 — X* — X° factor (properly) in C[X, ¥]? If not, explain. 3. Consider the curve ¥* — X* — X* = 0 asa real locus. Sketch the curve, especially near the origin. 4, Do Exercises 1, 2, and 8 for ¥? — X° = 0. 5. Let P(X, ¥)= X?4 ¥?+aqX+0F +e, PAX, ¥) = X?4+ ¥* + a,X + b¥ + cy with P, # P,. The locus of points (z, y) in the complex affine plane such that Pix, y) = Paw, y) is a straight line through the intersections of the curves P,(X, Y) = 0 and P(X, ¥)=0. Uf P,(X, ¥) = 0 is a circle in the Euclidean plane and (a, y) is a real point, then P,(x, y) is called the power of (x, y) with respect to the circle. It has a simple geometrical interpretation.] * Declarative sentences in the exercises (as, for example, the second sentence in Exercise 5) are assertions to be established. Chapter 2 HOMOGENEOUS COORDINATES In this chapter we recall some of the basic facts about affine and projective planes. We start with the Euclidean plane and take in it a coordinate system. Each point, then, is associated with an ordered pair of numbers (2, y), its coordinates; every ordered pair (x, y) gives the coordinates of some point. Every line has a linear equation aX + 6Y +c = 0 with a 40 orb 40, i.e., a point (2, y) ison the line if and only if ax + by +c = 0; the equation is essentially unique, ie., if a’X + 6'Y + c’ = 0 is another equation for the same line, then there is a number A such that a’ = Ja, b' = 4b, o’ = Je. Conversely, every linear equation aX + bY +c = 0 witha #0 orb 40 is the equation of some line. In the Buclidean plane, the coordinates x, y are real numbers, and one has the notion of the distance between two points. If one wishes to disregard this notion, one speaks of the affine plane (over the real-number field k). But then, more generally, one can start from an arbitrary field K, define a point as a pair (x, ) of elements from K and a line as the set of points satisfy- ing a linear equation aX + bY +c = 0 witha 4 0 or b ¥ 0, and, on the whole, retain the geometry of points and lines of the real affine plane. For example, we shall still have the following theorems: I) On any two (distinct) points there is one and only one line. II) (The Parallel Postulate) For any line land any point P not on 1, there is one and only one line on P that does not meet l. The proofs of these theorems are algebraic arguments which are familiar for k, but which work equally well for any field K. The object we have just described is called the affine (number) plane over K and is denoted ANK. Let (x, y) be a point in ANK. By homogeneous coordinates for (x, y) we mean the triple (1, x, y) or any multiple (¢, tx, ty) thereof provided that 6 PROJECTIVE PLANES 7 t40(tEK). Any triple (x, 2, 23) with x £ 0 is a set of homogeneous coordinates for just one point, namely (4/29, 23/%9). The relations between the original, or nonhomogencous coordinates, and the homogeneous coordinates are ea mt, y= 2)%— Hence if bX + cY¥ + a =0 is a line, the condition on the homogeneous coordinates (#9, 2%, v3) of a point P to lie on the line is 5(a/%0) + ¢(x4/%) + a = 0, or since x» # 0, bx, + cx, + ax) = 0, 80 that the homogencous equation for the line is bX, + cX, + aX, =0. A linear equation a)Xp + a,X, + a,X_, = 0 can be regarded either from a homogeneous point of view or from a nonhomogeneous point of view. Regarding it in the first way, we do not count (0, 0, 0) as a solution, or (for emphasis) we call it the érivial solution, where the word trivial is being used in a technical sense. If the triple (xo, 21, 23) satisfies the equation, then 80 does (to, tay, ta) for any t # 0 (in K); if xq, 2, 2» are not all zero, one calls the whole class (tz, t,, ta), where ¢ runs over K — 0, a (single) solution of the (homogeneous) equation. [Notation. (ay : a : 23) stands for this class.] Assuming, then, that a, # 0 or a, 0, one sees that there is a one-to-one correspondence between the solutions of ay + a,X + a,Y = 0 and those of aXq + aX, + a,X_ = 0 in which x #0. Considering now only triples (x9, %, %_) in which 2 = 0, one observes that (0, —dg, a,) satisfies the equa- tion, hence so does (0, —tag, ta;); and one proves that no other triple satisfies the equation. Thus there is one and only one solution to aoXp + 4X, + aX, = 0 in which 2 = 0, namely {(0, —tay, ta,)}. PROJECTIVE PLANES So far we have spoken only of the affine plane. By a point of the projective plane over K one means a class of triples (tx, tx,, tz), where 29, 2, 2, are in K, but are not all zero, and truns over K — 0. The triple (x9, 2, x2), abbrevi- ated 2, is called a representative of the point; fort # 0, (tz, tz, tz), abbrevi- ated tx, is a representative of the same point. The point with representative x is denoted P,; we sometimes also refer, though loosely, to the point x. The triple (x, x, 22) is also called a set of coordinates for the point Py. By a line of the projective plane over K one means the set of points P, such that gly + ayt, + agx, = 0, where ay, a,, a, are in K and are not all zero. The equation aX) + 4,X, + a,X, = 0 is called an equation of the line; ta,X, + ta,X, + ta,X, = 0, for any t in K — 0, is also an equation of the same line, and one proves that there are no others. Therefore one speaks, 8 HOMOGENEOUS COORDINATES though with a slight inexactitude, of ajX, + a,X, + a,X, = 0 as being the equation of the line. The triple of coefficients (a, 44, dy) = a is called a representative of the line; then (tap, ta,, ta,), 1 € K — 0, is also a representative of the same line, and there are no other representatives for the line. Notation- ally, we refer to the line L,. The point P, is said to be on the line L, if P, is one of the points of the line L,, or, in other words, if gy + Ay, + Agr, = 0. The notion made up of the set of points, the set of lines, and the relation on just defined, is called the projective (number) plane over K and is denoted PNK. It is usual to write the triples which represent lines as row triples and those which represent points as column triples; for typographical reasons, however, we will also write the point representatives in a row if no confusion is anticipated. THE LINE AT INFINITY It is clear that there must be a relation between the affine and projective planes. Start with the projective plane PNK and subtract the line X, = 0: every point not on Xj = 0 has a set of coordinates of the form (1, «, y). Then (1, x, y) — (2, y) is clearly a one-to-one correspondence between the points of PNK not on Xy = 0 and the points of ANK. Moreover the points of the projective line ayX) + a,X, + a,X, = 0, with a, #0 or a, #0, make up, with just one exception (namely, the point (0, —a,, a,)], the points of the affine line a)-+a,X-+a,¥ =0. Thus the correspondence (1, 2, y) — (2, y) induces a one-to-one correspondence between the lines of PNK, with the exception of Xj = 0, and the lines of ANK, the line ayXq + a,X, + a,X, = 0 corresponding to the line ay + a,X + 4,¥ = 0. Because of these correspondences, one says that the plane ANK is part of the plane PNK; one also says that ajXp + aX, + a,X_ = 0 (witha, £0 or a, # 0) and ay + a,X + a,¥ = 0 are the same line. Of course, one can start with ANK and add on the missing line X, = 0 to get the projective plane. From the point of view of the projective plane, any line is very much like any other: the line X, = 0, for example, is just as good, or bad, as the line Xj = 0. But from the point of view of the affine plane obtained by removing X, == 0, the line X, = 0 is special: it is then referred to as the line at infinity. Points on Xj = O are referred to as the points at infinity, while the points in ANK itself are said to be at finite distance (though no concept of distance is involved). From the projective plane one can subtract any of the lines X) = 0, X, = 0, or X, = 0 to obtain an affine plane. In this way three different PROJECTIVE PLANES OVER AN EXTENSION FIELD OF K 9 X1=0:X=0, Fo: 1,0, 0) Ei: (0, 1, 0) Fig. 2.1 (though isomorphic) ANK’s are obtained. (Two affine planes are said to be isomorphic if the points and lines of one plane can be put in one-to-one correspondence with the respective points and lines of the other in such way that a point and a line of one plane are incident if and only if the correspond- ing point and line of the other plane are incident.) The projective plane can be thought of as these three affine planes appropriately patched together. To have a diagram of the projective plane in relation to the affine plane, one draws a triangle as in Fig. 2.1. The line at infinity Xj = 0 is brought nearer, where one can see it. The X-axis, X,—=0 or Y =0, is drawn horizontally; the Y-axis, X; = 0 or X = 0, is drawn nearly, but not quite vertically, since angle, and in particular right angle, is not a concept of affine or of projective geometry; moreover the events taking place near Hy are just like those taking place near H,, or Hy, or, indeed, any other point. The point By: (1, 0, 0) is the origin (0, 0) of the affine plane. The point B,: (0, 1, 0) is the point at infinity on the X-ais, and By: (0, 0, 1) is the point at infinity on the Y-axis. Lines through Ez are called vertical, those through Ey, horizontal. The parallel but nonvertical lines Y = mX + 6, and Y = mX + b, mect at infinity in the point (0, 1, m). PROJECTIVE PLANES OVER AN EXTENSION FIELD OF K Let L be an extension field of K, ie., a field containing K. (Consider, for example, K = real field, L = complex field.) Then, obviously, ANK is contained in ANL; in symbols, ANK ¢ ANL. If in PNL we con- sider the points and lines having coordinates in K, then these points and lines form a plane isomorphic to PNK, and in this sense one can say PNK c PNL. Note, however, that a point in PNK may have coordinates not in K; for example, P :(1, 1, 1) is in PNK, but P also has coordinates (t,t) for any t¢ L — 0, and ¢ may not be in K. Exercise. By normalized coordinates for a point in a PNK one means coordinates in which at least one entry is 1. Every point has at least one and at most three sets of normalized coordinates. If K ¢ L, then a point in PNL is in PNK if and only if its normalized coordinates are in K. Chapter 3 ALGEBRAIC CURVES* A point of the projective plane is represented by a triple (x9, 11, %,), some x, % 0, and equally well by (pz, px, pz) for any p # 0: the point is the collection of triples (px, px,, pt,). If F(X, X;, Xz) €C[Xp, X, Xo] is a polynomial that vanishes at (29, 2%, 2), then we shall want F also to vanish at (py, px, px,); otherwise we could not speak of F vanishing or not vanish- ing at the point represented by (29, 2, 22). For this it will be sufficient (and, in fact, also necessary) to assume that F is homogeneous, i.e., that all its terms have the same total degree. Then let F(Xo, X,, Xp) € C[Xo, Xy, Xz] be a nonconstant homogeneous polynomial. The set of points (zz, 21, 2) such that F(z», x1, 2») = 0 is called an algebraic curve (in the projective plane). Now we ask the following question: Does every pair of algebraic curves (in the projective plane) have points in common? FIRST CASE: One of the curves is a line. One then has TP: F(X», X, X;) =0 (n = degree of F), D:agXq + a,X, + a,X,= 0 (some a; + 0). We may suppose that a, # 0. To solve simultaneously, we eliminate X», getting the equation F(X, X;, —(ao/as)Xq — (a,/a,)X,) = 0. Consider the polynomial PF (%. eee x) = H(Xy, X)). a, ay Two cases can occur: either H is the polynomial zero or H is different from zero. In the first case, every point of the line is a point of I, ie. L 0 for every element u, we find that 12 ALGEBRAIC CURVES deg e = 0, ie., ee K (and ¢ #0). Conversely, if e¢ K and e #0, then since K is a field, there is an f € K such that ef = 1. Thus: 3.4. An element ¢ in K[Yy,..., ¥,] is a unit in K[Y,,.. and only if ee K and e #0. Y,) if An element is called reducible if it can be factored into two nonunits; otherwise it is called irreducible. If u is factored: u = vw, with neither v nor w a unit, then deg v > 0, deg w > 0, and since deg wu = deg v + deg w, we have deg v < deg u, deg w < deg u. This shows by induction that any clement of w in K[Y,..., Yq] can be factored into irreducible factors. Now suppose we have factored u in two ways: use fe w= eg, where c and c’ are units and the f, and g, are irreducible nonunits. ‘Then it can be shown that s = ¢ and that for some rearrangement of the subscripts, 9: = Cif, Where c; is @ unit. Since a unit and its inverse can always be introduced into a factorization to give another, the above factorizations are regarded as essentially the same. These facts are summed in the following theorem, which we will not prove. Theorem 3.5. A polynomial ring in any number of letters over a field is a unique factorization domain. Polynomials f, g such that f = og, where c is a unit, are called associates. Given a complete factorization u = cf, +++ f, of an element w, i.e., a factoriza- tion with ca unit and the f; irreducible nonunits, it is convenient to group the associates among the f, together, to write each f; in a group as a unit times some one element of the group, and to let c absorb the new units thus arising. In this way, the complete factorization is rewritten as = ogti+ ++ gft u = egy K's where c is a unit, the g, are irreducible nonunits, and g,, g, are nonassociates if i #j. The r, are positive integers and g; is said to occur with multiplicity r; as a factor of u. Sometimes for notational reasons it is convenient to allow some r, to be zero, but of course if r, = 0, say, then g, does not occur as a factor of wu. We have already used above that deg fy = deg f+ degg. By the subdegree of f one means the degree of the lowest degree terms in f. It is also true, and easy to prove, that subdeg fy = subdeg f + subdeg g. Now if w = fg is homogeneous, then deg u = subdeg x, from which subdeg f + subdeg g = deg f + deg g, THE EQUATIONS OF A CURVE 13 and since subdeg f < deg f and subdeg g < deg g, we conclude that subdeg f = deg f and subdeg g = deg g. In other words: 3.6. If a homogeneous polynomial u is factored, then its factors are also homogeneous. Now let I’ be an algebraic curve and let F = 0, @ = 0 be equations for T. Write Fc FRB ++ Fry, G@ = d Guay Gt, where ¢, dare units, the F,, @, are irreducible nonunits, F, is not the associate of F, fori #j, and G, is not the associate of G; for i # j. By Theorem 3.3, since ( vanishes wherever F, does, @is divisible by F,,@ = F,H. Factoring H completely and applying the unique factorization theorem, we see that F, is the associate of some G,, and each F, is the associate of some G,. Similarly, each @, is the associate of some F,. Thus m = n, and writing each G, as a unit times an F;, absorbing the new units in d, and rearranging the subscripts, we see that the second factorization above can be rewritten: @ =d Fh--+ Fy. Thus we have the following theorem. Theorem 3.7. If T is given by F =0, where F=c Fu-++ Fp, cis a unit, the F, are irreducible nonunits, F; and F; are nonassociates for i # j, and the r; are > 0, then every other equation of Tis of the form G = 0 with G = d Fh-++ Fim, where d is a unit and the s; are > 0. Taking the r, all as small as possible, namely, 7; = 1, we see: Corollary 3.8. The equations of T' of minimal degree are all of the form cF,F,**+F, = 0. Thus disregarding the inessential factor c, we can say that the equation of T of minimal degree is essentially unique. The equation of I of minimal degree will sometimes be referred to as the equation of T’. Definition 3.9. A curve I is said to be reducible if it is the proper union of two other (algebraic) curves: in symbols PT =T, UT; P47), T £7, IfT is not reducible, it is called irreducible. Theorem 3.10. A curve I is irreducible if and only if its minimal equation F = 0 is such that F is irreducible. Proof. Let T' be given by F; +++ F, = 0, F; an irreducible nonunit, F; not an associate of F, for ij, so that F = F,--- F, is irreducible if and only if s = 1. Suppose first that s> 1. Let I’, be the curve F, = 0, I, the curve F,+++ F, = 0. Since F;, is not divisible by F,, by Theorem 3.3 F, cannot vanish over F, = 0 and a fortiori cannot vanish over I',; therefore 14 ALGEBRAIC CURVES T, is not contained in T',, in symbols: I, ¢ I. Similarly one proves T, ¢ 1. Therefore 1 =T, UT, with P41, FAT, whence I is reducible. Now suppose that s = 1, and let T=T, UT). Let G, = 0 by the minimal equation for I',. Since F = F, vanishes over I’, F, is divisible by each irreducible factor of @, and hence by G,. Thus F = hG,; and since F, is irreducible, 4 must be a constant and T’, is given by F, = 0, ic., T =P, Hence I is irreducible. Thus a curve is irreducible if and only if it is given by an equation F = 0 with F irreducible (or, as one says, by an irreducible equation). Theorem 3.11. Let Vy, 1 be irreducible curves with equations F, = 0, F, = 0, where F,, F, are irreducible. Then T, = Ty if and only if F, and F, are associates. Proof. If F,, Fz are associates, then clearly Ty = Ty. Conversely, if I, =, then F, = 0 and F, = 0 are minimal equations of I, and F, and F, are associates by 3.8. Theorem 3.12. Let T be a curve with F =0 as a (not necessarily minimal) equation and let A be an irreducible curve with minimal equation G@ = 0. Then A is contained in T if and only if F is divisible by G, i. if and only if @ is one of the irreducible factors of F. The irreducible curve A is contained in a union Ty UT, U+++ UT, of curves T, if and only if it is contained in one of the T,. Proof. If F is divisible by @ [in symbols, F = 0(@)], then F vanishes wherever G@ does, so A < I’. Conversely, if A < I, then F vanishes where- ever @ does, and since @ is irreducible, F = 0(@), by 3.3. As for the second part of the theorem, if A is contained in some I’,, then of course A is contained inT,U+++ UT). Conversely, if I, is given by F, = 0, then T,U ++ UT, is given by F,-++ F, = 0, and if ACT, U--- UT, then G is a factor of F,--++ F,, by the first part of the theorem, and hence is a factor of one of the F,, whence A is contained in one of the I’,. Corollary 3.124. If 1, Aare irreducible curves and A < T’, then A =T. Proof. If F = 0, @ = 0 are the irreducible equations for I’, A, then from A 0 (or, for notational reasons, > 0) effective, and otherwise one calls it virtual. As our cycles throughout are the so-called effective divisorial cycles, we may refer to these more briefly simply as cycles. INTERSECTIONS Theorem 3.3, the proof of which was postponed, can be reformulated and strengthened as follows: Theorem 3.14. Let F(Xq, X,, X,) be an irreducible homogeneous polynomial with F ¢C. If G(Xq, X1, Xq) is a homogeneous polynomial not divisible by F, then G vanishes at only a finite number of points of F = 0. 16 ALGEBRAIC CURVES Proof. Since C[X», X;, X;] is a unique factorization domain, we can speak of the greatest common divisor of F and G in C[X,, X;, Xz], and we have G.C.D. of F, @ = 1. Let us consider now the ring C(Xp, X,)[Xq], ie., the ring of polynomials in X, with coefficients in the field C(X,, X,) (= set of quotients f(X,, X,)/9(Xo, X,), where f, g€ C[Xp, X,]). InO(Xy, X,)Xq] the G.C.D. of F and @ is still 1. In fact, suppose that F and G@ have a nonunit common factor A(X», X,, X;) in C(Xp, X,)[Xz]; since & is a nonunit, deg x, > 0. Then F = uh, @ = oh, Shifting the denominator of h to u and 0, we may suppose that he C[Xq, X;, Xq]. Writing u, v as quotients of polynomials with a common denominator d(X,, X,)€C[X», X,] and multiplying through by d, we get dF = Uh, d@ = Vh, where now U, Ve O[X,, X,, X,]. The irreducible factors of h involving X, cannot go into (ie., factor) d, so they go into F; similarly they must go into G. This contradicts that the G.C.D. of F, @ in C[Xp, X,, Xa] is 1. The ring C(X,, X,)[Xq] is a polynomial ring in one letter over a field, and we recall that in such a ring, the G.C.D. of two elements F, @ can be written in the form uF + vG, where u, v are in the ring. Hence we have _ A(X Xr X) B'(Xo X1, Xs) G(X, Xy, X2) + LX, X) W(X, Ky) Xo Xv Xe) where A’, B’, L, M are polynomials (not necessarily homogeneous). Reducing to a common denominator d (Xo, X;) we can write A(X Xy) = A(Xo, Xi, X2)G(Xo, X1, Xa) + B(Xo, Xi, Xa) F (Xo, Xa, X2)- By retaining on both sides only the terms of a given degree, one can assume here that d, A, B are homogeneous (and d #0). Since d(Xy, X;) =0 represents a finite number of lines through the point (0, 0, 1), it follows from the last equality that the common points of F = 0 and G = 0 are on those lines. See Fig. 3.1. Reasoning in the same way with respect to the other variables one obtains that the points of intersection of F = 0 and @ =0 are among the points in which three finite sets of lines cut themselves (see Fig. 3.1) and hence F = 0 and @ = 0 intersect in only a finite number of points. (0, 0, 1) 1,0) AFFINE ALGEBRAIC CURVES 17 We give now another proof of a slightly weaker statement, namely, that if @ vanishes wherever F does, then@ = 0(F), given that Fis irreducible. In fact, suppose @ 4 0(F). As in the previous proof one shows that (Xo, X1) = A(Xo, Xy, Xp)O(Xo, X, Xa) + B(Xq, Xa, X2)F(Xo, Xz, Xo), with d, A, B homogeneous. We may suppose that degx, F > 0. Let F = ag(Xq, X,)X™ + a,(Xo, X,)XP-! + +++, a9 #0. Since d 4 0 as be- fore, the polynomial a d vanishes only for a finite number of ratios 29:2. Leto, x, be such that a9(%9, %,)d (x9, 2) # 0. Then one solves F (2g, 2, 22) = 0 to obtain xp:a,:2, such that F(z», 21,22) = 0. By hypothesis @(a», 21, 22) = 0 also and hence (gy Xy) = A (oy Xyy Le)G(Xqy Ly Xp) + Blsxq, 2, X_)F (Xo, %y, X) = 0, which is a contradiction. The same argument holds for any number of variables, both with Xp, X;,...,Xq in projective space as well as with Y, Y,,...,¥, in affine space. In nonhomogeneous or affine terms, the theorem says: Theorem 3.15. Let G, FEC(Y,,..., Yq] with F irreducible. If G vanishes wherever F does, then G = O(F) (in C[Yy,..., Yn). Exercises 1. One says that G(¥,,...,¥,) vanishes almost everywhere on F(¥y +++, ¥%,) = 0 if @ vanishes everywhere on F = 0 except perhaps where another poly- nomial H vanishes on F = 0, provided H does not vanish everywhere on F = 0. Show that if F is irreducible and G vanishes almost everywhere on F = 0, then G = 0(F). 2. In 3.15, if one drops the hypothesis that F is irreducible, one can still con- clude that a power of G is divisible by F. AFFINE ALGEBRAIC CURVES By a (plane) affine (algebraic) curve one means (as we have already said in Chapter 1) the set of points (x, y) with 2, y in C that satisfy an equation F(X, Y) = 0, where F is in C[X, Y] but not in C. Let F*(Xo, X;, X;) = 0 be a curve in the projective plane. Then F* is a homogeneous polynomial, of degree m, say, and we can write it in the form: F* = agX$ + a,(Xy, X_)XP! + ,(Xy, X)XPPF Hoo, where a,(X;, Xq) is homogeneous of degree i in X,, Xj. Now assunie that the curve is not the line X) = 0. Then F’* is not a power, nor a constant times 18 ALGEBRAIC CURVES a power, of X; hence a; 4 0 for some i > 0. Let a, 4 0, but yy = Oyyg = = 0. Then F*(1, X, Y) = ay + a(X, Y) + a,(X, Y) +--+ + a(X, Y), from which one sees that F*(1, X, Y)¢C. Clearly the point (1, x, y) on F*(X,, X,, X,) = 0 yields the point (x, y) on F*(1, X, ¥) = 0, and con- versely. Thus, taking X, = 0 as line at infinity of the associated affine plane, we have: 3.16. The points at finite distance on a projective curve different from 1, are the points of an affine curve. If F*(Xq, X, X;) = 0 is an equation for the projective curve, then F*(1, X, Y) = 0 is an equation for the affine curve. The converse of 3.16 also holds. Starting with the affine curve F(X, Y) = 0, where F(X, Y) = aq + a(X, Y) + a,(X, Y) +++ + a(X, Y) and a,(X, Y) is the sum of the terms in F of degree i, one merely writes F*(Xp, Xj, Xq) = aX} + a(Xy, X2)XQ* + +++ + a(Xp Xs) to obtain a projective curve, F*(X,, X,, Xz) = 0, whose affine part is the given curve F(X, Y) = 0. Note that (X, X, P*(Xp, Xp Xy) = maz z) : If F(X, Y) is of degree s, we call X, X, XjF @ 2) = F*(Xy Xy, Xs) the standard (or canonical) homogeneous polynomial associated with F; it is of degree s and is not divisible by Xp. If F*(Xo, X, Xz) is homogeneous of degree s and is not divisible by Xo, we call F*(1, X, Y) the standard non- homogeneous polynomial associated with F*; it is of degree s. Thus: 3.17. The standard (or canonical) operations just defined set up a one-to-one correspondence between the homogencous polynomials in Xo, Xy, X» not divisible by X, and the nonhomogeneous polynomials in X, Y. Correspond- ing polynomials have the same degree. A simple straightforward argument shows: 3.18, Let the homogeneous polynomial F*(X,, X;,X;) be # 0(X,). Then F* is irreducible if and only if F(1, X, Y) is. AFFINE ALGEBRAIC CURVES 19 Definition. Let I* be a curve in the projective plane not having X, = 0.as.a component. Then the curve I” consisting of the points not on X, = 0 is called the affine representative of I (relative to X) = 0 as line at infinity). Theorem 3.19. Let U'¥: Ft = 0, I: FE = 0 be two curves not having X= 0as component. Let T,, Vz be the associated affine representatives. Then T* = VF if (and only if) T, = Ty. Proof. Clearly X,F¥* = 0 and X,F¥* = 0 are the same locus. Hence X,F* and X,F¥ have the same irreducible factors. Since F¥, F* are neither = 0(X,), also F¥ and F} have the same irreducible factors. Hence 'f = TF. Definition. Let T* be a curve in the projective plane not having X, = 0 as component. If I* has I’ as affine representative, then (on the basis of 3.19) we call I* the projective completion of I’. Theorem 3.20. Let T,,...,1, be affine curves having I*,...,1* as projective completions. Then the projective completion of 1, U++ UT, isTPUs-UTs, Proof. T¥ U+++U TF is a projective curve meeting X= 0 in only a finite number of points and having [, U--- UT, as its affine part, whence the assertion follows. The notions of reducibility and irreducibility of affine curves are defined analogously to the projective notions: an affine curve I’ is reducible if it can be written in the form P=, UP, P¢T,, PAT, Ty, My affine curves; and irreducible otherwise. 3.21. An affine curve I is reducible if and only if its projective completion P* is. Proof. Suppose that T =T, UT, with PAT, and P #1). Then for the projective completions we have (by 3.19) P* #T'* and 4 TF. Hence (by 3.20) '*¥ = It U I} is reducible. The converse is proved in the same way. Theorem 3.22. Let T,,...,1, be distinct irreducible affine curves and [¥: FF = 0, i=1,...,8, their projective completions. By 3.21 the VF are irreducible, and the F¥ are taken irreducible. Let F(X, Y) = F¥(1, X, Y). (By 3.18 the F, are irreducible.) Then all equations for Ty Us UT, are of the form FuFy +++ Fl: = 0 with the r, > 0. Proof. Let F(X, ¥) = 0 be an equation for T, U+++UT,. Let F* be the standard homogeneous polynomial associated with F. Then F* = 0 is an equation of P¥ U-+» UT*. The I are distinct, hence F* is of the form F* = F¥"--- F¥ with the r; > 0. Hence F = Fp-++ Fy. QED. 20 ALGEBRAIC CURVES Defining cycle in the affine plane as the notion made up of a finite number of distinct irreducible affine curves T',, , I’, and associated positive multiplicities r,...,7, and taking F--- Ff 0 as the equation of nl, +++++7,0, one obtains: Corollary 3.24. There is a one-to-one correspondence between the cycles in the affine plane and those in the projective plane not containing X = 0. The one-to-one correspondence between the cycles is obtained through the one-to-one correspondence of their equations. The last few theorems and definitions are simply translations into affine terms of previous projective theorems and definitions. Such translations, and the reverse translations, are usually (though not always) a straightforward and simple matter. Usually, then, we will make and use these translations without much ado, Exercises 1, Sketch the curve X¥ = 1 in the real affine plane. Sketch its projective completion X,X, — X} = 0 in the real projective plane (especially near the line at infinity). 2. Find the points in which (the projective completion of) X* — 6X*¥ + 11X¥*—6Y*41=0 meets the line at infinity, Zo. 3. Sketch (the projective completion of) Y? — X* = 0 in the real projective plane. Do the same for ¥ — X* = 0. . All the curves X? + ¥? + aX + bY + ¢=0 go through the same pair of points I, J on the line at infinity. (These are the so-called circular points at infinity.) 5. Let F(X, X;, Xz) be homogeneous. Show that if F(1, X, ¥) = 0, then F(Xy, X;, Xz) = 0. 6. Let F(X», Xy, Xz), G(Xp, Xi, X_) be homogeneous polynomials with deg F > degG Assume that F(1, X, ¥) = Gl, X, ¥). Then F (Xp, Xy Xz) = Xj G(Xy Xy KX), where d = deg F — deg G. 7. a) Show that - Xn Xue Xn Xez is irreducible in C[X,,---, Xeal- b) Show that Xn Xe Xes {Xu Xn Xs |: Xs Xss is irreducible in C[Xy,.-., Xsal- Chapter 4 RESULTANTS Consider two polynomials $= X" GX" $+ ay, G = BX* + OXON boot bys with indeterminate coefficients a = (ao,...,@m), 6 = (bg---5 0p). Let us suppose that for certain special values a = d, b = 5 of the coefficients, these polynomials, f and g, have a root X in common, that is: GX + GMA fe + dy = 0, bX" + 6X41 4 ---4 5, = 0. Multiplying the first equation successively by X*-1, X*-2,...,1 and the second by X"-1, X"-2,..., 1, we obtain the equalities GXmina 4 Gmina g Sb dig Xd =0, Gykmin-2 4. =0, =0, ByXmin-z 4 Hy Nmine 4 =0, ByXMiO® pees eeveeeceeeeeeeees + 5,2 =0, BX" 4 BRA peeeeeeeeee +6, =0. Considering the homogeneous linear system with @ and 5 as coefficients 21 22 RESULTANTS obtained by replacing in these equalities the distinct powers of X by in- determinates, we see that the system thus obtained has a nontrivial solution, namely, Xm+n-1, Xm+n-2,___ 1. Hence the determinant of the system is equal to 0. This determinant is called the resultant of f and g, and one has proved that: 4,1. A necessary condition for f and g to have a common root is that the resultant be zero. [Multiply the left-hand sides of the above equations respectively by the corresponding cofactors of the last column of the above determinant and sum: by familiar theorems on determinants, the coefficients of X™+"—,...,X in the sum will be zero and the constant term will be R(f, 7). Moreover, the sum is a linear combination of X"-4f(X), X"—1g(X),..., G(X). Hence one obtains: Rf, 9) = A(K)f(X) + BX) G(X), and, for indeterminate a, 6, X, 4.11. Rif, g) = A(X)f(X) + B(X)g(X), where A(X), B(X) are polynomials of degrees n — 1 and m — 1 respectively.] 4.2. If R(f, G) = 0 and deg f = m or deg g = n, then f = Oand g = have a common solution. Proof. f = 0 and gj = 0 have a common solution if and only if f and g have a common factor h of positive degree. Suppose that f = c(X)h, g = d(X)h with deg h > 0. Then df = cg with deg c < deg f and c 40 (also, of course, deg d < deg g and d #0). Conversely, if df = cg with deg c < deg f and c + 0, then f, g must have a common factor of positive degree. In fact, we factor both sides completely, cancel the factors from c, and are still left with a factor of f, which goes into g. Thus /, g have a common factor if and only if there are polynomials c #0, d #0, dege <™m, deg d A(px) = pAx, which also represents P,,. The transformation obtained in this way will be denoted Ly. Such a trans- formation is called a homogeneous nonsingular linear transformation. The transformation L, is univalent (i.e., it transforms distinct points into distinct points) and it is onto (i.c., every point of the plane is an image of some point under L,). It transforms straight lines into straight lines. Associated with 7 we consider a transformation on the set {¥(X)} of homogeneous polynomials according to the rule: F(X) > F(A>X) = F(X). 27 28 LINEAR TRANSFORMATIONS We agree to call 0 homogeneous of any degree. Then we have the following properties: 1) If F is homogeneous of degree n, then F’ is also homogeneous of degree n. 2) If F = 0, then F’ 3) If F’ = 0, then F x—> Ax] 4) The value of F at x is equal to the value of F’ at x’ = Ax. Hence the locus of F = 0 is transformed (under L,) into the locus of F’ = 0. 5) If F is reducible, then also F’ is reducible. 6) If F’ is reducible, then also F is reducible. [One applies (5) to the transformation x —» A-'x.] Hence if F is irreducible, then also F’ is irreducible. 7) An equation for a curve I’ of minimal degree is transformed into an equation of minimal degree for the transformed curve I’. 8) If F and Gare associates, then F’ and G' are associates and conversely. Thus if F and G are not associates, then F’ and G’ are not associates. 0. [One applies (2) to the transformation Definition. The degree of an equation of minimal degree for a curve ° is called the order of I’. 9) The order of I’ is equal to the order of I’. The assertion (9) can be expressed by saying that the order of I is a projective invariant. The second part of (4) is expressed by saying that the concept of an algebraic curve is a projectively invariantive concept. According to (6), the irreducibility of an algebraic curve is a projectively invariantive concept. 5.1 (Fundamental theorem of linear transformations). Let Ay Ay, Aa, Ag be four points no three of which are collinear; let Bo, By, By, Bg be another four points no three of which are collinear. Then there exists one and only ‘one homogeneous nonsingular linear transformation sending A; into B,, i = 0, 1, 2, 3. (We omit the proof.) Therule L, = L, 4 isimmediate. The converse is also true: if Ly = Ly, then B = pA for some p. Exercise. Let ae No three of the A, and no three of the B, are collinear. Find a matrix A such that L,(A,) = B, i = 0,1, 2,3. COORDINATE SYSTEMS 29 COORDINATE SYSTEMS Let Ao, Ay, Ag, Ag be four points no three of which are collinear. Let Ly be such that E, is sent into A;, i = 0, 1, 2, 3 under Ly. Here Ey: (1,0, 0), E,:(0, 1, 0), Ey: (0, 0, 1), By: (1, 1, 1). Given a point P, we assign new coordinates to it according to the rule: new coordinates of P = coordinates of L4+(P). That is, designating the original coordinates of P as x and its new coordinates as x’, one obtains x’ = Aol, One will say that Ag, 4,, Ay, 4g is a system of coordinates: Ag, Ay, Ay are called the first, second, and third vertices, respectively, and Ay is called the unit point. Let a curve IP’ be given in the old (or original) coordinate system by F(X) = 0. Then one sees that in the new coordinate system I’ is given by F(AX) = 0. The algebraic formulas that we use in considering transformations and. in considering changes of coordinates are the same, but the point of view is different. In the latter case, we speak of an alias transformation and in the former, of an alibi transformation. The basic facts are the same in either case, but sometimes it is convenient to adopt one point of view, and some- times the other. In particular, to study I’ in coordinate system AyA,494y is the same as to study L,(I’) in the coordinate system E,Z,E,E,. In more technical terms, this assertion can be formulated as follows: The equation of I’ in the coordinate system A,4,A,A, is also the equation of L,.(P) in the original coordinate system. The preceding are preliminaries to case (b) of the last theorem of Chapter 4. We proceed as follows. The idea is to transform I’, A by means of a transformation L4 into curves ',, A, that will be in permissible position, i.e., into curves for which the hypothesis of case (a) holds (see the first sentence of the proof of Theorem 4.6). The intersections P,,...,P, of T and A are transformed into the intersections P#,..., P4 of Py and Ay. The order is a projective invari- ant; thus ord, = m, ord Ay =n. Hence applying case (a) to Ty and Ag, we obtain 1 (Xo, Xi, 3), a = Layey, in affine terms. We take Xj = 0as line at infinity and confine our attention to points P, such that P, and its transform P,, are at finite distance. Then P, has the affine coordinates (y;, yo) = (x,/9, %2/%)) and P,, has the affine coordinates (yi, yi) = (2/24, «i/z)). Thus yf = 1 + nt + Ait, _ io + ati + tine © Agog + Aoytr + Moat, — Goo + dort + Aine i = 1,2. As we have already said, we are not considering points P, on the line do9Xo -+ G,X1 + doeXe = 0. If ap, or dog # 0, then this line has the affine equation ay) -+ 4,¥1 + dog¥2 = 0. Otherwise it is the line at infinity, and the transformation takes the simpler form: Yi = bro + bts + brave, Ys = bao + bath + beater where b;; = @,;/a9. Here j1 0 0 |A B10 by Ors sou | bao bay bap whence b1:0.9 — bexbyg # 0. Hence one can solve for (y,, yz) in terms of (yt, yj); the transformation is onto the affine plane. These transformations are called (nonsingular) affine linear transformations. They are the trans- formations induced by the nonsingular homogeneous linear transformations of the associated projective plane that leave the line at infinity invariant. #0, 3 Ty Exercise. Let B,, B,, By be three points on the irreducible conic dgX3 + --- = 0. By a linear transformation taking By, B,, B, into Ey, Ey, E,, the equation of the transformed conic takes the form ¢,X 9X, + coX,X, + ¢,X;X)=0 with Cocca # 0. The linear transformation x/ = s,/¢; (i.e., x > x’) transforms this conic into X)X, + X,X, + X,X, = 0. Hence any irreducible conic can be sent into any other by a linear transformation taking any three given points of the first into any three given points of the second and, moreover, this can be done in just one way. Chapter 6 SIMPLE POINTS AND SINGULAR POINTS Consider the intersection of a curve I’ having F(X, X,, X_) = 0 as equation of minimal degree with a straight line L. Let x° = (29, 2%, 22), x! = (2}, x}, 23) be two points of ZL. An arbitrary point of L will then be of the form Ayx® + Axl. This will lie on I’ if and only if F(Agx° + 4.x?) = 0. Thus to find the intersections of ' and L one solves the equation F(A x® + A,x!) = 0. It is possible that F(A x + A,x1) is the element 0 in C[Ay, A,]. This will be so if and only if every point of Lis on I’. Let us consider the case that Lis not contained in I. Let k B(x? + Aye!) = H (Mo, Ay) = ¢ TT (Ao? — Ara)" with 49:20 4 AD:4D for i #j,cEC,c¢ #0. Definition, The number s; is called the intersection multiplicity of P and Lat the point P; determined by A{): 4, ie., the point Aix? + AMx1, A similar definition can be given for a cycle F = 0. We shall show later that the intersection multiplicity of T and L at P; is a projective invariant and does not depend on the choice of x? and x! (but only on L (and I)]. We suppose now that the point x° is taken to be on the curve I. Then x° = 1+ x? + 0+ x1 and corresponds to the ratio A,:4) = 0:1, and hence to the factor Ay-0 — Ay+1 = —A, of F(Agx® + Ayx}). Let us place Ay = 1, Ay = A. (This is similar to the passage from the projective plane to the affine plane explained in Chapter 3.) Then we have: 6.1. The intersection multiplicity of L with V at x° is equal to the maxrimum power of A that can be factored from F (x9 + Ax?), Moreover we have F(xp + As), x} + Act, 2} + Ax3) = F(x?) + A(Ho(x)x) + Hy(x°)x} + Halstad) + AN) po 31 32 SIMPLE POINTS AND SINGULAR POINTS Two cases may present themselves: a) (H((x°), H,(x°), Ha(x°)) = (0, 0, 0), b) the remaining case. In case (a) every line L through x® cuts T' at x° with multiplicity > 2. In case (b) L intersects I at x® with multiplicity > 1 if and only if H,(x®)xb + H,(x%)z) + H,(x°)z} = 0, ice., if and only if x? is on the line Ho(x°)Xo + Hy(x)X, + Hy(x°)X2 = 0. In case (a) one says that x° is a singular point of T' and in case (b) that it is simple. The line mentioned in case (b) is called the tangent to T’ at x°. One will see below that these concepts are projective invariants. Exercises 1. Find the intersections of 25X} — X? — X}=0 and the line joining Po: (1,7, 1), Py: (1, -1, 7). 2. If F has two (not not necessarily distinct) factors that vanish at x°, then the intersection multiplicity of the cycle F = 0 with any line through x° is greater than 1. 3. Let I’ be the cycle having X$+ X}+ X{=0 as equation, and let (a) 2, 22) be a (L.e., any) point on it. Show that: (a9, 2;, x2) is simple on I’. Hence the cycle I’ is an irreducible curve. (See the remark at the end of Chapter 1.) Write the equation of the tangent to T’ at (#9, a, #2). 4, Do Exercise 3 for the cycle X} + X} + X}+ XX, =0. We now study the nature of a curve I’ at (or, as one also says, near) a point P of the curve. Since any point is in an affine plane (for some choice of the line at infinity), we make an affine study: this has at least the advan- tage that there is one less variable to deal with. Ee Fig. 6.1 Let F(X, X;, Xz) = 0 be the given curve and M, the point we wish to study; here F(X) = 0 is an equation of minimal degree for I. We may suppose that M, = Ey = (1, 0, 0), that is, we take M, as first vertex of a coordinate system (see Fig. 6.1). Let E, = (0, 1, 0), Ey = (0, 0, 1) and A=(0, 1, m). Let L be the line E,A: its equation is X, — mX, = 0. Passing to affine coordinates, the affine part of the curve I’ is given by the equation F(1, X, Y) = 0; we write F(1, X, ¥) = F(X, Y). The affine SIMPLE POINTS AND SINGULAR POINTS 33 equation of L in Y — mX = 0. From 6.1, the intersection multiplicity of E,A with I’ at Ey is the maximum power of A that can be factored from F(l+A-0,0+A-1,0+ A+m) = F(1, A, Am). Thus: The intersection multiplicity of T and L at (0, 0) is the maximum power of X that can be factored from F(X, mX). Notation. The intersection multiplicity of T and L at P is denoted by iD, L; P). Exercise, Let I’, A be curves without common component meeting at a point P and let L be a line through P. Show that ul UA, L; P) = ul, L; P)+ aA, L; P). We write F(X, Y) = gg -+ GyoX + ay ¥ + aggX? + ay XY + agg ¥? + +++ The point (0, 0) is on Tif and only if ay) = 0. We suppose this to be the case. Corresponding to the alternatives (a) and (b) above, we assert: 6.1.1. a) If a) = do, = 0, then every line through E, cuts T at Ey with multiplicity > 2. b) If ayo or ag is not zero, then every line through Ey with just one exception cuts T’ at E, with multiplicity 1. This exception, the tangent (by the defini- tion already given), has as equation: aX + ay,¥ = 0. In fact, substituting Y = mX in F(X, Y), one obtains in case (a) F(X, mX) = dy X* + aymX? + agm®X? $+, ie., at least X? can be factored out. This takes care of every line through (0, 0) except the vertical line X = 0. The result for this line follows by interchanging the roles of X and Y (i.e., by what has already been proved). In case (b), if, say, dy, # 0, then one obtains F(X, MX) = (ayo + magy)X + X*--+) and X but no higher power can be factored out, except in the case that yy + may, = 0, ie., if and only if m = —ayo/ay. This takes care of every line except the vertical line X = 0; here the intersection multiplicity is given by the maximum power of ¥ that can be factored out of F(0, Y), which one sees is 1. This gives Y = —(ayo/a,)X or aX + a,Y¥ =0 as the tangent. Ifa = 0, then ay, % 0, and the desired result follows by an inter- change of the roles of X and Y. In case (b), i.e., when the point is simple, one can say, speaking without precision, that the tangent approximates the curve near the given point. We may write our polynomial F(X, Y) as F(X, ¥) = FAX, Y) + FralX ¥)+-++, F(X, ¥) #0, 34 SIMPLE POINTS AND SINGULAR POINTS with F,(X, Y) homogeneous of degree i and with r > 1. The point is simple if and only ifr = 1. For r > 1 one has 7 F(X, ¥) = ¢ T] (mPX — mPY)%, t=1 where c € C and mi: m{ ¢ m{):m) for i # j. In this case, i.e, when the point (0, 0) is singular, every line through this point cuts I’ at the point with multiplicity r except for the lines mPX —mOY =0 (Gi =1,2,...,8), which are called the tangents to Tat (0, 0). In this case one says that (0, 0) is of multiplicity exactly r. One says that the point is ordinary if every 5; = 1. Examples 1) The curve Y? — X? — X5 = 0 passes through (0, 0). The poly- nomial Y* — X* — X3, which determines it, is irreducible. In this case r = 2 and every line through (0, 0) intersects the curve at (0, 0) with multi- plicity 2 except for the tangents, ie., the lines given by the factors of Y? — X*=0,or Y — X = Oand Y + X = 0. This is shown in Fig. 6.2. This type of singular point is called ordinary double point. The curves Y?— X24 X3=0 and Y? 4+ X?+4 X%=0 also have at the origin ordinary double points. Fig. 6.2 2) Let I have at the origin a nonordinary double point. Then at the origin there is just one tangent to I’; let this be the X-axis. Let F = 0 be the minimal equation for T. Then F = ¥? + ayX* + ayX?Y + ayX¥? + ayg¥2 +-°° The origin is called an ordinary cusp if the tangent cuts the curve there with multiplicity 3. Necessary and sufficient for this is that aj 40. Thus Y? = Xhas a cusp at (0, 0). The appearance is illustrated in Fig. 6.3. Not all nonordinary double points are ordinary cusps. For example, the origin is not an ordinary cusp of ¥? — X4 — X5= 0. The appearance here is shown in Fig. 6.4. SIMPLE POINTS AND SINGULAR POINTS 35 Fig. 6.3 Fig. 6.4 Remark. The above does not yet say what is essential about a cusp. Later we shall define branch, and a cusp will be defined as a double point at which there is only one branch. If the double point is a cusp at the origin, then (in an appropriate coordinate system) for points near the origin, y is given by a power series in x'/2; whereas if it is not a cusp, then the points near the origin are given by two power series in z, ie, y = P,(x), i = 1, 2. Theorem 6.2. If T and A are two irreducible curves, then every common point of T and A is a singular point of the curve TU A. Proof. We may suppose that the common point is the origin. If F = 0 and @ =0 are the (minimal) equations for I’ and A, then F- @ = 0 is the minimal equation for I’ U A (we are supposing that I’ # A). Since subdeg FG = subdeg F + subdeg @ > 2, the origin is singular for FG = 0. Q.E.D. Remark. One can define simple and singular points for cycles. The theorem corresponding to Theorem 6.2 would then say: If I’ and A are cycles and P is a point common to T and A, then P is singular for the cycle T' + A. Corollary 6.3. Every point of a curve that belongs to two distinct com- ponents of it is singular. Let F(X, Y) €CLX, Y], (a,b) a point. Then F(X, ¥) = ¥ ejX'¥) = Y ofa + (X — a)} [6 + (Y — by = LY eya'bi + Y cyiat 0X — a) + Y eyja/-(Y — 6) + terms of degree > 2 in X — a, ¥ —6 (Y¥ — 6) \(a,b) + higher-degree terms. oF oF| = Fa, b)+ a (XK —a) +55 \(a,b) Here oF| OX la) is the partial derivative of F(X, Y) with respect to X evaluated at (a, 6). 36 SIMPLE POINTS AND SINGULAR POINTS (In writing @F /@X, we do not have to enter into the notions of the calculus, but merely work formally, using the familiar symbols and rules.) If (a, 6) is on the curve, then F(a, b) = 0 (and conversely). Then writing X' = X —a, Y' = Y — b, the equation oF aF| eX\an) Fan is simply the equation of F = 0 in a new coordinate system having (a, 6) as origin. Thus: ¥'+ 0 Theorem 6.4. The singular points of F(X, Y) = 0 are the common points of F(X, Y) =0, 6F/aX =0, and aF/2Y =0. At a simple sian es i thee ocean eo une tel oF oF S|) (xX-a+S) (y-b=0. @X\ian Xian) Corollary 6.5. A curve T' has at most a finite number of singularities. Proof. First suppose that I": F = 0 is irreducible. F involves at least one of the letters X, Y,say Y. Then 2F/@Y # 0, hence deg (@F/2Y) < deg F. Hence @F/2Y cannot be divisible by F, so by 3.14 (in nonhomogeneous form), 2F/@¥ can vanish only at a finite number of points of I’, which must include all the singularities of I’ (at finite distance—and at infinity P only has a finite number of points). If I is reducible, the corollary still holds, as the singularities of Pare the singularities of its individual components plus points of intersection of pairs of components. Corollary 6.6. Let 0: F(X, Y) = 0 be a curve and P: (a, 6) a point thereon. Then oF| ay if and only if either P is singular for Tor it is simple and the tangent to T° at P is vertical. =0 \a,b) Proof. If P is singular, then Fi eG O¥ an by the theorem. If P is simple, then the tangent at P is oF| Senet gy eae (Lie pyltat gy aX 0,0) aX | a This line is vertical if and only if oF| a = 0. 6Y| (a,b) SIMPLE POINTS AND SINGULAR POINTS 37 Corollary 6.7. Let T be a curve of order n, La line not a component of T. Then L cuts ' in at most n points, and this number is reached for some L. Proof. We already have the first assertion (see 3.1). For the second, take the coordinate system so that E,: (0,0, 1), the point at infinity on the Y-axis, does not lie on I’. Through E, take a line not passing through any singularity of I’ and not tangent to I: to see that this can be done, consider first the case that I’: F(X, Y) = 0 is irreducible. Then F(X, Y) involves Y: otherwise F(X, Y) would be a polynomial G(X) in X alone, in fact of the first degree, since G(X) is irreducible; F(X, Y) = 0 would be a vertical line, therefore a curve through E,, and this is not so. Hence 2F/@Y 4 0. Now aF /Y # O(F), as deg (@F/2Y) < deg F. Hence aF/@Y = 0 at only a finite number of points P,,...,P, of T. Any other point of Pat finite distance is simple with nonvertical tangent. Take, then, a line through E, not through P,,...,P, and # L,,. Such a line will satisfy the require- ments. If 1 =I, U-++ UT, with the T irreducible, one takes a line through E, not tangent to any T’,, through none of the singularities of the T,, and through no intersection of a T'; and a T;, i #j. Such a line L will meet I’ in n points each counted simply, hence in n distinct points. Exercises 1. Show that the polynomials H,, H,, H, on p. 31 are 8F/0X,, 2F/@X,, OF /2X, respectively. Hence in homogeneous terms, the singularities of F (Xo, Xy, X2) = 0 are given by the simultaneous solutions of F = 0, 2F/2X, = 0, 2F/2X, = 0, aF/aX, = 0. 2. Let F(Xo, Xi, Xz) be homogeneous of degree n. Show that X, OF/@X, + X, OF/0X, + Xz OF/aX, = nF. (This is known as Euler’s Formula.) Use this identity and the affine criterion for singularities (Theorem 6.4) to get the projective criterion given in the last exercise. 3. Let F(Xq, X;, Xq) be homogeneous of degree 2. Then F can be written in the form (Xo, Xi, Xa) [oo 01 Goa\ [Xo Ao 4%, Ae] [ X1 420 a1 ae) \ Xe, with aj; = ay; (all i, j). Show that F(X) = 0 has a singularity, hence is reducible, if and only if det (aj) = 0. 4. Find the singularities (in the complex projective plane) of a) X94 Y3 = 3X¥ (Folium of Descartes), b) (X? + ¥2(X —1)?= X? — (Conchoid of Nicomedes), c) (X?-4 ¥2)) = 4x27? (Four-leafed rosette), and sketch the curves (in the real projective plane). 38 SIMPLE POINTS AND SINGULAR POINTS 5. Sketch the curve ¥? + X* + X= 0 in the real affine plane, especially near the origin. 6. Let Xo, Xj, X, be indeterminates over C and let x, = X3/Xy. a) Show that if f(Xo, Xj, X:)€C{X, Xj, Xz] vanishes (Xp, X,, 2), then f = 0(X} — XyX;) in CLXp, X, Xs). b) Show that C[X,, Xj, a4] (= set of elements in the field C(Xp, X,) that can be written as polynomials in X», X;, 2, With coefficients in C) is not a unique factorization domain. THE POLAR Definition 6.8. Let I’: F(X, X, Xz) = 0 be a curve (or cycle) and let P: (ap, a, a) be a point. If oF re te ete #0, 8X, i then by the polar of P relative to T one means the cycle oF oF oF % ox, + “ax, | ax, = Theorem 6.9. The notion of a polar is projective. In more detail, if X' = AX, where Aisa3 x 3 nonsingular matrix and F(X’) = F(AX’), then oF’ oF oH ax; La 2x,’ where a’ = Aa. Proof. Let deg F = n. By a computation like the one leading to 6.4 (or by the directly ensuing computational rules) F(X +a) = F(X) + rae ++ terms of degree ai(oF’/0X!). QED. Let us take a coordinate system in such way that P gets the coordinates (0, 0, 1), ie., P = By, Then the polar cycle, if defined, is F /2X, = 0. The polar cycle will not be defined if and only if F does not involve X,, or, expressed invariantively, if and only if F = 0 consists of a number of lines through P. SOME COMPUTATIONAL RULES 39 OF (Xo, X1, Xs) _ @F(1, X, Y) i eY ox, Ix=(1,x,¥) one sees that the affine part of the polar of G(X, Y) = 0 with respect to By: (0, 0, 1) is aG/a¥ = 0. Exercises 1. If Pis on ,, then the polar of P (if defined) passes through P. 2. If Pis simple on I’, and A is the polar of P, then A has P as simple point and Tand A have the same tangent at P. Theorem 6.10. The polar of P with respect to T (assuming it is defined) passes through the singular points of T', through the points of contact of the tangents from P to V', and through no other point of T’. Proof. We may suppose that P is at E,. Let Q be any other point of F(X, X;, X,) = 0. We may suppose that Q is at finite distance: let Q = (a,6). Then Q lies on the polar if and only if @F(1, X, Y) oF =0. \(a,b) By 6.6, then, Q is either singular or is a point of contact of a tangent from P. This takes care of every point Q of I’ except P. If P is not on I’, there is nothing to prove; and if P: (a, a, a) is on I, then P is on the polar > a, @F /eX, = 0 by Buler’s Formula, which says that YX, 0F/0X, = nF. The proof is complete. SOME COMPUTATIONAL RULES We summarize some rules familiar from calculus. Throughout we deal only with polynomials; as a consequence these rules are really merely com- putational rules. To emphasize this, we start from the following definition: Definition. Let {(X) € O[X] and let h be a new indeterminate. Then S(X +h) € CLX, h): S(X + A) = fo(X) + AXA + fa(X)M +--+, f(X) eC[X], and by definition the derivative of f(X) is f’(X) =f,(X), ie., the coefficient of hin f(X + h). Placing h = 0, note that {(X + 0) = f,(X), so that we write: F(X + A) = f(X) + F(X Yk + >»). a) f +9 =f +9, b) (fo) =f'9 + fy’, c) c =0,for ced, d) X’=1. 40 SIMPLE POINTS AND SINGULAR POINTS Proof. Letus prove (b); the others are proved similarly : F(X +h) = f(X) + f'(X)hk + F-+), g(X + h) = g(X) + g'(X)hk + W---), hence F(X + h)g(X +h) = f(X) G(X) + (F(X) G(X) + f(X)g'(X)M + W--), and the coefficient of / is the familiar expression. Q.E.D. For e) (X")' =X the proof is by induction, using (b). Hence f) (SaXi) = > ie. XA, Now for the function of a function rule: 8) f(9(X)) = f'(9(X)) - 9(X)- Proof SK +h) = f(X) + f' (Kh + >). Therefore, S(u + v) = fw) + fiw + o%---) (u,v OLX), g(X + h) = g(X) + g'(X)h + W--+) = gX)+v=u+y, S(GX + h)) = f(G(X)) +f ( GX) G(X yh + b+ -)] + [g'(X)h + YPC) = f(W(X)) + f'(G(X)) 9 (Xb + A+). If instead of C' we have any other coefficient domain, say C[Y], where Y is a second indeterminate, then the above rules and proofs continue to hold. In the stated case, the derivative is denoted 2/@X. Similarly, looking at C{X, Y] in the form C[X][Y], one defines o/@Y. o0F «. or axay * denoted 5xOY* Since oF é gray (2 wX'Y!) = oe (Z jeuX'¥) = Y tjeyX OY, SOME COMPUTATIONAL RULES 41 and similarly for 63/2Y@X, we have: ae @ exeY ayax’ in other words, the operators [0X and 0/2Y commute. h) Now for Taylor's Theorem: ie 1 i) f(X +h) =f(X) +f (Xp +i") y +i@ a eee Proof. f(X +h) = f(X) +f’ (X)b-+ SAX) + h-++). Applying the fune- tion of a function rule on the variable h, we get P(X WL = F(X) + (Xb + WO), whence from the definition SX) = (FY = 2X), showing that Taylor’s Theorem is correct so far as the coefficient of A? is concerned. Now we make an induction, assuming that pen @—D! $F A(X) + WAG), for some i and every polynomial f(X). Applying the function of a function rule, we have he 7 SK $B) = FR) + PORE IR) bo ELEM) hi @=2! + i f(XWO + hE), and applying the induction assumption to f’(X), we find i f(X) = (F(X))G — 1h, and the induction is complete. Writing Df for f’, D*f for f”, etc., we can write Taylor's Theorem symbolically in the form SUR EM LHX) 4 PRM $$ fMR) eD* SR +A =F AD + SPE + SK)- Reealling from calculus that for any number x 2 @altetyts 42 SIMPLE POINTS AND SINGULAR POINTS we also write the above (more symbolically) in the form (*) S(X +h) = eb 7(X). For several variables, let X abbreviate (X,,..., X,) and h abbreviate +h»), Where the X, and h, are 2n indeterminates or letters, and let f(X;,.-., X,) by a polynomial in C[X,,... , X,]. Let Dy = 0/0X,,..., Dy = 0/0Xy. Then F(X + hy, X_ + hg, Xn + In) = ehPif(X,, Xp + hg, Xp + hy», Xp + en) = enPa(ehPa( F(X, Xyy Xq + hy 1 Xn + fa) and, since (,D,)!((tgD,)if) = (hiALD! Dif, this = (en eePs) F(X, Xap Xp + hey Xu + Is where = 14 AD, + BPE... and Ons = 1 hyDy + BPE... are to be multiplied out according to the usual rules for working with numbers. Now one knows that e% + ev = ert, when « and y are numbers, and one checks directly that also Ds «ghar = ghDithDs, Continuing, then, our previous computation, we find F(Ky A yy Xp ig) = em Pet HD al(X, Ky) An abbreviation for h,D, +-++-+ h,D, is BD, so we obtain Taylor's Theorem for several variables in the form (x) S(X +h) = ePP4(X). Exercise. Let I': F(X) = 0 be a cycle with deg F = n and let a be a point. Then F(X + a) = F(X) + F,(X,a) + F,(X,a) +--+ + Fy(X,a) + Fla), where F; is homogeneous of degree n — i in X (and homogeneous of degree i in a). If F,, F;41 define cycles (i.e., are not = 0), then F,,, = 0 is the polar of F, = 0 with respect to a. Chapter 7 BEZOUT’S THEOREM In Chapter 5 we said that we would prove the following results: Theorem 7.1. The intersection multiplicity of T and L at P; (a point of intersection of T and L) is independent of the points x°, x1 used to determine L. Theorem 7.2. The intersection multiplicity of T and L at P, is a pro- jective invariant. However we still postpone the proof for reasons that will shortly become clear (see the remark at the end of the chapter). Now suppose we have two arbitrary curves (or cycles) instead of a curve anda line. Let these be TP: F(X, X,X,)=0 and A:G(X, X,, X,) =0 of orders m and n and without common components. Let P,,..., P, be their points of intersection. Problem. To define the intersection multiplicity uD, A; P;) of I’ and A at P; in such a way that ‘ Dil, A; P;) = mn. j=l Suppose first that ' and A are in permissible position with respect to the coordinate system Ey, Hy, By, Ey; that is, at least one of the curves does not pass through E, and no two points of intersection of and A are collinear with Ey. Let P; = (ti ta, tq). Then R,( F(X), @(X)) = cT] (Xizig — Xora)", ce, c #0. 43 44 BEZOUT’S THEOREM We can then propose: uD, A; Pi) = Observe that degy, F < deg F, degy, @ < deg @ and that equality holds at least once, since I’, A are in permissible position. In forming Ry(F, @) we agree to give to F and G formally the degrees m and n in X. If I’ and A are not in permissible position, then one considers a linear transformation L, such that if Ly: 1 —+>T4, A— Ay, and P; — PA, then Ty and A, are in permissible position. Then one can define iP, A; P) = iP y Ay; PA. It is clear that for this definition to be correct we shall have to prove that if Lg is a linear transformation sending I’ and A into curves I, A? in per- missible position, then (*) iy, Ags Pf) = ig, Ags PP). Once this is proved, we shall have: je Theorem 7.3 (Bezout’s Theorem). Two curves of orders m and n without common components meet in mn points, counting their multiplicities. By a complete set of intersections of T and A one means the system Py, Pz, ..., Pn of intersections, taken in any order, where a point P of multiplicity r occurs r times in the system. The equality (*) amounts to saying that the proposed definition of intersection multiplicity is a projective invariant. This in turn could be said to be the main point of Bezout’s Theorem. To prove the equality we will introduce a general linear transformation Ly. This is determined by the matrix Ugo or Wor U=|% wn the} > 29 a1 Ua: where ‘go, ..-, Ug are nine indeterminates over . In order to prove the desired equality (*) we will prove: (#%) iy, Ags Pf) = Ty, Aus PH). Now we will work over the field C, = C(t...» Ugg) = algebraic closure of Clg, «++; Ug2), Le., Cy i8 algebraic over C(t, ..-, Ugg) and C, is algebraically closed.t All that has been proved up to now for (is valid for Cy. We will designate by PNC the projective number plane over C and similarly PNC,. One has that PNC < PNC,. Let us suppose that F = 0 isa curve in PNC. In PNC,, F = 0 defines a curve that contains the points of F = 0 in PNC. To continue the proof of (+*), we first prove the following theorem. BEZOUT’S THEOREM 45 Theorem 7.4. Let T and A be curves in PNO without common com- ponents and let T', A be given by F = 0, @ = 0, F, @EC[Xy, Xy, Xp]. Then the intersections of F = 0 and @ = 0 in PNC, are the same as the intersections of F = 0 and @ = 0 in PNO. We may take the curves I’, A in permissible position with respect: to E,H,E,B, and suppose moreover that no intersection is on ByH, (the line at infinity). Using the resultant Ry,(F, @), we see that the points of inter- section of F = 0 and @ = 0, whether regarded in PNC or in PNC,, are on the complex lines through E, given by the factorization over C of Ry(F, @) = H(X,, X;) into linear factors, A similar statement is true for Ry,(F, @). Since the intersections of these two finite sets of lines are points with complex coordinates, the theorem is proved. We return to the linear transformation Ly. Let Ty be the curve F(U“X) = 0 and Ay the curve G(U-X) = 0. Let PY be the transform of P;. We now assert that I, and Ay are in permissible position. In fact, if (0, 0, 1) satisfied F(U-!X) = 0, then, specializing U to A (ie., substituting ay; for u,;), we would obtain that (0, 0, 1) is on F(A-1X) = 0 for every A, which is surely not so. Analogously, one shows that PY PY does not pass through E,. Since the PY are the points of intersection of 'y and Ay, we have Ry (F (UX), GUAX)) = TT (Xi (woorio + tortin + Uox%is) — Xoluretig + Marta + tata) with c’ in C,. Observe first that c’ © C(tg, ..-, Ugg): one sees this by comparing the coefficient of any power product of X,, X, occurring on the right with the coefficient of the same power product on the left. We write then c’ = c(U)/d(U), where c, d are polynomials having no common factors (other than units, of course). Next observe that each factor of the product [T is irreducible in O[Xp, X;, too.-+-y Usa]; for since it is homogeneous of degree 1 in Xp, X, and homogeneous of degree 1 in tg, ..- Ugg, it could, at worst, factor into a polynomial in Xo, X, alone times a polynomial in Ugg, +» Ugg alone; but then the coefficients of X, and of X, would differ only by a factor in C, in particular, would involve the same w,,, and this is not so. Finally, observe that on the left only powers of the determinant |U| occur in the denominator (since this is true for the entries of U-). Crossmultiplying and using the unique factorization theorem, one sees that d(u) is a factor of [U|*, for some e. Hence c’ can be written as c,(U)/|U |*, ¢, a polynomial. Now we specialize U > A. Observe that the operations of forming the resultant and of specializing taken in either order give the same result. (This observation depends on our agreement to give F and @ formal degrees in X, equal, respectively, to their total degrees in Xo, X,, Xj. See p. 44.) 46 BEZOUT’S THEOREM Then we get oA) |Al° where (a, xf, xf) are the coordinates of P#. Here X,x4 — X,r4 and X,x4 — Xyef are nonassociates for i # j because I’, A, are in permissible position. Hence Rx( F(A7X), (ATK) = Tl as —Xerhl, iy, Ags PP) = Py, Ags Pf), and the proof is complete. The main object of the preceding proof was to establish the projective invariance of the intersection multiplicity, but it shows quite generally that on specializing U to A, the i(I'y, Ay; PY) factors of Ry,( F (UX), G(U-~X)) corresponding to PY go over into that number of factors of Ry (F(A“X), G(A~X)) corresponding to PA; the curves I'y, Ay need not be in permissible position, though Ey is not to lie on both of them. Hence we get: Theorem 7.5. Assume that By: (0, 0, 1) does not lie on both T: F =0 and A: @ = Oand that T and A are without common component. Then the multiplicity with which Xyr_ — Xr occurs in Ry (F, @) is the sum of the intersection multiplicities along the line Xyvq — Xyr, = 0. Exercises 1, Let SUF) = ag¥™ + YM Eo, g(¥)=¥ —-b. Show that in this case R(f, 9) = +f(). 2. Let I’ have at P an ordinary cusp, and let A have a simple point there. Assume that I’ and A have a common tangent at P. Show that i(I', A; P) = 3 (and not more!). 3. Prove that ¢(U) 6|U |4, € eC (see p. 45). Now we have a definition of intersection multiplicity (at a point) of two curves I’ and A and, in particular, for A = L, of a curve I’ and a line L. We already had a definition for the latter case, but we see that the two definitions coincide in that case: for let TP: F(Xy X,, X,) = 0, L:X,—mX,=0. According to one definition, i(I', L; H,) equals the maximum power of X, that can be factored from F(X, X,, mX,); according to the other, it equals the maximum power of X, factorable from Ry(F(Xo, X1, X,), X, — mX,). BEZOUT’S THEOREM 47 Hence by Exercise 1 above, the two definitions coincide. Since all of our results concerning singularities follow from the theorem that the intersection multiplicity of F(X, Y) = 0 and Y = mX at E, is the maximum power of X that can be factored from F(X, mX) and since we have obtained this result starting from the second definition of intersection multiplicity we may now simply abandon the first definition and then no longer need to prove Theorems 7.1 and 7.2. Exercise. Prove Theorems 7.1 and 7.2. Chapter 8 THE BASIC INEQUALITY Theorem 8.1. Let T and A be two curves without common components, each containing P as a simple point. Then i(I', A; P) > 1 and equality holds if and only if the tangents to T and A at P are distinct. Let F(X, Y) = 0 and G(X, Y) = 0 be the minimal equations of I and A. We may suppose that I’ and A are in permissible position and that P = (0,0). We write F(X, ¥) = aX +ay¥ +++, AX, Y) = bX + Og¥ toe, ayyX fre dg fee Expanding the determinant by the first two columns according to Laplace’s rule,* we find that one of the terms will be Mobo ay bos bye bots by te The other four terms, as one sees, have X* as factor. Thus one obtains x + cofactor. Ayo Aor bd + cofactor + X%(+++), bio bo whence the first point, ie. ([, A; P) >1. If the tangents are equal, then @9by, — 41049 = 0, and then ur, A; P)> 1. If the tangents are distinct, then 4499, — 4,9 # 0. We shall show that the cofactor mentioned above is not divisible by X, i.e., that evaluated at 48 THE BASIC INEQUALITY 49 0, it is 4 0. The cofactor evaluated at 0 is the resultant of the polynomials F(0, Y)/¥, @(0, Y)/¥: F(0, Y) (0, Y) a (292, 22.0)» Now F(0, Y) = 0 gives the points of intersection of I’ with X = 0; the same is true for (0, Y) = 0. Thus F(0, Y)/Y and (0, Y)/¥ have no root in common except possibly 0, because I’ and A are in permissible position; but at least one of F(0, Y)/¥, G(0, Y)/Y¥ does not have 0 as a root (since at. least one of the tangents is not vertical). Hence 2(*5 Y) Go, ”) gt Peery Exercise. Let I" have an ordinary double point at P and let P be simple for A. Show that i(I', A; P) > 2 and = 2 if and only if the tangent at P to A is distinct from the tangents to T' at P. Generalizing Theorem 8.1, we have: Theorem 8,2, Let T and A be curves without common components and let the point P be an exactly r-fold point for T and an exactly s-fold point for A. Then aT, A; P)>rs, and inequality holds if and only if T and A have a common tangent at P. Remark. If one fills out the determinant % G+ + Oy My My + + Om by bp tt ty by bb ty by by bn with the letters a_,, a», ...,b, b-g,-.. in the form Mg Am Amy * 8 t Omanmt a, % a ae at Omens aoe gett St Opies Dy atts Da tuerieec eeaeiee cco bya by bes tt bmanae 50 THE BASIC INEQUALITY then, just as in the proof of 4.5, one sees that this last determinant is isobaric of weight mn. The same is true if we assign a, the weight m — i and b, the weight n — j; see the remark at the end of Chapter 4. We will prove 8.2 for r = 2, s = 3, though the method is the same in the general case. We may suppose that I’ and A are in permissible position and that P = (0,0). We write F = dyX? + ayXY + ag:¥? +>, G = dypX? + by X*Y + byXY? + bog¥* +--+ Then Ry(F,@) = [yg X2 fe ayX fees agg fre Ogg frre Gogo ee eres gee ae Meege eT e laa teie ny dygX* poe dK foe gy bot oe FE bak? foes Bak poe bygK bee by bee We make a Laplace expansion by the first r + s (=2 + 3) columns. One of the (r + 8) x (r + 8) determinants is formed from the first s rows of the a’s and the first r rows of the b’s, namely, AygX? fv dy X free gg ttt Meg tt Mog tee yyX* + My X tes gg +0 Og + QygX? + +++ ay X + on + By X28 foes By XP Hees ByeX pees by fees Oy tere ByoX8 pore ByX*® foes byK fee byy bos If we give the entry a,X? + +++ the weight 2, a,X + -+- the weight 1, etc., as in the initial remark, we see that weight is < subdegree for each entry and hence that the weight (= rs) of the determinant is < the subdegree of the determinant. So rs is < the subdegree of the determinant. Moreover, if weight is < subdegree of some entry, then weight < subdegree of any term containing that entry as a factor. For example, weight (ay, + +++) = —1 <0 < subdegree (a5 + ---). Hence any term in the expansion of the determinant containing ay. + ** as a factor has subdegree > rs; and similarly for the other terms. Hence the above determinant is equal to AygX? 8+ ayX f-e* dog fee GyX* $+ aX foe panies bygX8 fete bygXB foes byX pore ByK8 oes by KE fo THE BASIC INEQUALITY 51 In this determinant the term cX* comes from the lowest-degree terms of the various entries. Hence the coefficient of X® is [42 1 doe Aq M1 Fon | xq Ay, Aon | > bap ba Pie Pog S59 Bay Bye os which is R(R,, G,), and which vanishes if and only if the homogeneous polynomials F,, @, have a common factor. We have to examine the other subdeterminants of the first r+ s columns. Let us take s rows of the a’s not the first 8, say rows i,, ig, ..., i, with i, subdegree of the corresponding elements of row j, and inequality holds at least once. Hence the subdegree of any such subdeterminant is >rs. We still have to consider taking s + k rows of a’s and r + k rows of b’s, say the first s + 1 of the a’s and the first r — 1 of the b’s. We recall that we give to zero every degree. Giving the appropriate degrees to the zeros, we find that the subdegrees for row s + 1 of the a’s are: pce eee The subdegrees for row r of the b’s are: Sere meer rere Hence here too, we replace elements by elements having greater subdegree, and hence the subdeterminant is = 0(X’*+!), The same is true for the re- maining subdeterminants. Resuming, we have Ry(F, @) = R(F,, G,)X"* + cofactor + XP(-++), In every case, then, i(I', A; P) > rs. The cofactor evaluated at X = 0 is R(F(0, ¥)/¥*, G0, Y)/¥*). Since I’, A are in permissible position, F(0, Y)/¥* and G(0, Y)/Y¥* have no common roots other than 0. Choosing the coordinate system conveniently, we may suppose that the Y-axis is not tangent either to [or to A at P = (0,0). In that case, F(0, Y) is exactly divisible by Yt and G0, Y) by ¥*, whence neither F(0, Y)/¥* nor G(0, Y)/¥* has 0 as a root, and hence R(F(0, Y)/¥", 40, Y)/¥*) 40. Consequently i(I', A; P) = rs if and only if R(F,, G,) # 0, ie., ifand only if T’and A have no common tangent at P. 52 THE BASIC INEQUALITY Theorem 8.3. If T', A are two cycles having no common component with a third cycle E, then ul + A, £; P) =i, £; P) + (A, E; P). This theorem is an immediate consequence of the following: Theorem 8.4, If f,,h, are polynomials in Y with indeterminate coefficients, then R(f(Y)9(¥), MY)) = RPL), MY)) + R(g(Y), AY). Thus let. L(Y) = ag¥™ +--+, WY) = HY" $00, WY) = c¥? + +++, a,b; cy indeterminates. We shall need first: Lemma 8.5. R(f, 9) is irreducible in Clg, ... , Bg... J. We first prove 8.4 using 8.5. Let f*, g*, h* be, respectively, the homo- geneous polynomials corresponding to f, g, h (then, for example, we have R(f*, h*) = R(f, h)]. Applying 4.3, we find that if for certain values Ogee Ogres Cg the resultant R(f, 4) vanishes, then also R(fg, h) vanishes, since if f*, h* have a common solution [4(0, 0)], then so do f*g*, 7 Hence by Theorem 3.15, in Cg, ..- 5 Boy «+ +5 Cos «+ R(fg, h) = O(R(f, h)). Also R(fg, h) = 0(R(g, h)) and since C[ay,...,b9,-++,¢y,++-] is a unique factorization domain and since R(f, h), R(g, h) (which are distinct) are irreducible by 8.5, one has R(fg, h) = ORS, b)R(g, h)). Then let R(fg, h) = HRS, W)R(g, h). The degree of R(f, h) in the cy is m and that of R(g, h) in the o, is m. Moreover the degree of (fg, h) in the c, is m + . Hence H cannot involve the ¢,. A similar argument holds for the a,, b;. Hence H is a constant; and moreover THE BASIC INEQUALITY 53 this constant is 1, but we omit the proof as this fact is not vital for the moment. Finally we prove 8.5. Let R(f, g) = P-+Q+R-... be a complete factorization of R(f, g). Making an induction on deg f + deg g, after checking 8.5 for deg f = 1 and deg g = 1, we assume (as inductive hypo- thesis) that R(f, g,) is irreducible, where g, = b,¥"-! + -++ Since ay and R(f, g,) are irreducible, and R(f, 9)|,,-0 = 4oR(f, gx), we must have P|,,-0 = 4% * const or = R(f, 9) - const or = a9R(f, 9) + const or =const (const £ 0), and the same may be said of Q|,,-9, -.. Hence the complete factorization of R(f, g) has the form (1) RF, 9) = [ao + bol + YILR(F, 92) + bo(-- +) X [const + b9(-+*)] +++ [const + bo(-++)], or possibly (2) RUF, 9) = [aoR(f, 91) + bo(+**)) X [const + By(+++)] +++ [const + Bg(-+ ye Since R(f, g) is homogeneous in the b,, the same is true for each factor and hence the coefficient of by in each factor [const -+ b,(-+-)] must be zero. In other words, except for units (constants # 0), one cannot have such factors. Now in case (2) one sees that K(f, g) is irreducible (since there is only one factor in the complete factorization). In case (1), as argued before, the coefficient of by in [ay + bo(++*)] is zero. Hence R(f, g) has ay as factor. But this is certainly not so as R(f, g)|o # 0- (This last point already occurred previously in the argument, when we used R(f, 9)|p,-0 # 0-] Exercises 1, Let I’ bea curve, F(X, X;, X;) = 0 its minimal equation, and a = (ay, 4, 43) a point. The cycle oF oF oF A:ag— +a,— +4 2— =0 eX ax, eX, is the polar of a with respect to I’ (see 6.8). By 6.10 A (if it is defined) meets T in only a finite number of points. Show that if I has as singularities only ordinary double points, d in number, and ordinary cusps, k in number, then if a is not on I, A intersects I’ at its singularities with a total multiplicity of 2d + 3k at least. Hence if ais not on any tangent to I’ at a singular point, then from a one can draw at most n(n — 1) — 2d — 3k tangents to I. 54 THE BASIC INEQUALITY 2. Let S(X) = a(X — XX — Xy)+-- (KX — Xp) g(X) = b(X — F(X — ¥y)-+ (KX — Fy), where dp, by, Xy,.+., Yq are indeterminates. Show that R(X), 9(X)) = agogll(X;, — ¥,) = afl g(Xy). ia j Following this up, one could get another proof that Riffs 9) = Rf 9) > Rhy 9) Chapter 9 CUBICS Definition. By an ordinary point of inflection of a curve one means a simple point P of the curve such that the tangent to the curve at P has intersection multiplicity 3 with the curve at P. If the intersection multiplicity is 2, then the simple point P is called ordinary. The following theorem is immediate: Theorem 9.1. Every simple point of an irreducible cubic is either an ordinary simple point or an ordinary point of inflection. Lemma 9.2. Let P be a simple point of curve T and let L be the tangent toT at P. Let K be a curve. Then if one of the numbers i(K,T; P), i(K, L; P) is less than i(D', L; P), 80 is the other, and thé two are equal. This lemma holds for curves I’ of arbitrary degree, but for the moment. we will prove it only for cubics. There are three cases: a) (I', L; P) = 2, ie., P is an ordinary simple point; b) (IP, L; P) = 3, i. c) (I, L; P) = co, whereby we mean that L is a component of I’. ., P is an ordinary point of inflection; Case (c) is trivial. Take P: (1, 0, 0), L: ¥ = 0, and T: F = 0, K:G@ =0 in permissible position, By 4.1.1 R(F,@) = AF + BG: placing Y = 0, one sees that i(K,T; P) >min {i(, L; P), (K, L; Py}, so that if (K, L; P) >i(0', L; P), then (K,T; P) >i, L; P). It thus remains to show that (K,L; P) n, and zero if m n. Taking a permissible coordinate system for @ = 0 and F = 0 (without changing X, = 0), we note that the system is also permissible for Xj, = 0 and F = 0, since the same intersections are involved. To complete the proof we will use: Lemma 9.4. Let f,g, hbe polynomials in ¥ with indeterminate coefficients. Then R(f, g + hf) is divisible by Rf, 9). Supposing that the lemma is known, we take f = F(X,, X;, X,) and g = G(X, X;,Xq) with X, = Y and find R(F,X,G,) = R(F,G — HF) = 0 (R(F, G)). Since both resultants have degree mn, R(F, X_G,) = const - RUF, G). Hence the intersections of @ = 0 and F = 0 are the same, including multi- plicities, as those of X,G, = 0 and F = 0, and since deg @, = m — 1, the proof is complete. Proof of the lemma. R(f(¥), 9(Y) + WM Y)f(¥)) vanishes (for special values of the coefficients a;, b; of f, g) whenever R(f(¥), g(¥)) does. Since R(f, 9) is irreducible by 8.5, the first resultant is divisible by the second. Theorem 9.5. An irreducible cubic has at most one singularity. The proof is trivial. Notation. Let 8, T be sets. By S x T, the so-called cartesian product of S and T, one means the set of (ordered) pairs (P, Q) with P in S and Q in T. Fig. 9.1, Theorem 9.6. Let T be an irreducible cubic without singularities and Oa point on I. Let A and B be arbitrary on T. We define an operation T x P+ Tas follows (see Fig. 9.1): the line determined by AB (or the 58 cUBICS tangent in the case A = B) cuts T ina third point R. The line OR cuts T in a third point which we will call A + B. Then V with this operation becomes a commutative group. Commutativity, the existence of a neutral element (namely, 0), and the solvability of the equation P + X = Gare all immediate. All that remains is to establish associativity. Examining Fig. 9.2, one sees that the problem of proving that (4+ B)+0C=4+(B+0) comes to proving that A, B + C, and Q are collinear. Consider the cubic K = ABP + 0(A + BQ + OB + O)R. The intersection of K and I’ (counting multiplicities) consists of nine points: KT ={A, B, P,C, A + B,Q,0, B+C, RB}. Now the points C, B, Rare collinear. These are three of the nine mentioned points. Hence by Theorem 9.3, the remaining six, i.e., A, P, A+B, Q, 90, B+e are cut out on I’ by a conic. Of these six, 0, A + B, Pare collinear. Hence, again by Theorem 9.3, A, B + C, Q are collinear. Q.E.D. In the case that the cubic I’ has a singular point D one can define a sum in exactly the same way on the set I’ — {D}, and I’ — {D} becomes an abelian group. Exercises 1, Let 0, 1, 00 be three points on an irreducible conic IT. Then T — {00} can be made into a commutative group by defining X + ¥ as in Fig. 9.3; and T’ — {0, 00} can be made into a commutative group by defining X- ¥ as in Fig. 9.4. CUBICS 59 Fig, 9.3 Fig. 9.4 2. Let 0 be the zero of the group G on a nonsingular cubie I, If A, B, C are collinear (i.e., the complete set of intersections of some line with T), then A+ B+C = K, where K is the third point of intersection with I’ of the tangent at 0. Conversely, if A + B + C = K, then A, B, C are collinear. The points of inflection of I’ are given by the solutions of 3X = K. 3. Let G, I’ be as in the previous exercise. Then there is a point X on I’ such that 0, X, 2X are collinear, and hence I’ contains at least one point of inflection (namely X). 4. There is a line through any two given points in the plane, a conic through any five, a cubic through any nine, and a quartic through any fourteen. Theorem 9.7. Let 0 be the zero of the group @ on a nonsingular cubic T and let H = 0, where deg H = m, cut out Py,..., Py, on I. Then P, +:+* + Pom = mK, where K is the third point of intersection with T on the tangent at 0. Proof. The proof is by induction on m and is checked immediately from the definition of @ form = 1. Let P,P, cut I in a third point R and let P,P, cut Tin S. Let RS cut Tin 7; R4+84+T7=K. Then Py, P,..., Pam, R,S, T are cut out by an (m + 1)-ic (i.e., a curve of order m + 1). Since P,, P,, Rare collinear, Ps, Py,..., Pan, S, T are cut out by anm-ic; and since Ps, Py, S are collinear, Ps, Ps,-.., Pam, T are cut out by an (m — l)-ic. Hence P, + Py -+ +++ Pam + T = (m — 1)K; using P, + P,+ R= K and P, + P, + 8 = K, we get (Py + P, + BR) + (Ps + Py +8) + (Ps + Po +++ + Pam + T) = (m+ 1)K. Now using R + S + 7 = K, we get the desired result. The converse of 9.7 is proved similarly. 60 CUBICS RATIONAL FUNCTIONS ON AN IRREDUCIBLE CURVE Let I’ be an irreducible affine curve, F(X, Y) = 0 its minimal equation. By a rational function on T' we mean, at least roughly speaking, a function given by a rational expression a(z, y)/b(e, y), where (x, y) varies over I’ and a(X, Y), b(X, Y) are polynomials in C[X, Y]. If the expression is to be meaningful for even one point (z, y), we must have &(X, Y) 4 0(F(X, Y)), and we impose this condition; the function is then defined at all but perhaps a finite number of points, namely, the points where b(X, Y) = 0 intersects F(X, Y) = 0. However, we wish to allow ourselves to rewrite the expression a(e, y)/b(x, y) taking into account relations which hold for a point (z, y) in general position on I’. For example, let I" be the curve ¥ = X?, and consider the expression y/x. This expression is undefined at (0, 0), but if we write yz =, which holds for point in general position, we will be able to say that the rational function y/z is defined at (0, 0), and that its value there is 0. In order, then, to define rational function with appropriate precision, we first define a representative of a rational function on T’ as a function given by a rational expression a(z, y)/b(x, y) where (x, y) varies over I’ and a(X, Y), 0(X, Y)eC[X, ¥] with 6(X, Y) # 0(F); this function is defined at the points (x, y) of I’ for which b(z, y) # 0, and nowhere else. ‘Two representatives a/b, o/d are called equivalent if they are equal wherever both are defined, i.e., for points (x, y) on I’ for which (x, y)d(x, y) # 0. Note that a/b and c/d are equivalent if and only if a(X, ¥)d(X, Y) — 0(X, Y)c(X, Y) = O(F); in fact, if a/b and cfd are equivalent, then a(z, y)d(z, y) — b(a, yo(x, y) = 0 except at a finite num- ber of points, so a(X,¥)d(X, ¥) — (X, ¥)c(X,¥) = O(F). Conversely, if this congruence holds, then a(z, y)d(x,y) — B(x, y)c(x, y) = 0 for all points of I. The same argument shows that a/b and c/d are equivalent if they are equal at infinitely many points of Y'. It follows that if a/b is equivalent to ¢/d and o/d is equivalent to e/f, then a/b is equivalent to ¢/f, for all three will be equal where bdf 4 0. We then define a rational function on Tas an equivalence class of representations of a rational function. The value of a rational function r at any point (x, y) is the value of its representatives at (x, y). Since none of the representatives may be defined at (2, y), it is possible for r not to be defined at (x, y); for example, if: ¥ — X? = 0, then 2/y is not defined at (0, 0). Rational functions are added, subtracted, multiplied, and divided via their representatives. The situation for an irreducible projective curve I’ with minimal equation F(X, X,, Xq) = 0 is quite the same except that in the definition of a repre- sentative a/b one requires a, 6 to be homogeneous of the same degree. Now let I', A be two irreducible affine curves; the curves need not be distinct or even in the same plane, in fact, for notational purposes it is RATIONAL FUNCTIONS ON AN IRREDUCIBLE CURVE 61 convenient to think of them as being in distinct planes and then write their minimal equations as F(X, Y) = 0, G(U, V) = 0. We now wish to define a rational function for the variable (x, y; u, v), where (x, y) is on T’ and (u, v) on A; or, in other words, we want to define a rational function on T x A. We need: Lemma 9.8. Let T: F(X, ¥)=0, A:G(U,V) =0 be irreducible curves. If H,(X,¥; U,V)+H,(X,¥; U,V) vanishes on Tx A, where H,, H, are polynomials, then H, or Hy vanishes on T x A. Proof. Assume that neither H, nor H, vanishes on I. x A. Then there is a point (x,y; u,v) in’ x A where H, does not vanish. Thus there is a point (u, v) on A such that H,(X, Y; u, v) does not vanish on I’. Similarly, there is a point (u’, v') on A such that H,(X, Y; w’, v') does not vanish on I. Hence H,(X, Y; u, »)H,(X, ¥; wv’, v’) does not vanish on I’. Hence there is a point (x,y) on I’ such that H(z, y; w, v)* Hale, y; w’, v') #0. Then H,(z,y; u,v) £0, so H,(z,y; U,V) does not vanish on A, and for a similar reason H,(z, y; U, V) does not vanish on A. Hence it follows that Hx, y; U, V)+ H,(x, y; U, V) does not vanish on A. Hence there is a point (u,v) on A such that H,(z, y; u, v)+ Hy(x, y; u,v) 40. This contradicts the assumption that H,H, vanishes on I x A. The definition of a rational function on T x A can now be given: one proceeds very much as with the definition of a rational function on I. A representative of a rational function is a function given by an expression a(x, y; u, v)/b(x, y; w, v), where a, 6 are polynomials and 6 does not vanish everywhere on I’ x A. Two representatives a/b, o[d are said to be equivalent if ad — be vanishes over! x A. The proof that if a/b and fd are equivalent and if old and e/f are equivalent, then a/b and e/f are equivalent is slightly different, since a polynomial 6 may vanish at infinitely many points of T’ x A without vanishing everywhere: assuming ad — bc and of — de vanish on I’ x A, one finds that adf — bde (=f(ad — be) + Blof — de)) vanishes on I’ x A, and since d does not, af — be does, by the lemma. Now we define a rational function on Tx A as an equivalence class of representatives. If alb ~a'jb’ (i.e. a/b is equivalent to a’/b') and cfd ~c'/d’, then aclod ~ a’c'[b'd’. For the proof, one first shows ac/bd ~ ac’/bd’ and then ac'|bd’ ~ a'c'[b'd’, whence ac/bd ~ a’c'/b'd’. Hence the product of rational functions can be defined via their representatives. Similarly for sum, difference, and quotient. A similar situation obtains for irreducible projective curves I’, A, given by F(X», X;,X;) = 0, GU, U,, Uz) = 0. Here a, b should be homogeneous of the same degree in Xo, X,, X, and homogeneous of the same degree in Us, Uy, Uy. In the following theorem we take T = A. 62 cUBICS Theorem 9.9. Let T be an irreducible affine curve with F(X, Y) =0 as minimal equation. Let (1, y1), (te; Yo) be points on I’. Then the rational function given by (yo — y:)[\% — %) (which, of course, is the slope of the line joining the points if a, # 2,) is still defined for (x, Y) = (ta, Y2) provided that (x, y,) is a simple point of T with nonvertical tangent; and, in that case, the value of the function at (ar, y,) is the slope of the tangent to T at (x1, 1)- Proof. By the computational rules given in Chapter 6, 0 = P(ts ys) — F(t xn) = (Flea yo) — F(a ys)) + (Fn y2) — F(t m)) ar + (ey — PE A+ (Ye 0) 55 ave) aoe oF = 8) Kaya) + (Y2 — wi)" 1 where the [+ -] stand for the polynomial expressions in 2, y;, 2, Yp- Then ar | t+ @—a team Xba ty — a, ar : ay + (Ye — wl] Ven The theorem follows from this way of rewriting (yp — y,)/(, — 2). Let I’, A be irreducible affine curves. By an obvious generalization of rational functions, we say that a point R (varying in an affine plane) is a rational function of points P € A, Qe A if the coordinates of R are rational functions of the coordinates of P and of Q. We frequently write R : (s, t), where s, ¢ are representatives of the functions in question. The situation for projective curves is quite the same. Here P: (29, 2, 23), Q: (tg, ty, Ug), Ri (ro 7, 72), and ro, 7%, 72 are to be rational functions of P, Q, at least one of which does not = 0, ie., does not vanish every- where on IT x A. We say that R = R(P, Q) is defined at P, Q if for some r, that does not vanish on I’ x A, the functions rofr,, ry/r;, ral; are all defined at P,Q. [If this happens for two values of i, say i = 1, 2, then the points obtained are the same. In fact, one first observes that if g and h are rational functions defined at P,Q, then so is gh, and the value of gh at P,Q is the product of the values of g and h at P,Q. Then the values of r4/r, and r4[ryare both + 0, as the product of these values = 1. The assertion now follows upon observing that (rs/r,)/(ra/r2) = r[rg and that (n/n)lagllraln lao = (n/re)|B,9-] By a complete set of representatives for a function r we mean a set of representatives of 7 such that at least one of them is defined at any point (of PNC or of PNC x PNC) at which r is defined. RATIONAL FUNCTIONS ON AN IRREDUCIBLE CURVE 63 Let w,,..., wy be indeterminates over C and let C, = C(w, ..., wy). We wish to define a rational function on T’, or on I’ x A, over C,. The in- dependent variables will be the same as before, but the values of the functions may involve the w,. The procedure is quite the same as before once we have the following: Lemma 9.10. Let F be an irreducible polynomial in C[X, Y]. Then F remains irreducible in C,[X, Y]. Proof. This can also be formulated: if Hy, Hy € CX, Y]and H,H, = 0(F) in O,[X, Y], then H, or Hy = 0(F) inC,[X, Y]. The H, may have denomina- tors involving the w,, but this is of no importance, and we may suppose the H, are in Ofw,,...,wy; X, ¥]; we consider H,, H, as polynomials in the w; with coefficients in C[X, Y]. If HH, = O(F), then Ay(wy . +5 ni ty) * Haly --- sn 2 y) = 0 for every point (x, y) (in ANC) on F = 0. Then Hy(w,,...,W,3 %, y) = 0 or Hy(wy,..., 0,3 2, y) = 0. Thus for any (x, y) on F = 0, all the co- efficients of H, or all those of Hy vanish. Hence the product a,(X, Y)a,(X, Y) of any coefficient a, of H, and any coefficient a, of H, vanishes everywhere on F = 0, whence a,a, = 0(F). If H,# 0(F), then some coefficient a, of H, is 4 0(F), whence every coefficient of H, is = 0(F). Q.E.D. Theorem 9.11. Let T be a cubic curve in the projective plane free of singularities. Let P,, P, be points on T' and let P, be the third intersection of P,P, with T. (As usual, we understand P,P, to be the tangent at P, if P, = P,). Then Py is an everywhere (on T' x 1’) defined rational function of Py, Px. Proof. Let P;: (x\, x, xf), i= 1, 2, 3. Let X,=0 be taken as the line at infinity; and consider for a moment the case that the P; are at finite distance. Let (1, 2/26, x{/al?) = (1, 2, y;) be the normalized coordinates of the P,;. Assume, too, that P,P, is not vertical and even, at first, that P,# P,, Then the slope of P,P, is m = (yz — y,)/(t, — 2), and the equation of P,P, is Y — y= m(X — Let F(X, Y) =0 be the minimal equation for the affine part of I’. Eliminating Y, we find a cubic in X whose roots are 2, 2, %3: let this equation be aX* + a,X? + aX + a, = where the a; are polynomial expressions in m, 21, yy. Here a) + 0, since the equation has 3 roots, and —a,/ay = 2, + %_ + 25. Solving for x, we find that 2, is a rational function of 2, 41, %2, Ya; the same now holds for ys from Ys = mtg — 2,) + yy. Thus Py, for the case considered, is a rational fune- tion of P,, Py. Moreover, by 9.9, Py is still defined for P, = P,, provided the tangent at P, is not vertical. 64 CUBICS The possibility that P,P, is vertical and the possibility that P,, P,, or P, (or some two or all three of them) are at infinity require further considera- tion. As in Bezout’s Theorem (7.3), we introduce nine indeterminates Uo +++» Uae over C and consider the linear transformation Ly. Let PY: (yf, y®, y?), where y®) = Ux), be the transform of P;. As explained for 7.3, P? PY does not pass through Z, : (0, 0, 1); @ similar argument shows that PY cannot be at infinity (if it were, then P4 would be, too, for any nonsingular matrix A with entries from C, and this is clearly absurd). If I’: F(X) = 0, then the PY lie on F(U-1X) = 0, and by the considerations given just above, PY: (1, y?/y®, y/y) is an everywhere (on I’ x I’) defined rational function of P,, P,. Note that we need only two representa- tions of PY to bring this definition to expression: one for the case that P, # P,, in which one writes the slope of PY PY in the obvious (i.e., universal) manner, and one for the case that P, = P,, in which one writes the slope as in 9.9. Applying U-} to these representations, we get two representations (Ag, Ay, Ag), (Bo, By, Ba) of Ps: (x, x), 29). Here A; and B; are repre- sentatives of the same function, i = 1, 2, 3. The proof would be complete, except that the A,, B; involve the U,,. To complete the proof, we first establish : Lemma 9.12. Let r be a rational function on T x A (say) over O, = O(w,, ..., wy) and assume that (P,Q) €O whenever defined. Let Ty, +++ 5% be @ complete set of representatives of r; each r; can be written as De tem ke KEKE n= et, Latins where the t, are power products of the w; and the ayy, by: are polynomial expressions in P,Q (PET, Qe A). Consider all the ay;[by; with by; # 0. Then these form a set of representatives for a function r' on Tx A over C defined wherever r is; and r'(P,Q) = r(P, Q) wherever r is defined. Proof. Consider one of the a;/byi, 84Y a4,/b,,- We say that 4 (X fib) — bu( hin) = 0 onI x A. In fact, for any point P,Q on I x A, if ¥ tibi(P, Q) = 0, then by(P, @) = 0 and, in this case, the equality follows; if ¥ tuba (P,Q) 4 0, then there is a number # € C such that ? ¥ tibia(P, Q) — > titia(P, Q) = 0, whence 7b,,(P, Q) — a,(P, @) = 0, and again the equality follows. Thus ,,/6,; is a representative of r. As this holds for all the a,,/b,;, they are all equivalent and represent a function r’ over C. If, now, r is defined at P,Q, then ¥ t,b,(P,Q) # 0 for some i, say i = 1, Then forsome k, b,(P,Q) 4 0, say k = 1. Then ay/b,, is defined at P, Q, and the proof is complete. ABELIAN VARIETIES 65 To complete the proof of 9.11, note that 4», 4;, Az, and not merely their ratios, are defined for P, # P, and have values not all = 0. Let a, k=1,...,gi, be the complete set of representatives of A; given by the lemma. Then the g.g,g2 representations (a, a(*, aff), where k; runs from 1 to g,, are over C; and for any P, P, in x I’ with P; # Pz, at least one of these yields P,. These g,go9. representations, together with a similar set derived from By, B,, By, yield a point (ry, ry, 73) which is everywhere defined on I’ x I’ and which for any P,, P, has the value Py. Remark. Let T' be given by an irreducible equation F(X) = 0. As in 7.3, F =0 defines a curve I* in PNC,, where C, is the algebraic closure of O, = C(Uq, ..., Ugg). To define rational functions on '* we must know that it is irreducible. It is a property of any irreducible polynomial in C[X] to remain irreducible over any extension field K of C (ie., in K[X]); this is its so-called absolute irreducibility. In the case of a curve : F = 0 without singularities, a simple proof of this can be given by observing that cannot acquire singularities by being regarded over an (algebraically closed) extension field of C. We could, then, have defined rational functions on I'* for our cubic F = 0; but, in fact, this was not necessary. ABELIAN VARIETIES Let I’ be a cubic curve without singularities in the projective plane and let a line cut P'in the points P;, P,, P. Then, as we have just seen, P, is an every- where (on I’ x T’) defined rational function of P;, Pz. From this one sees that in the group defined on I one has: a) in P+ Q= R, R is an everywhere (on I’ x T’) defined rational function of P, Q, and b) in P + Q@=0 (the neutral element), Q is an everywhere (on I’) defined rational function of P. We will now make some observations with the aim of relating the above example to a branch of algebraic geometry that has attracted much attention, namely, abelian varieties. We need some definitions. An affine variety is the locus of a point satisfying a finite number of simultaneous polynomial equations: F,(Y,,...,¥,) =0,-..) Fi(Yus---+¥n) =0, F,eC{Y;,...,Yal (where F,eC is allowed). A projective variety is similarly defined by homogeneous polynomials @,(X,,.-.,Xn)- ‘A variety V (affine or projective) is called reducible if V = V,U Va» VV, V#V_ where V,, V_ are varieties (affine or projective according as V is affine or projective). A variety is called irreducible if it is not reducible. It would be convenient to introduce two more terms: abstract variety and complete variety. We will not give the general definitions, but will limit ourselves to some special cases. 66 cUBICS By a special abstract variety we mean an irreducible variety (projective or affine) minus a proper subvariety. Thus every special abstract variety has the form V — F with V, F varieties and V irreducible. By a special complete variety we mean an irreducible projective variety. Thus every special complete variety is a special abstract variety of the form V — F with V projective and F empty. By a special group variety V we mean a special abstract variety that is also a group with the conditions: a) in P-Q = R, Ris an everywhere (on V x V) defined rational func- tion of P, Q, and b) in P-Q= E (the neutral element), Q is an everywhere (on V) defined rational function of P. The main point of this definition is that the group operations are not to be taken arbitrarily, but must be defined algebraically. A special group variety need not be commutative. For example, let us consider the points (%, tor, so tx) in projective 3-space such that Ugorls, — %ro%o, # 0; these form a special abstract variety. It can be made into a group according to the rule. (%oos Mors “ro M2) * (Yoo Yor P10» Y21) = (Moo Yor» M10» Mrs) ee ) e >) : e =) Wy Wy, Uo Un, 0 Yn, (ordinary matrix multiplication). ‘A nonsingular cubic made into a group is a special complete group variety. There is a known theorem assuring us that a special complete group variety is necessarily commutative, and because of this the special complete group varieties are called special abelian varieties. Thus it was no accident that in our example of the cubic there was commutativity. In fact, it is impossible to make a nonsingular cubic into & noncommutative group variety. where Chapter 10 CUBICS (CONTINUED) POINTS OF INFLECTION Let I': F(X», X;, X;) =0 be an irreducible curve of order n. Our first question is: Does I’ necessarily have a point of inflection (ie., a simple point whose tangent to I’ cuts P there with multiplicity > 3)? If I’ is a line, then every point of I’ is a point of inflection; if I’ is a conic, then no point is a point of inflection: hence we will assume ord I’ > 3. If x, y are two points, then (as in Chapter 6) oF 1 er F(x + dy) = F(x) + (Sz nat sil eoan ) ae an (Here @F/éx; is a standard notation for @F/@X;|x_,; and similarly for @F/éx,0x,.) If xis onT, then F(x) = 0; ifitissimple, then > (@F/ax,)X; = Ois a line, the tangent line. If y is on this line, then (2F/éz,)y, = 0; if the tangent line is to cut with multiplicity > 3, then also @F —— yy; = 0, > tage, ice., y lies on the locus OF z x,0x; As this is to hold for each point of 3(aF/éx,)X; = 0, we find: Theorem 10.1. The simple point x of I: F = 0 is a point of inflection of T if and only if OF oF Daa eh o(sZx,). In particular, therefore, if x is a point of inflection then the conic er 2 baee, XX, =0. X,X,=0 67 68 CUBICS (CONTINUED) must be reducible. The condition for this is that oF det lea an, or, in other words: Corollary 10.2. A point x of T’ is a point of inflection only if it lies on the curve H = 0, where @F @F @F @X2 aX0X, eX,0X, er @8F @F @X,0X, OX? 8X,0X, |" er er er @X,0X, @X,0X, ax? We will see in a moment that not every point of Tis a point of inflection. Hence the above determinant is not the polynomial 0, and the use of the word “curve” is justified. Hence also not all 2F/2x,z, can vanish over I’, which justifies the word “conic” in our reference to the locus er Xx: 2 tee, ey =0. Corollary 10.3. Conversely, if x is a simple point of I on H = 0 (the so-called Hessian of I’), then x is a point of inflection of T. Proof. Multiplying the first column of H by X, and adding to it X, times the second and X, times the third, and applying Euler’s formula, one finds yok ar _ ar ("—D5x oxox, Ox0x, oF @F @F Re eye ee 7 ™—Use axe oxax, | (ey) eR eEOU aX, 0X0X, x3 Applying a similar operation to the rows of this determinant, placing X,=1, X,=X, X,=Y, and writing F for F(1, X, Y) and H for H(1, X, Y), one finds H = 0 as the affine equation of the Hessian, where nn —1)F (nw — 1)F/OX (n — 1)0F/aY H=|(n—0P/oX Pox? ePlaxoY (n —l)aF/oY eFjayex ePloy® POINTS OF INFLECTION 69 Now assume that the origin (0, 0) is a simple point but not an inflection. Let the tangent at the origin be ¥ = 0. Then I’ is given by F = 0 with F=Y+cX*+---,¢ #0. Then 0 0 n-1 HO,0)=| 0 2% + |= 2m —1)% 40, n—1 so that the Hessian does not pass through the origin. Remark, Tn assuming that the simple point is at the origin, we are using the fact that the Hessian curve is a projective notion: this should be checked. We are also using Theorem 7.2 (see the exercise on p. 47). ‘As the order of the Hessian (more precisely: the cycle H = 0) is 3(n — 2), we have: Theorem 10.4. A nonsingular curve of order n > 3 has at least one point of inflection. Let I’ be a nonsingular cubic. Take the coordinate system in such a way that E,:(0, 0, 1) is on I. With reference to X, = 0 as line at infinity, the equation of I’ takes the form a(X)¥? + 6(X)¥ + c(X) = 0, where a, b, ¢ are polynomials without common factor; the degree in Y is 2, since some vertical lines cut I’ in two points and none cuts in more than 2 (at finite distance). Taking (0, 0, 1) to be a point of inflection with X, = 0 as tangent, every vertical line cuts I’ at finite distance in two points (counting with multiplicity), and hence a(X) = 0 can have no root, i.e., a(X) = const. Since i(I', 1.3 H,) = 3, one finds that deg (X) <1. Making the linear transformation X' = X, Y’ = Y + 6(X)/2a, one eliminates the Y’-term: the equation of I’ then has the form Y? = g(X). Here g(X) is a cubic, since isa cubic. Thus I’ is given by Y* — d(X — a)(X — 6X —c) =0. Replacing Y by Y’Vd, we may suppose that d = 1. The roots a, b, ¢ of g are distinct, since if a = 6, for example, then I’ has a multiple point at (a, 0). Moreover, by taking a new coordinate system, keeping (0, 0, 1) and (0, 1, 0) as vertices, but taking (a, 0) as a vertex and the unit point collinear with B, and (6, 0), the equation takes the form ¥? = X(X — 1)(X — ec). Hence: Theorem 10.5. With an appropriate choice of the coordinate system, every nonsingular cubic can be put in the form Y? = X(X —1)(X —c), c #0. (See Fig. 10.1.) Let P, be the point (a, 0), for arbitrary a. Then c gives the cross-ratio of Py, Py, P., Pin some order. Since EPs, H,P;, EP, 2,P.,, are the four tangents from E, to I’, c can be described directly in terms of I’, namely, as the cross-ratio of these tangents at H,. In the above construction (of the 70 CUBICS (CONTINUED) E» Pf) E\=Po i V Y?=X(X-1)(X-0) Fig. 10.1 equation Y? = X(X — 1)(X —c)), the roles of P,, P,, P, can be arbitrarily interchanged. For the six possibilities, one gets the cross-ratios c, 1/c, 1 — c, 1/(1 —c), —c/(1 —c), (1 — c)/—c, which are in general distinct. Thus in general one gets six different standard equations for I’. Let S be the system of values c, l/c,...,(1 —¢)/—c. These are defined in terms of a selected point of inflection, but we will show that S depends only on T' and not on the point of inflection. Theorem 10.6. The Hessian of T cuts T' at an ordinary point of inflection with multiplicity 1. The theorem is true for I’ of arbitrary order n > 3, but we will prove it for the present only for n = 3. Proof. Taking a flex at (0, 0) with Y = 0 as tangent (and applying 10.5 with change of notation), the equation of I’ can be written: Y —X(X — Y\X —cY) =0. In the Hessian determinant H, dropping terms divisible by Y or by X? we get io ie, H(X, Y) contains the term —24X. Hence H = 0 has a simple point at the origin and Y = 0 is not the tangent. Hence the intersection multi- plicity is 1 as stated. Corollary 10.7. T' has only a finite number of points of inflection; H#O(F). Corollary 10.8. A nonsingular cubic has exactly nine points of inflection. POINTS OF INFLECTION 7 Theorem 10.9. Let Tbe a nonsingular cubic. Then any two points of inflection are collinear with a third. For any two points of inflection A, B, there is a linear transformation taking T' into itself and A into B. Hence the cross-ratios of the tangents at A are the same as those at B. Proof. Let A be placed at E, and let IT have the equation ¥? = X(X —1)(X —0). This curve is symmetric in the X-axis (ie., is sent into itself by the linear transformation X’ = X, ¥Y’ = —Y). The points P,, P,, P, are not points of inflection, since the tangents there go through Z,. Hence B is not on the X-axis. Hence the symmetric point, B’, is also a point of inflection; and BB" passes through E,. This proves the first point. Now let A, B be two given points of inflection and let C be the third point of inflection collinear with A and B. Take C at Ey, Then A, B become symmetric points of inflection on I’, and the transformation X’ = X, Y’ = —Y interchanges A and B and sends I’ into itself. Corollary 10.10. The system S = Sy depends only on T', and not on the point of inflection by means of which it is defined. Theorem 10.11. Two nonsingular cubics [', 1’ are projectively equivalent if and only if Sp = Sp. Proof. If a linear transformation takes I’ into I’, it takes the points of inflection of I’ into those of I’, and the tangents to I’ from a point of in- flection P into the tangents to I’ from the corresponding point P’; hence Sp = S,. Conversely, let Sp = Sp. Ifc is an element of S,, then I can be given the equation Y? = X(X — 1)(X —c); and for the same reason, T’ can be given the same equation. Hence I’ and I” are projectively equivalent. Ifc is the cross-ratio of four lines taken in some order, then the cross-ratios for various orders are c, l/c, 1 — c, 1/(1 — ¢), 1 — Ife, 1 — 1/(1 — ¢), all of which can be obtained from c by a succession of substitutions of the form o> lJeandc +1 —c. Since A =4(L — 0 + c%)%/(c + 1)%(2e — 1)%e — 2) is invariant under these substitutions, the six cross-ratios yield the same value of A. Conversely, given A, the above equality yields a sixth-degree equation, which can have at most six roots. Except for the cases 1 — c + c? = 0 and (c + 1)(2c — 1)(c — 2) = 0, which may be handled separately, the values ¢, Vc, ete., are distinct, hence any other cross-ratio c yields a different value of A. In the case of a nonsingular cubic I’, then, Sp determines and is 72° cUBICS (CONTINUED) uniquely determined by A: this number is called the modulus of A. Theorem 10.11 can also be phrased: 10.11.1. Two nonsingular cubics T, 1’, are projectively equivalent if and only if they have the same modulus. Exercise. Let I’ be a nonsingular cubic and define a group on it with 0 at a point of inflection. Show that the nine points of inflection are 0, X, —X, Y, —Y, -X—Y,X+¥, X—Y, —X + Y, the nine solutions of 3Z = 0. Call these points A,, 49, 4g, By, Cy, Cz, By, Cy, By (in the given order). Take a coordinate system in such way that 4,:(1, 1), 42:(1, —1), By:(—1, 1) By:(—1, —1). One then finds that Cz: (0,0), 4s: (1, ®), By:(—1, —a), C1: (B, Ds Cz:(—B, —1) with « = V—3, 8 = —V—8. Thus the nine points are deter- mined, up to a projective transformation, independently of I. Standard forms for cubics with a singularity are found by methods similar to the one used for a nonsingular cubic. If T has a cusp, place it at E, with X) = 0as tangent. Then I’ takes the form Y = g(X), where g(X) isa cubic. Taking some point on I’ as H, and the tangent there as the X-axis, we see that I’ takes the form Y = aX*(X —c). Here c = 0 is possible, but taking Z, to be not a point of inflection, we have c # 0. By appropriate choice of the unit point, we may suppose that c= 1. Replacing Y by Y’ = Y/a, we have a = 1. Therefore all cubics with a cusp are projectively equivalent. Hence: Theorem 10.12. Every cubic with a cusp is projectively equivalent to Yt = X8, Exercise. If a cubic has a cusp, then it has a point of inflection. Now let T have an ordinary double point. Place it at , with Xy = 0, X, = 0s the tangents. Then I’ takes the form X,X,X, -+ a cubic form in X,, X; (as one sees by referring I’ to X, = 0 as line at infinity and then restoring homogeneous coordinates). Thus TP: XpX:X_ + aX3 + OX2X, + cX)X? + dX? = 0. Replacing X, -+ 6X, + cX, by X, (and then d by 8), we get 1: XpX,X_ + aX3 + bX? = 0. Here a # 0 and b #0, as I’ is irreducible. Applying the transformation Xi, = Va, Xo, Xi = Vb X,, Xi = X,, we may achieve a=1, b=1. Hence: Theorem 10.13. All irreducible cubics with an ordinary double point are projectively equivalent. CUBICS THROUGH EIGHT POINTS 73 Exercise. Obtain another proof of this therom as follows: Show that the Hessian of a cubic I’ with an ordinary double point has intersection multiplicity 6 with T there. Hence T’ has a point of inflection. Place a point of inflection at By, ete. CUBICS THROUGH EIGHT POINTS Every conic has an equation of the form yoXG + to XoXy + MoeXoXs + MX] + ayeXiXe + ayeXZ = 0. There are six coefficients here. If we ask that this conic pass through a point (2, 21, %,), we impose the condition Aggt +++ = 0. Regarding the a;, as unknowns, we find that this is a linear condition on the ay. From the theory of linear equations, it is known that any five such conditions can be satisfied nontrivially. Hence: Theorem 10.14. Through any five points in the plane one can pass at least one conic. As the condition aggx? +--+ = 0 and the condition Aaoots, + are the same, we associate the condition with the point (ag, xory,... , 243) in projective 5-space. Then we use the terminology of this 5-space also for the conditions; e.g., we say that one condition is (linearly) dependent on others, that several conditions are (linearly) independent, ete. Let Pi, P2,...,P; be five points and let L,,L,,...,L; be the corresponding conditions on a conic to pass through them. Then there will be only one conic through the points if and only if Ly, ..., Ls are independent: this is a direct translation from a familiar theorem on homogeneous linear equations (namely, that r linearly independent equations in n unknowns has n —r, but not more, linearly independent solutions). Now if Ly,..., Ls are independent, then none is a consequence of the others (i.e., by definition, any four have a solution which is not a solution of the fifth—this will be so because any four have 6 — 4 = 2 linearly independent solutions), and there will be a conic through any four of P,,...,P; not passing through the fifth. Conversely, if through every four there is a conic not through the fifth, then the conditions are independent. If four of the points are on a line m, then any conic through three of them contains the line m and hence the fourth point; the four (also the five) points impose dependent conditions on a conic to contain them, and there will be infinitely many conics through the five points. On the other hand, if no four of P,,..., Ps are collinear, then 0, AFO(EO), 74 CUBICS (CONTINUED) one can pass a conic (even a reducible one) through any four of them not through the fifth; the five points impose independent conditions; hence one can pass only one conic through the five points. (We say that a linear condition L is a consequence of linear conditions Ly,..., L, if every solution of L,,..., L, is a solution of L. By a theorem from the theory of linear equations, this will be so if and only if L is a linear combination of L;,...,L,.) The same ideas apply to cubics. The general cubic equation has ten coefficients. Hence: Theorem 10.15. A cubic can be made to pass through any nine points. Lemma 10.16. Let P,,...,P, be eight distinct points and let Ly,..., Lg be the conditions imposed on a cubic to pass through them. If no four of the P; are collinear and no seven co-conical, then Ly,..., Ly are independent. Proof. We show that a cubic can be put through any seven of the P,, say P,,..., P;, without passing through the eighth. On P,P, there is at most one other P;, hence there is a P,, say P,, such that P,P, does not pass through P,. A similar argument shows that there is a third line P,P,, say P,P,, not through P,. Now pass a conic C, through P,,...,P; and P,, and a conic C, through P,,...,P, and P,. Then not both C, and C, can pass through P,. If they did, then C, 4 C,; otherwise the points P,,..., Py would be co-conical. Then the distinét conics C',, C, pass through P,,..., Ps, whence some four of P,,..., Ps are collinear, which is excluded by hypo- thesis. Thus C, or C, does not pass through Ps, say C, does not. Then C, and the line P,P, form a cubic through P,,..., P; but not through P,. Theorem 10.17. Let two cubics meet in just nine (distinct) points, Py,...,Py. Then any cubic through any eight of the P, also passes through the ninth. Remark, Let Ly,...,Ly be the conditions imposed on a cubic to contain P,,...,P». Since the conditions L,,...,Z,) have two solutions, Ly,...,D, are dependent, and one of them must depend on the others. Hence all the cubics through some eight of the P; also pass through the ninth. This doesn’t quite prove the theorem. Proof of the theorem. No four of the P, are collinear and no seven co-conical, since otherwise the two cubics would have a common component. Hence any eight of the conditions L, are independent, so the ninth must depend on them. A slightly deeper statement is the following. THE MODULUS OF A NONSINGULAR CUBIC 75 Theorem 10.18. All cubics through any eight given (distinct) points, P,,..., Ps, also pass through a ninth. Here the ninth point may be “infinitely near” one of the P,, say Py, whereby we mean that there is a line m through P, such that i(T', m; P,) > 1 for every cubic T through Pee Py Proof. If four of the P; lie on a line g, then any cubic through the P, contains g as a component, and any point of g (#P;, i = 1, ..., 8) can be taken as the ninth point. If no four P, are collinear but some seven lie on a conic C, then C is irreducible, since otherwise some four of the P; would lie on a component of (. Then any cubic through the P; contains C as a component, and any point of C (#P,, i = 1,..., 8) can be taken as the ninth point. We may assume, then, that no four of the P, are collinear and that no seven are co-conical. Now take two cubics I',, I, through P,,..., Pg and suppose they meet in a ninth point P,. Since, by Lemma 10.16, the conditions Iy,...,Lg are independent, but conditions Ly,..., I, are dependent, Ly depends on Iy,...,L,, and all cubics through P,,...,P, also pass through P,. However, two cubics through P,,..., P, may meet just in those points, therefore with multiplicity > 1 at one of the points, say P,; in fact, since the cubics meet just in the points Py, ..., Ps, they have no common component, and the intersection multiplicity will be 1 at seven of the points and 2 at the eighth (P,). Thus at least one of ',, P, has a simple point at P,, say I’, does, with m as tangent; then I’, must have at P, either a double point or a simple point with m as tangent. Place P, at (0, 0) with mas X-axis and write the general cubic in the form Boo + BigX + byY + + = 0. The condition that a cubic pass through (0, 0) is that by) = 0, which is a linear condition; the condition that it pass through (0, 0), be simple there, and have m as tangent (or be a double point) is byg = 0 and byy = 0, ice., two linear conditions. The condition by) = 0 is L,; designate the condition by = 0 by Ly. Then Ly,...,L) are dependent, since they have two solutions, and as L,,..., D, are independent, Ly is a consequence of them. Hence any cubic through Py, ..., P, either has a simple point at P, with m as tangent or has a double point there. This completes the proof. THE MODULUS OF A NONSINGULAR CUBIC Above (see 10.5 through 10.11) we classified the nonsingular cubics from a projective point of view. Later (see Chapter 20, Example 6) we shall want to do this from a so-called birational point of view and shall need a result of Salmon on the nonsingular cubics. 76 CUBICS (CONTINUED) Let T': F = 0 bea cycle and Pa point. If F(X», X;, Xq) is irreducible (or even if it merely has no multiple factors) and if P is not on I’ (so that the polar is certainly defined), then the polar A meets F = 0 in n(n — 1) points Q, where n = degree of F. From P, then, one can draw n(n — 1) tangents PQ to F = 0. These may not be distinct, but if we count PQ i times, where i = i(T’, A; Q), we get the full number n(n — 1) of tangents. If ': F = 0 is an irreducible nonsingular cubic, then from P one can draw six tangents. For orientation, consider the remark: if P approaches a point A of the curve, two of the six tangents coalesce with the tangent at A; thus one can draw four tangents from A to I’, not counting the tangent at A. We now make this remark precise. One checks the following statements: a) The polar A of A passes through A. The intersection multiplicity iI’, A; A) is 2 if and only if A is not a point of inflection, in which case it is 3. b) IfQ # A is on T and on A, then i(T, A; Q) = 1. Exercise. Establish the following statements: a’) [and A both have simple points at A and have a common tangent there and nowhere else. b’) A splits into two lines, one of which will be the tangent at A, if and only if A is a point of inflection of D. Establish (a) and (b). Thus if A is not a point of inflection, one can draw precisely four distinct tangents from A to T’, not counting the tangent at A. If A isa point of inflection, then one can draw precisely four, this time counting the tangent at A. Note that no two of the intersections Q,, Q. of I and A can be collinear with A, as otherwise Q,Q, would intersect I’ more than three times. Appropriately interpreted, we have: From every point A of T’ one can draw precisely four tangents to T. Let P, Q be points on a nonsingular cubic I" and consider the construc- tion illustrated in Fig. 10.2. Through P draw a line, cutting I’ again in A and A’; let QA cut I’ in the third point B, and Q4’ in the third point B’. Then we say: The lines BB' pass through a fixed point R on T. Proof. Setting up a group on I’, let 0 be the zero and K the third intersection with I’ of the tangent at 0; by taking 0 to be a point of inflection, we may also arrange to have K = 0. Now by 9.7 we have P+A+A'=K, Q4+44B=K, Q+A 4B =K, R+B+B=K, THE MODULUS OF A NONSINGULAR CUBIC 77 Fig. 10.2 from which one deduces P — Q = Q — R, whence R = 2Q — P is fixed. Moreover, P and R can be taken arbitrarily on T. That is, we are to show that R + P = 2Q can be solved for Q. Write this in the form: D+ 2Q=0. This is equivalent to (D + K) + 2Q= K. This comes to saying that D + K, Q, Q are collinear, i.e., that Q is the point of contact with a tangent to T from D + K; we know that such tangents exist. Q.E.D. The above construction obviously sets up a one-to-one correspondence between the lines on P and those on R; moreover A = A’ if and only if B = B’,so that the tangents from P to I’ correspond to those from R to I’. We will prove that this correspondence is a projectivity, and then we shall have: Theorem 10.19. There is a projectivity between the pencils at P and R such that the four tangents from P to 1’ correspond to the four tangents from R to T. Hence the cross-ratio of the four tangents from P (taken in some order) is the same as the cross-ratio of the four tangents from R (taken in some order). We first give a rough indication of why the above correspondence is a projectivity. Take a coordinate ¢ for the lines of the pencil of center P, ice., let fy = 0, l, = 0 be two lines through P and write the lines through P in the form J + tl, = 0. Let ¢’ be similarly defined for the pencil at R. Now we can find the intersections of J, + tl; = 0 with I': F(X, Y) = 0; eliminating Y, say, from F(X, Y) = 0 and 1,(X, Y) + #,(X, Y) = 0, we would get a cubic in X, yielding the abscissas of P, A, A’ (see Fig. 10.2). Since P is known, we have one of the roots of the cubic, and so have only to solve a quadratic to get the other roots; this will introduce a square root of an expression in ¢ into the computations. Now we can compute the coordinates of B and of B’ (we only have to solve linear equations for this), 78 cuBIcS (CONTINUED) and then can write the equation for BB’. In this way we can write ¢’ in terms of t; a square root may enter the expression for ¢’ in terms of t, but if 80, we eliminate it, and get a polynomial relation G(t, t’) = 0. The poly- nomial G(T, 7’) may have multiple factors, but we keep each irreducible factor only to the first power: in this way we get an equivalent condition on (t, t’), and may suppose that (7, 7’) is free of multiple factors. Consider the curve G(T, T’) = 0. For each value of t, we get just one value of t’. The values of ¢’ are obtained from G(t, T’) = 0. Let G(T’, T’) be of degree nin T’. Then, with the exception of a finite number of values of t, G(t, T'') is still of degree n in T’. Now G(t, T’) = 0 may have multiple roots: this will happen if and only if the line 7’ — ¢ = 0 is tangent to G(T, 7’) = 0 or passes through one of its singularities (at finite distance), Thus avoiding a finite number of further values of ¢, we get distinct values for t’. Thus n= 1. Similarly one argues that G(T, 7’) is of degree 1 in T. Thus G(T, T’) = 0 is of the form a + bT + cT’ + dTT’ = 0. Then pe ah e+ dt” Here ad — bc # 0, otherwise ¢’ would be constant. Hence the correspondence is a projectivity. The above argument, though quite suggestive, does contain a slippery point or two, since the nonhomogeneous treatment does not take into account all the elements of the situation. We give a precise proof. The proof will show quite generally that if a one-to-one correspondence between the lines of two pencils can be described in an algebro-geometric manner, then the correspon dence is a projectivity, though we stick, at subsidiary points, to our present situation with a cubic. We use homogeneous coordinates. Let P, Q, R, A, A’, B, B’ have coordinates (Po, Pr» Pa)s (Gor G1» Ya), etc. Let lo(Xo, X, Xa), ly(Xo, Xi, Xz) be two linear forms yielding two (distinct) lines through P, so that every line through P has an equation of the form fl) + tl, = 0. Similarly let U, = 0, U =0 be lines through R, so that every line through R has an equation of the form (iJ, + til, = 0. Let (to, t,) be the coordinates of the line PAA’ (ie., PAA’ is given by tal -+ tl, = 0) and let (t%, t{) be the coordinates of RBB’. Let I be given by F(Xo, X;, X;) = 0. Now we write equations mirroring the geometric facts: (1) Fd, dy, %) = 0, F(a, a}, 3) = @) tolol@os Ay, 42) + tyly(Ao, a1, 42) = 0, tolo(™, @, 4%) + th(ap, a, a) = 0. These equations say that A and A’ are on the intersection of I and THE MODULUS OF A NONSINGULAR CUBIC 79 tolo + tl, = 0, but they do not say that (P, A, A’) is a complete set of inter- sections of I’ and the line; for example, Eqs. (1) and (2) are both satisfied by the triple (P, A, A), or even by (P, P, P). For a reason that will be apparent in a moment, we do not wish to write inequalities. We proceed as follows. Let a coordinate system be taken in such way that Z,: (0, 0, 1) and E, : (0, 1, 0) are not on I’ and such that P,Q, R are not on H,Z,. Then Ry,(F, tolo + th) = 0 yields the ratios py: p1, dia, a): a; of the points P, A, A’ as solutions (see 7.5). Therefore we wish to say that this resultant is const (pyX, — pyXq)(4X1 — a4Xy)(ajX, — ajXq). Let uy, %4,... be the various coefficients of Ry, = P(X», X,) and let v», v,,... be the correspond- ing coefficients of Q(Xo, X1) = (PoX1 — PrXo)(4oX1 — 4Xo)(4HX1 — 4X). We do not say that (w%, %4,...) and (v%, 24,...) are equal, but only that they are proportional. This is expressed by: Uy My Ug Ue (3) =0, =0, ete. Mm 1 Um Ye The w, are homogeneous and of the same degree (= deg F) in (to, t,), and the v; are homogeneous of a degree 1 in a and in a’. Consider also Ry(F, tolo + th) = P'(Xo, X2) and Q (Xo, X2) = (PoX2 — P2Xo)(aoXz — AX o)(apXz — a5X,). Let (uj, wi,...) and (vj, v{,...) be the coefficients and consider the conditions : Uy Uy % % Ug My (4) =0, ete. % % Of course, we also have (8) F (Po, Py P2) = 0. Throughout we are speaking of points (ag, 41, 42), (Po, Py» Pa), ete., i.e., not all the entries are zero; and the same goes for (tp, f) and (ts, 4). The conditions (1) through (5), all equalities, say that (P, 4, 4’) is a complete set of intersections of I’ and fly + tl, = 0. Conditions (1), (2), (3), and (5) do not quite do it, as one sees by considering the case that to + th = 0 passes through By. Write, now, (6) F(b)=0, — F(b!) = 0, (7) tolo(b) + e{;(b) = 0, HlG(b") + ty; (b") = 0, and, introducing coordinates 4”, t” for the pencil at Q, write conditions (8) 80 CUBICS (CONTINUED) and (9), like (3) and (4), saying that (Q, A, B) is a complete set of intersections of line BA with P; similarly write conditions (10), (11) for line A’B’ and conditions (12), (13) for line BB’. Statements (1) through (4) and (6) through (13) are a precise counterpart of the construction leading to the correspondence PA —> RB. [Statement (5) and corresponding statements for Q, R need not be made.] Now by the result quoted in (4.4), one can eliminate any of the unknowns a, b, ete. In this way we can successively eliminate a, a’, b, b’, t”. This yields a condition (14) Glo 5 ty H) = 0. This condition is equivalent to conditions (1) through (13): any solution t,a,..., t” of (1) through (13) yields a solution (t, ¢’) to (14); and for any solution (t, t’) of (14), there exist a,...,b’ satisfying (1) through (13). Thus the correspondence PA — RB is given by (14). The polynomial G(T, T’) is homogeneous in T = (Z', T;) and in T’ = (1%, T!). By 3.6, the irreducible factors of G(T, 'T’) will also be homo- geneous in T and in’. Keeping each irreducible factor only to the first power, we may suppose that @ is without multiple factors. Also, @ is not divisible by 7; otherwise for (tg, t) = (0, 1), (14) would have infinitely many solutions (t, #1). Hence degy G1, T; To, Ti) = degen, 7y4(To, Tr; To, 71). Similarly, degy GL, Ts 1, 7") = degen, ny A(T, Tr; Ty, T%)- By considerations similar to those in the last part of Chapter 3, one also sees that G(1, 7; 1, 7’) is without multiple factors. Moreover, any solution of (15) Gi, t; 1,¢) =0 yields a solution of (14). Hence by an argument already made in the pre- liminary remarks, G(1, 7; 1, 7’) is of degree 1 in 7 and of degree 1 in T’. A similar statement can be made for G(T, T’) and G(T, T’) is of the form ATT, +b TT, +c TT +a TyT'. Then (15) is a+bi+ ct + du’ =0, and, as before, ad — be # 0. Hence (14) yields a projectivity. Q.E.D. Remark. The difficulty mentioned concerning the lack of generality in proceeding nonhomogeneously can also be overcome by a device employed in Chapter 9, as follows. One first writes the equation F = 0 for T homo- geneously, i.e, F = F(X, X;, X,), and one takes homogeneous coordinates (to, ty), (to, #4) for the pencils at P and at R. Then one subjects I’ to a general linear transformation. In the transformed situation one considers only THE MODULUS OF A NONSINGULAR CUBIC 81 transformed elements, i.e., points and lines in the original plane PNC. The transform of I’ and any transformed line will be in permissible position. Let PY be the transform of P; then (ty, ¢,) are also coordinates in the pencil at P’. We wish to pass to the nonhomogeneous coordinate t/t, but have to face the exceptional case f, = 0. To meet this one also subjects f, t, (and similarly (4, t;) to a general linear transformation: (3) i le Pa a M0 nu) \h)’ and now (* # 0 for any transformed line. Now one dehomogenizes and uses the argument given at the beginning. The result is that corresponding lines (to ty), (th, #4) satisfy a relation G(T, T,; 7), 7%) = 0 of the desired form, except that @ may have coefficients in an extension field C, of C. To meet this new point, one writes @ in the form @ = oly + +++ + c,G,, where c;€C, and the @; have coefficients in C. If a c, can be written linearly in terms of the others with coefficients in C, one can eliminate it; in this way, we may suppose that the c, are linearly independent over C. But then corre- sponding lines (fo, t,), (tj, ¢1) satisfy all of the G,. Chapter 11 EXPANSION AT A SIMPLE POINT We start with some motivating remarks. Let P': F(X, Y) = 0 be an irreducible curve having a simple point at the origin with Y = 0 as the tangent. Then F(X, Y) = c¥ — G(X, Y), c€C, and subd @ > 2. Without loss of generality, we assume that c = 1. Then any point on I’ satisfies qd) Y = G(X, Y). If we substitute this expression for Y into the right-hand side, we get (2) ¥ = GX, Q(X, Y)), which is also, clearly, satisfied by any point on I’. Now in (1), Y appears on the right only in terms of degree > 2, and hence in (2), Y appears on the right only in terms of degree > 3. And we can repeat the process, getting relations Y = H,, which are satisfied by all points of I’ and in which Y appears on the right only in terms of degree > n + 1. For example, starting with Y=X%+4XY, we first get Y= X%+4X(X*4XY), ie, Y = X84 X44 XY; after another step we get Y = X84 X84 XAX84 X44 XY), ie, Y= X%+4 X4+ X54 X® + X*Y, ete. Since Y — G(X, G(X, Y)) vanishes over I’, we have Y — G(X, G(X, Y)) =0(¥ — @(x, ¥)), ie., Y — G(X, O(X, ¥)) = (¥ — @(X, Y))+ H(X, Y). Here H(X, Y) = 1+ +--+, since Y occurs on the left, and hence must occur on the right. The curve (2), then, consists of I plus, perhaps, a component: not passing through the origin. Thus (2) has exactly the same appearance as (1) near the origin. And the same can be said for the various curves Y = H, obtained by repeating the process. 82 FORMAL POWER SERIES 83 Now roughly speaking, if |z| and |y| are small, we can neglect higher- degree terms relative to lower ones. For example, we could say of the curve Y =X%4 X44 X54 X64 XY by neglecting the terms of degree > 5, that it is approximated by the curve Y = X%+ X* near the origin. If we make the substitution mentioned, we get Y= X94 X¢4 X54 X84 X4. (X34 X84 X54 X84 KY), and neglecting the terms of degree > 9, we could say that the original curve (Y = X84 XY) is approximated by Y = X* + X#+---+ X*% But without neglecting anything, if we repeated the process infinitely many times and could use some kind of limiting process, we would say that T: ¥ = X* + XY is given near the origin by Y = X*+ X44--- Surely, working with complex numbers, we would conjecture that XS 4 X4+ is convergent near X = 0 and that Y = X*+ X4+--- gives the points of Y = X* + XY near the origin, and the conjecture would be correct. If we could replace Y = X*+ X*?¥ by Y= X#+4 X4+--- in investigating phenomena near the origin, it would be quite helpful. For example, in studying intersection multiplicity, the resultant of G(X, Y) and Y — X3— X# —..-is easy to handle—it is simply G(X, X*+ X#+--+) —the unwieldy determinants would be eliminated. The above suggests that the proper tool for a study near the origin is the power series—we should work with power series rather than polynomials. It turns out that we need not be concerned with questions of convergence, but can work formally with the power series. How this is done, we now explain. FORMAL POWER SERIES Let K be a field. By K[[X, Y]] we designate the set of objects of the form F = gy + QyyX + ay ¥ + GogX? + ay XY + ay¥2 +++, aye K, among which are defined the operations of sum and product according to the rules: (Aq + a1oX + Gor ¥ + +++) + (boo + broX + bor ¥ + +++) = (oo + boo) + (410 + O10) X + (dor + bo) ¥ + °°, (4o9 + @yoX + aq ¥ + +++) * (boo + b10X + bY + +**) = Agqboo + (oobro + Ar0bo0)X + °° With these two operations K[[X, Y]] becomes a commutative ring whose elements are called formal power series. It is customary to write the power series in the form PoRth+h +o, 84 EXPANSION AT A SIMPLE POINT where F; is a homogeneous polynomial in X, Y of degree i; (also, if an F,=0, it may be left out of the above expression). Starting from this convention, we define (Fy + F,+-+:)+(@+G+°"") as (Fo +) + (Fr + Gy) +++ and (Fy + Fy +o) Got G+ °°) as FyG, + (FoG, + F,G)) + -*++ If one has FHF,+ Fate, F, #0, then one calls r the subdegree or order of F. It is immediate that ord (F + @) > min {ord F, ord G} and ord FG = ord F + ord G. Taking into account this last property, one sees easily that K[[X, Y]]is an integral domain. One designates its quotient field by K((X, Y)). If one has a sequence of power series F; such that ord F;— 0 as i> co, one may define dF. ¢ K(X, Y]] as follows: Let Fo=Fe+ Fates with F,, homogeneous of degree j. Then we place @ Fe = (LF) + (LPFa) ++. if which makes sense since each sum ,(F,,) is finite by the hypothesis that ord F; > wasi— o. Remark, If F = Fy + Fy-+-++, then F = ¥2qF, in the sense just defined. One also has the properties LF + LG = LF + @), PYG, = YFG,. Theorem 11.1. The units of K[[X, ¥}]are the power series of order zero. Proof. If FG = 1, then ord F + ord @ = ord 1 = 0, whence ord F = 0. Conversely, let F = aq + dX + ay ¥ + +++ = dy(l — @), where dy #0 and ord @ > 0. Now 1 — G@ has $254" as inverse, as one easily proves using the properties mentioned above. Hence F has ago! @ as inverse.

You might also like