Автор: Hodge W.V.D.   Pedoe D.  

Теги: maths  

Год: 1953

Текст
                    METHODS
OF
ALGEBRAIC GEOMETRY
by
W. V. D. HODGE, Sc.D., F.R.S.
Lowndean Professor of Astronomy and Geometry, and
Fellow of Pembroke College, Cambridge
and
D. PEDOE, B.Sc. (Lond.), Ph.D. (Cantab.)
Sometime Charles Kingsley Bye-Fellow
of Magdalene College, Cambridge
VOLUME I
BOOK I: ALGEBRAIC PRELIMINARIES
BOOKII: PROJECTIVE SPACE
CAMBRIDGE
AT THE UNIVERSITY PRESS
1953
*


* PUBLISHED BY THE SYNDICS OF THE CAMBRIDGE UNIVERSITY PRESS London Office: Bentley House, n.w. i American Branch: New York Agents for Canada, India, and Pakistan: Macmillan Printed in Great Britain at the University Press, Cambridge (Brooke Crutchley, University Printer)
* CONTENTS PREFACE page vu BOOK I ALGEBRAIC PRELIMINARIES Chapter I: RINGS AND FIELDS 1. Groups 2. Rings 3. Classification of rings 4. The quotient field of an integral domain 5. Polynomial rings PAGE PAGE 2 6. The division algorithm 24 6 7. Factorisation in an integral 12 domain 27 8. Factorisation in polynomial 15 rings 33 19 9. Examples of fields 37 Chapter II: LINEAR ALGEBRA, MATRICES, DETERMINANTS 1. Linear dependence 41 2. Matrices 49 3. Square matrices 54 4. Transformations of a matrix 58 5. The rank of a matrix 65 6. Homogeneous linear equations 68 7. Matrices over a commutative field 71 8. Determinants 73 9. A-matrices 86 10. Miscellaneous theorems 95 Chapter III: ALGEBRAIC DEPENDENCE 1. Simple algebraic extensions 2. Extensions of a commutative field 3. Extensions of finite degree 4. Factorisation of polynomials 5. Differentiation of polynomials 99 6. Primitive elements of algebraic extensions 126 . Differentiation of algebraic functions 128 . Some useful theorems 136 101 105 111 119 Chapter IV: ALGEBRAIC EQUATIONS 1. Introduction 139 2. Hilbert's basis theorem 142 3. The resultant of two binary forms 146 4. Some properties of the resultant 152 5. The resultant of a system of binary equations 156 6. The resultant system for a set of homogeneous equations in several unknowns 159 7. Non-homogeneous equations in several unknowns 162 8. Hilbert's zero-theorem 165 9. The resultant ideal 166 10. The w-resultant of a system of equations 171
vi CONTENTS BOOK II PROJECTIVE SPACE Chapter V: PROJECTIVE SPACE: ALGEBRAIC DEFINITION 1. Introduction 2. Projective number space 3. Projective space of n dimensions 4. Linear dependence in PJ,(i£) 5. Equations of linear spaces. Duality PAGE 175 176 178 180 6. Desargues' Theorem 7. Some fundamental constructions 8. The condition for a commutative field; Pappus' Theorem 9. Some finite geometries 186 10. r-way spaces PAGE 191 196 202 205 206 Chapter VI: PROJECTIVE SPACE: SYNTHETIC DEFINITION 253 1. Propositions of incidence 208 2. Desargues' Theorem 213 3. Related ranges 216 4. Harmonic conjugates 224 5. Two projectively invariant constructions 231 6. Reference systems 246 7. The algebra of points on a line 8. The representation of the incidence space as a P^K) 9. Restrictions on the geometry 10. Consequences of assuming Pappus' Theorem Chapter VII: GRASSMANN COORDINATES 5 1. Linear spaces 286 2. Grassmann coordinates 288 3. Dual Grassmann coordinates 292 4. Elementary properties of Grassmann coordinates 297 Some results on intersections and joins Quadratic p -relations The basis theorem 259 272 274 304 309 315 Chapter VIII: COLLINEATIONS Projective transformations 322 Collineations 327 United points and primes of a non-singular collineation 333 4. The united fc-spaces of a non- singular collineation 5. Cyclic collineations 6. Some particular collineations 7. Singular collineations Chapter IX: CORRELATIONS 1. Correlations 362 2. Polarities 367 3. Null-polarities 378 4. Simple correlations 390 5. Transformation of the general correlation 395 Bibliographical Notes Bibliography 6. Canonical forms for correlations 7. Some geometrical properties of correlations 344 352 354 358 411 419 430 431 Index 433
* PREFACE This volume is the first part of a work designed to provide a convenient account of the foundations and methods of modern algebraic geometry. Since nearly every topic of algebraic geometry has some claim for inclusion it has been necessary, in order to keep the size of this volume within reasonable limits, to confine ourselves strictly to general methods, and to stop short of any detailed development of geometrical properties. We have thought it desirable to begin with a section devoted to pure algebra, since the necessary algebraic topics are not easily accessible in English texts. After a preliminary chapter on the basic notions of algebra, we develop the theory of matrices. Some novelty has been given to this work by the fact that the ground field is not assumed to be commutative. The more general results obtained are used in Chapters V and VI to analyse the concepts on which projective geometry is based. Chapters III and IV, which will be required in a later volume, are devoted to a study of algebraic equations. Book II is concerned with the definition and basic properties of projective space of n dimensions. Both the algebraic and the synthetic definitions are discussed, and the theory of matrices over a non-commutative field is used to show that a space based on the propositions of incidence can be represented by coordinates, without the introduction of any assumption equivalent to Pappus' theorem. The necessity of considering a large number of special cases has made Chapter VI rather long, but some space has been saved in the later parts of the chapter by merely mentioning the special cases and leaving the proofs to the reader, when they are sufficiently simple. It is hoped that this will not cause any difficulty. This Book concludes with a purely algebraic account of collineations and correlations. Certain elementary geometrical consequences are indicated, but a complete study of the geometrical problems involved would have taken us beyond our present objective. It is hoped that Volume II will appear shortly. This will be devoted to the theory of algebraic varieties, and to the study of certain loci which arise in many geometrical problems.
Viii PREFACE We wish to express our thanks to ProfessorT. A. A. Broadbent of the Royal Naval College, Greenwich, who has read this volume in manuscript and in proof, and to Mr D. B. Scott of the University of Aberdeen, who has read it in proof. We must also thank the staff of the University Press for the care they have taken in the production of this book, and for their ready courtesy in meeting our wishes. Note. The reference [II, §4, Th. II] is to Theorem II in §4 of Chapter II. If the reference is to the same chapter or section, the corresponding numeral or numerals will be omitted. w.v. D.H* D.P. CAMBRIDGE November 1946 *
* BOOK I ALGEBRAIC PRELIMINARIES CHAPTER I RINGS AND FIELDS The reader is assumed to be familiar with the use of homogeneous and non-homogeneous coordinates in geometry, when the coordinates are real or complex numbers. When geometry is developed with the help of these coordinates, results are obtained by methods which belong to algebra, the differential calculus, and so on. Those results which can be obtained by purely algebraic processes (and they include many which are usually obtained by the methods of the calculus) make up the subject with which we are concerned in this work. The operations of algebra we shall study are those of addition, subtraction, multiplication, division, and the solution of algebraic equations. While ordinary complex numbers are the most familiar elements for which these operations are defined, there are more general sets of elements for which it is possible to define them. By allowing our coordinates to belong to these sets, a more general geometry is obtained. We thus arrive at the definition of a more general space than that considered in elementary geometry, and the study of this space is the purpose of this work. In this and succeeding chapters we consider sets of elements for which some or all of the algebraic operations cited above are defined, and step by step we arrive at a characterisation of the sets of elements from which our coordinates may be chosen. These sets are known in algebra as fields. In order, however, that the geometry which we derive may conform to the general pattern which appears when the field is that of the complex numbers, we shall find it desirable to impose certain restrictions on the fields considered. These restrictions are not all imposed simultaneously, but in succession; and only when we have proceeded under the Umitations already adopted as far as our subject demands do we impose a new condition. The mathematical advantages of such a method are evident. HPI i
2 I. RINGS AND FIELDS 1. Groups. Consider a set S of elements, which we denote by a,6,c,.... A law by which, given any ordered pair of elements a and 6 of /S, possibly not distinct, we can derive a unique element c of S, is called a law of composition for S. A non-vacuous set S of elements with a law of composition which satisfies certain conditions, explained below, is called a group. We denote the element resulting from the combination of a and b (in the given order) by ab; if the resulting element of 8 is c, we write ab = c. ab is a uniquely defined element of S, but may be different from ba. The law of composition is said to be associative if, given any three elements a, 6, c of S, we have the equation (ab) c = a(bc). We then write this element as abc. The conditions that a set S, with a given law of composition, should form a group are that (i) the law of composition is associative; (ii) given any two elements a, 6 of S, there exist elements x, y such that ax = b and ya = 6. As an example of a group, consider the possible derangements of the integers 1, 2, 3. If a, /?, y is any derangement of these integers, we denote the operation of replacing 1, 2, 3 by a, /?, y by the symbol /1 2 3\ [a /? Y)' The successive application of any two of the six possible operations is equivalent to some single operation of the set, and hence a law of comppsition for the set of operations is defined. If we denote the six operations /1 2 3\ /12 3\ /1 2 3\ \1 2 3/' \2 3 1/' \3 1 2J' /1 2 3\ /12 3\ /1 2 3\ \1 3 2/' \3 2 l/' \2 1 3/ by a, 6, c, d, e, f respectively, and denote the result of performing
1. GROUPS 3 first the substitution x and then the substitution y by xy, the complete law of composition is given by the table of double entry: a b c d e f a a b c d e f b ~b~ c a f d e c c a b e / d d ~T e / a b c e e / d c a b f f d e b c a where the entry in the row containing x and the column containing y is xy. From the table, or directly from the definition, it is at once seen that the law of composition is associative. Again, since each row contains all six elements, the equation px = q always has a solution; the corresponding result for the equation yp = q follows from the fact that each column contains all six elements. Hence the six elements, with the law of composition, form a group. We notice in this group that bd = e, but db = /. Hence the equation xy = yx does not hold for all pairs of elements in the group. Such a group is called non-commutative. If, in a given group, the equation xy = yx is always true, we say the group is commutative, or Abelian. A very simple example of such a group is provided by the natural integers (positive, zero and negative), the law of composition being ordinary addition of integers. It is often convenient, when dealing with commutative groups, to use the symbol of addition for the law of composition, writing a + b instead of ab. We then call the group an additive group. It is important to remember that this notation is never used for a non-commutative group. We now obtain certain properties common to all groups. From condition (ii) we know that, given any element a, there exist elements e and/such that ae = a, fa = a. Let 6 be any element of S. Then there exist elements c, d such that ac = 6, da =* b.
4 I. RINGS AND FIELDS From condition (i) we have be = (da) e = d(ae) = da = 6, fb = f(ac) = (/a)c = ac = 6. In particular, in the first of these equations put 6=/, and in the second put 6 = e. Then / = /e, and fe = e. Hence e = /. If there is another element e' with the properties of e, e'e = e', and e'e = e, and therefore e is unique. We have thus established the existence of a unique element e of the group such that ae = a = ea for every element of the group. This element e is called the unity of the group. In the case of an additive group it is usually called the zero of the group, and denoted by 0. Now consider the equation ax = e, where a is any element of the group, e being the unity. By (ii) this has a solution x. Then xax =— xe -5 Xy and therefore the element / = xa has the property fx = x, for the x considered. But, by an argument used above, fb = b for any b in the group. In fact, let c be an element satisfying the equation xc = 6. Then fb =fxc = xc = b. It follows, taking 6 = 6, that / = e. Therefore xa = e. If y is any element such that ay = e = ya9 then 2/ = 2/^ = i/aa; = ea; = #. Hence a; is uniquely defined by the equations ax = e = xa. This element is called the inverse of a, and is denoted by a"1. (In the case of an additive group it is called the negative of a, and denoted by — a. We then write 6 - a for 6 + (- a).)
1. GROUPS 5 We now show that the equations ax = 6, ya = b, where a, b are any elements of the group, serve to define x and y uniquely. For x = ex = a~lax = a~16, and i/ = ye = j/oa^1 = 6a""1. Hence # and y are determined explicitly. In particular the equation a~xx = e has a unique solution. But a^a = e. The solution is therefore x = a. Therefore (a-*)-* = a. In the case of an additive group this becomes — ( — a) = a. A non-vacuous subset s of S may, with the law of composition assumed for the elements of S, also form a group. This is called a subgroup of the given group. The following conditions are evidently necessary and sufficient for the elements of s to form a subgroup: (i) if 8 contains elements a, 6, it contains ab; (ii) if 8 contains an element a, it also contains a*1. We conclude this section with a brief reference to an evident generalisation of the example of a group described above, namely, the symmetric group whose elements are the permutations of the numbers 1,2,3,...,n. The law of composition is defined as in the example. An important subgroup of this group is the alternating group. To define the alternating group we assume the definition of polynomials in the n indeterminates xv ...9xn which will be given in § 5. Consider the polynomial A = II (*<-**) (*'»*= 1,2,...,n). i<k If the suffixes 1,2,...,n are permuted, it is easily seen that the polynomial A is either unchanged (except for the order of the factors) or becomes — A. Permutations which leave A invariant are called even permutations, the others are called odd permutations.
6 I. RINGS AND FIELDS A transposition, that is, a permutation which interchanges two suffixes only, is seen to be an odd permutation. Any permutation can be regarded as the product of transpositions. Such a decomposition of a permutation into transpositions is not unique, but the parity of the number of transpositions for a given permutation is independent of the method of decomposition. The product of two even or two odd permutations is even; the product of an even and an odd permutation is an odd permutation. Hence the set of even permutations of the symmetric group forms a subgroup, which is called the alternating group, and the number of even permutations in the symmetric group is equal to the number of odd permutations. 2. Rings. A set of elements may have more than one law of composition. We shall be particularly concerned with sets having two laws of composition, under one of which the set forms a commutative group. We write this group as an additive group, and refer to the corresponding law as the addition law of the set. The zero of the group is called the zero of the set. The second law of composition is called the multiplication law, and the result of combining elements a, b of the set by this law is denoted by the product ab. Multiplication need not be commutative, but we shall require it to be associative. It is said to be distributive over addition if a(b + c) = ab + ae, and (b + c)a = ba + ca9 for all a, b, c in the set. A ring, then, is a set of elements with two laws of composition, addition and multiplication, with the properties: (i) the set is an additive group with respect to addition; (ii) multiplication is associative, and distributive over addition. The following examples will illustrate the various possibilities which may arise in the study of rings. In the first four cases the laws of composition for the elements involved are addition and multiplication as usually defined: I. The set of all complex numbers. II. The set of all integers, positive, zero, and negative. III. The set of all even integers. IV. The set of all integers, reduced modulo the integer m.
2. RINGS 7 V. The set of all matrices of q rows and columns whose elements are complex numbers. A matrix of this type is the array fan a12 . alq\ ^21 ^22 • &2q • • • • ^aql ag2 • aqq/ where each a{j is a complex number. Addition is defined by the rule /an a12 • ala\ /fin fi12 a21 a22 . <z2q I I /?21 /?22 \&ql <%q2 • &qq/ \Pql Hq2 • A« A*U + All «13 + As • "lq + Plq\ a21 + Al a22 + ^22 • a2q + 02q • • • . ^Pql+Pql <*q2 + fiq2 • <*qq +fiqq/ and multiplication by the rule /Ai As • Ai A2 a«i ag2 • aw/ \Ai A* • Aa/ v«i ^«2 where r«sS^«. fc-i The reader may easily verify that these sets, with the prescribed laws of addition and multiplication, form rings. Further study reveals certain features common to some of the rings, but not necessarily to all. (i) In the cases I, II, III and IV the zero of the ring is the number 0. In case V the zero is the zero matrix In all five cases for all elements a 0 = 1 ! a. of the ring (° (o ii 0 = ,• 0 0 • 0 = 0 = • • 1 • = 0 0 0 0, .a
8 I. RINGS AND FIELDS (ii) In cases I, II and IV let e be the integer 1, and in case V let ^0 0 . 1; in each of these cases e has the property given by the equations ae = a = ea for every element a of the ring. When such an element e exists, it will be shown to be unique. It is called the unity of the ring. In case III there is no element of the ring with this property. (iii) In all but the last case multiplication is commutative. In case V it is non-commutative, since, if q > 1, . 0 and 0 0 0 (iv) In cases I, II and III, if a and b are two elements of the ring such that , ~ ab = 0, then either a = 0 or b = 0. This property holds for the ring IV if and only if m is a prime number. If m = pq, where neither p nor q is 1, then p and q are two non-zero elements of the ring such that pq = 0. On the other hand, if m is prime and ab = 0, so that ab = cm, it follows, by the unique factorisation properties of ordinary integers, that either a or 6 is divisible by m. The property we are discussing does not hold in case V if q > 1. For „ A A, #A Q t ox = 0> 1 ^0 0 , 0/ \0 0 , 0j and neither factor is the zero of the ring.
2. RINGS 9 (v) In case I there is associated with every element a of the ring, other than the zero, a unique inverse ar1 such that aa*1 = arxa = e. The rings II and V do not have this property, and it can be shown quite simply that the ring IV has the property only when m is a prime number. The property is meaningless in case III since the ring has not unity, (vi) If we write a + a = 2a, a + 2a = 3a, etc., the elements a, 2a, 3a,... (a + O) are all distinct, except in case IV, when ma = 0 for all elements a in the ring. It will be observed that only in (i) did we have a property common to the five rings, namely, the zero has the property aO = 0 = Oa. We now show that this property holds for all rings, and deduce other elementary results true for any ring. Let a, b be any two elements of a ring i2, and let 0 be the zero of R. Then a + 0 = a, and therefore ba = 6 (a + 0) = ba + 60, and also ab = (a + 0) b = ab + 06. From the uniqueness of the zero of a ring it follows that 60 = 0 = 06, these equations holding for any element 6 in R. Again, we know that any equation a + x = 6 has a unique solution x = 6 + (-a) = 6 —a in R. Now a(6-c) + ac = a(b-c + c) = a6. Hence a(6 — c) = a6 — ac, and similarly (6 - c) a = 6a - ca.
10 I. RINGS AND FIELDS Therefore, taking 6 = 0, we obtain the equations a( — c) = —acf { — c)a = —ca for all a, c in B. Again, (-a)(-6)-a& = (-a)(-6) + (-a)6 = (-a)(-6 + &) = (-a)0 = 0, and therefore (— a) (— b) = aft. We have thus shown that the usual multiplicative properties of the minus sign hold in any ring. We now define the relationship between two rings known as isomorphism. Let B and i2* be two rings such that to each element a of B there corresponds a unique element a* of i2*, and such that any element a* of B* arises from exactly one element a of B. Such a correspondence is said to be one-to-one. Now suppose, in addition, that the correspondence is such that if a, 6 correspond respectively to a*, &*, then a + b and ab correspond respectively to a*+ 6* and to a*6*. The correspondence is then called an isomorphism. Isomorphism between rings is a relation of the class known as equivalence relations. Consider any set 8 of elements a, yff, y,..., and let there be a relation, which we denote by ~, between the elements of S, so that, given any two elements a, ft, we know whether a~ p is true or false. If the relation ~ is: (i) reflexive, that is, a ~a for all a in 8; (ii) symmetric, that is, a ~ ft implies f}~ a; (iii) transitive, that is, if a ~ /? and /? ~ y, then a ~ y\ we say it is an equivalence relation. An equivalence relation between the elements of a set 8 divides 8 into subsets, no two of which have any elements in common. If a, fi are in the same subset, <z~fi. Every element of 8 lies in one of these subsets. It is clear that if S is the set of all rings, and if a ~/? means that the ring a is isomorphic with the ring /?, the relation is an equivalence relation. We shall often speak of two isomorphic rings as being equivalent, implying that if in our discussion we replace one ring by the other (making any necessary consequential substitutions) nothing in our conclusions is altered.
2. RINGS 11 A subset S of the elements of a ring R is said to form a subring of R if the elements of S form a ring under the addition and multiplication laws of B. For this to be so it is necessary and sufficient that if a and b are any two elements of S, then a — 6 and ab belong to £. If S is a subring of 22, B is said to be an extension of 8. The following theorem is frequently used: Theorem I. If A and JB* are two rings, and A is isomorphic with a subring B of J5*, there exists an extension A* of A which is isomorphic with J5*, this isomorphism including that between A and B. Let C be a set of elements in a one-to-one correspondence jTwith those elements of J3* which are not in B, and let A* consist of the elements of A and C. We have merely to define addition and multiplication in A* so that A* becomes a ring with the required properties. Let al9a2,... be elements of A*. To each at there corresponds a unique element bt in J3*. If at is in A, bt is the element of B corresponding to ai in the isomorphism between A and B. If ai is in C, bt is that element of the set of elements of J5* not in B which corresponds to ai in the correspondence T. Conversely, to any element bi of J5* there corresponds a unique element ai of A*. Now, ifal9 a2 are any two elements of A*, we define ax + a2 and axa2 as follows. To ax and a2 there correspond bx and b2 in J5*. Let bx + b2 = 68, bxb2 = 64, and let a3 and a4 be the elements of A* which correspond to 63 and 64. We then define ax + a2 = a3, axa2 == a4. Clearly A * is a ring isomorphic with B*. Moreover, the construction shows that A is a subring of -4*. Hence A* has all the required properties. We conclude this section by describing a particular type of sub- ring which we shall meet in later chapters. Let R be any ring, I a subset of R with the following properties: (i) if a, b are in I, then a — b is in I; (ii) if a is in /, r in R, then ra is in I. We call I a left-hand ideal in i2; a right-hand ideal is defined by reading ar instead of ra in (ii). The two ideals coincide when multiplication in jR is commutative.
12 I. RINGS AND FIELDS To show that if I is not vacuous it is a subring of R we remark that if r is in I conditions (i) and (ii) coincide with the conditions stated above for a subring. On the other hand, not all subrings are ideals, since the condition (ii) need not hold. R itself is an ideal in R, and is called the unit ideal. The subset of R which consists only of the zero of R is also an ideal. These two ideals are usually called the improper ideals of R. A ring containing a proper ideal is the ring of integers (Example II, above). If m is any integer greater than 1, the set I of all integers of the form + ma, where a is an integer, clearly forms a proper ideal in this ring. The existence of an ideal I in R enables us to define an equivalence relation in R. If a, 6 are two elements of R, we write a = 6 (I) if a —6 is in I. The reader may easily verify that this is an equivalence relation. It is a type of equivalence which will often appear subsequently. 3. Classification of rings. Rings which are subject to no other conditions than those imposed by the definition given in §2 are too general to be of much interest in algebraic geometry. We limit the rings with which we deal by imposing fresh conditions which are, in fact, equivalent to requiring the rings to possess one or more of the properties noted in the remarks (ii), ..., (vi) on the examples of rings given in § 2. Different restrictions lead to rings with quite different properties. We now consider some of the more important types of rings. We saw that a ring R may or may not possess an element c with the property r r j ea = a = ae for every element a in R. We now show that R cannot have two distinct elements e,/ with this property. For, if ea = a, bf = 6, for all elements a, b in R, we may take a = /, b = e, and obtain the equations , _ « If this unique element e exists in R, it is called the unity of jB. All the rings we shall consider will be rings with unity, unless the contrary is stated explicitly. For the present we continue to denote the unity by e.
3. CLASSIFICATION OF RINGS 13 If a is any element of a ring J2 with unity, there may or there may not exist an element a*1 of R such that aa*1 = e = a^a. There cannot be two such inverse elements, for if ax = e, then a~"xax = a^e, that is, x = a"1. When such an inverse element a""1 exists, a is said to be regular, or to be a unit of R. Clearly e and — e are units of R. On the other hand, the zero 0 is not a unit, since for every element a of R a0 = 0. Again, we have seen [§ 2, (iv)] that there are rings R in which there exist pairs of elements a, b, neither of which is zero, such that ab = 0. We call a a left-hand divisor of zero, b a right-hand divisor of zero. The divisor of zero a cannot be regular. For, if a has an inverse a-1, 6 = a~xab = a~x0 = 0, contradicting our assumption that 6 is not zero. Similarly, 6 cannot be regular. The two types of ring with which we shall be particularly concerned are the integral domain and the field. An integral domain is defined as a ring with unity, in which multiplication is commutative, and which has no divisors of zero. Examples I, II, and IV (in the case m prime) of § 2 are integral domains. A field is defined as a ring with unity in which every non-zero element is regular. The ring of all complex numbers is clearly a field. By the result proved above a field cannot contain any divisors of zero. A commutative field—that is, a field in which multiplication is commutative—is an integral domain. We shall eventually be able to confine our attention to commutative rings, but it is convenient to obtain certain results which are valid for non-commutative rings, as these will be used in an analysis of the postulates on which projective geometry is usually based. Therefore, except when we say so expUcitly, we shall not assume our rings to be commutative.
14 I, RINGS AND FIELDS In a ring R consider any element a. Write a + a = 2a, a + 2a = 3a, ..., noting that na is not a product, but the sum of n elements each equal to a. We have seen by an example that there exists a ring R for which we can find an integer m such that ma = 0 for all elements a in R. Now, let R be a ring with unity e, and suppose that for some positive integer n, __ 0 Let m be the smallest positive integer satisfying this condition. Then, first of all, if a is any element of R, ma = a + a+ ...+a = ea + ea + ... + ea = (e + e+...+e)a = (me) a = 0. If m is not prime, and m = pq, then #e=#0, je + O, but peqe = pqee = me = 0, and R has divisors of zero. The number m is called the characteristic of the ring R. We have proved that the characteristic (if there is one) of an integral domain or a field is a prime number. If, for all non-zero integers n, ne + O, the ring is without characteristic. (It is then often said to be of characteristic zero, but this terminology is not strictly correct.) If a ring R with unity has characteristic m, the subset 0, e, 2e, ..., (m —l)e forms a subring. For ae — be = (a — 6) e = ce, where c = a — b (modulo m\ and (ae) (be) = de, where d = ab (modulo m).
3. CLASSIFICATION OF RINGS 15 Thus R contains a subring isomorphic with the ring of integers reduced modulo m. By Th. I, § 2, there is an extension of this ring isomorphic with R. Similarly, if R is without characteristic, it is isomorphic with an extension of the ring of natural integers. It follows that a ring without characteristic contains an infinite number of elements. In each case we identify R with its isomorph, and so in future we shall write the unity of the integral domain or field we are considering as 1. Later on we shall confine ourselves to integral domains and fields without characteristic, but at present we do not make this restriction. 4. The quotient field of an integral domain. It is a familiar result that the integral domain of the natural integers is embedded in a field, namely, the field of rational numbers. This is a special case of a general theorem which we shall frequently use. Theorem I. Given an integral domain I, there exists a commutative field K which, regarded as a ring, contains I as a subring. The construction which we give for K gives the minimum field having the required property. This minimum field is defined uniquely, to within isomorphism. The elements of I are denoted by a, 6, c,..., the zero and unity by 0 and 1. Consider the set of all ordered pairs (a, 6) in which 6#0, and the relation denoted by ~ between pairs where (a,b)~(c,d) stands for the equation ad — bc = 0. Since multiplication is commutative, it follows at once that this relation is reflexive and symmetric. It is also transitive. For, if (a,b)~(c,d) and (c,d)~(e,/), ad = be and c/ = de. Therefore adf = bef = bde, and so # d(af— be) = 0. From our definition of an ordered pair, d is not zero. Since an integral domain contains no divisors of zero it follows that af = be, that is, (a>yb)~(e,f).
16 I. RINGS AND FIELDS The relation is therefore an equivalence relation, and divides all pairs under consideration into classes. The class to which (a, 6) belongs will be denoted indifferently by [a, 6] or [c, d], where (a,6)~(c,d). We define the sum of a pair {a, b) and a pair (c, d) by the equation {a, b) + (c, d) = {ad + be, bd). By hypothesis, b and d are not zero, and therefore bd =f= 0. The pair denoted as the sum is therefore admissible. It is evident that addition of pairs is commutative. Now let {a, 6) -(a', 6'), {c,d)~{cf ,d'). Then (a', 6') + (c', d') = (a'd' + 6V, 6'd') - {ad + be, bd), since {ad + 6c) 6'd' - {a'd1 + 6'c') 6d = {aV - a'6) dtf' + (cd' - c'd) 66' = 0. Hence we may define the addition of classes by the equation [a, 6] + [c, d] = [ad + be, bd]. Since {[a, b] + [c, d]) + [e, /] - [adf+ bcf+ bde, bdf] = [a,b] + {[c,d] + [e,f]), addition is associative. Also, the equation [a,b] + z= [c,d] has the solution x = [6c — ad,bd]. Hence the classes form a commutative group under addition. The zero of the additive group is [0,1] = [0, a], where a is any non-zero element of I. We now define multiplication of pairs by the equation {a,b){c,d) = {ac,bd). Multiplication is clearly commutative. Also, if (a, 6) ~{a', V), {c,d)~{c',d')y then {a'c', b'd') ~ (ac, bd), since a'c'bd — ac6'd' - {a'b-ab')c'd + {c'd-cd')ab' »0.
4. THE QUOTIENT FIELD OF AN INTEGRAL DOMAIN 17 Thus we can define multiplication of classes by the equation [a, 6] [c, d] = [ac, bd]. Multiphcation can immediately be proved to be associative. Again, [a, b] ([c, d] + [e, /]) = [a, 6] [cf+ de, df] = [acf+ade,bdf] = [abcf+abde,b2df] = [ac, bd\ + [ae, 6/] = [a,b][c,d] + [a,b][e,f]. Hence multiphcation is distributive over addition. The set of classes of equivalent pairs, with addition and multiphcation defined as above, therefore forms a commutative ring If*. If a is any non-zero element of /, (a,a)~(l,l), and so [«,«] = [1,1]. Since [a, b] [1,1] = [a, 6], [1,1] is the unity of K*. Finally, if [a, 6] is any non-zero element of K*, both a and b are non-zero. Therefore [6, a] is in K*. But [6,a][a,6] = [o6,a6] = [l,l]. Therefore every non-zero element of K* has an inverse. Thus K* is a commutative field. Now consider the elements of K* which can be written as [a, 1], and let them form the subset 8. Since [a,l]-[6,l] = [a-6,l], and [a, 1] [6,1] = [ab, 1], S forms a subring of K* isomorphic with I. By § 2, Th. I, there is an extension K of I which is isomorphic with If*, and which is therefore a field. The element of K corresponding to [a, b] of If* is usually written a/b. If a _ c 6~^ then orf = 6c. We also have a — = a. 1 HP I 2
18 I. RINGS AND FIELDS The field K is the one required. It is usually called the quotient field of I. To complete the proof of our theorem we now show that any commutative field K' which contains I as a subring contains a subfield isomorphic with K. We first observe that, by the definition of a field, the non-zero elements of Kf form a group under multiplication. Hence the equation bx = a (6 # 0), where a, 6 are in K'y has a unique solution in K'. We now show that the equations bx = a, dx = c, where a, 6, c, d are in I, and 6 and d are not zero, have the same solution in K' if and only if ad = be. First, if this condition is satisfied, and # is a solution of the first equation, we have bdx = ad = be. Therefore b(dx — e) = 0, that is, dx = c. Secondly, if the two equations have the same solution x, then ad = bdx = be. Thus to every element x of K' which satisfies an equation bx = a, where a, b are in /, and b # 0, there corresponds a unique element afb of If, and conversely. Moreover, if x, y correspond to a/6, c/d respectively, then bx = a, dt/ = e. and therefore d6(a; + y) = od + 6c, and 6d#2/ = a6. Hence a? + y and o^/ correspond to r + i and y-z respectively. The field K' therefore contains a subfield isomorphic with K.
4. THE QUOTIENT FIELD OF AN INTEGRAL DOMAIN 19 From this theorem we deduce that K is the smallest field which contains I as a subring. Finally, we note that if I is of characteristic p, the sum to p terms 1 + 1 + ... + 1 = 0. Hence [1,1] + [1, l]+... + [l, 1] = 0, and it follows that K is of characteristic p. On the other hand, if I is without characteristic, [l,l] + [l,l]+... + [l,l] = [l + l + ... + l,l]#[0,l], and hence K is without characteristic. 5. Polynomial rings. We now discuss a method of extending a ring by the adjunction of an element x, called an indeterminate. The only case of any importance which we shall meet in this book is that in which the ring is an integral domain, and it will make for simplicity if we assume throughout this section that the basic ring from which we start is an integral domain 7. Let us consider the ordered sets (aQ,av...,ar) which consist of a finite number of elements ai of I. Two sets {aQ,al9...,ar)9 (b0,bv ...,b8) are said to be equivalent if «*< = &< (i== l,2,...,min(r,a)), a. = 0 (i = 5+1,...,r) when r>8> bt = 0 (i = r+1, ...,5) when s>r. This relation is evidently reflexive, symmetric and transitive, and is therefore an equivalence relation. We define the addition of two sets by the equation (a0,av ...,ar) + (60, bl9..., b8) = (c0, cv ..., ct), where t = max (r, s), and ct = at + 6i5 defining a{ = 0 for i > r and 6i = 0 for i > 8. It is easily seen that if either of the two sets is 2-2
20 I. RINGS AND FIELDS replaced by an equivalent set, the sum is replaced by an equivalent set. Multiplication of two sets is defined by the equation (a0, av..., ar) (60, bl9..., b8) = (d0,dl9..., dr+8), where d< = a<60 + a<_161+...+^6^ (i = 0, ...,r + s), and, as above, we define at = 0 for i > r, and bi = 0 for i > s. Again, the replacement of either set by an equivalent set converts the product set into an equivalent set. If we now consider the classes of ordered sets defined by the equivalence relation, we see that they form an additive group with respect to addition, the zero of the group being all sets equivalent to the set (0); also, multiplication is associative, and distributive over addition. Hence our classes of equivalent sets, with the two laws of composition we have defined, form a ring R, and it is easily seen that (1) is the unity of R. This ring R contains a subring which is isomorphic to I. The elements of this subring are (a). For (a) + (6) = (a + 6), and (a) (b) = (ab). Hence an isomorphism is set up if (a) is mapped on the element a of I. Let us now consider the set (0,1). We see that (0,1)2 =(0,1)(0,1) = (0,0,1), (0,1)3 =(0,1)2 (0,1) = (0,0,0,1), • ••••• (0,1)^=(0,1)^(0,1) = (0,0,0,...,1), where in the last set written the unity is preceded by n zeros. Since (a) (0,1)» = (0,0,0,..., a), we may express the set (a0,av ...,ar) as follows: (a0,^,...,af) = (a0) + ^ Now let us denote by I[x] the ring which is an extension of I and which is isomorphic with R [§2, Th. I], x being the element of the extension which is mapped on the set (0,1) of J?. Then, from the above reasoning, the set (a0,al9...,ar) of R is mapped on the element a0 -+-axx + a2x2 + ... -+-arxr of I[x\. These elements are called polynomials, with coefficients a0,av ...,ar. If ar#= 0, the polynomial a0 + aix + • • • + ar& iQ sai(* to ^6 °f degree r.
6. POLYNOMIAL RINGS 21 We note that the zero (0) of R is mapped on the zero of 7, and therefore the equation a^ + axx+...+arsif = 0 (1) is satisfied if and only if a0 = a± = ... = ar = 0. Because of this property of the element x, it is said to be an indeterminate over I. In the representation of a polynomial we usually omit any term a^1 in which at = 0, whether i exceeds or is less than the degree of the polynomial. In this sense anxn is a polynomial whose complete representation is 0 + Ox + ... + Ox11*1 + anxn. Furthermore, a0 + a1x1+... +arof can be regarded as the sum of the r + 1 polynomials o> 0/]X) ..., a^x . Since also a^ is the product of at and xi9 and since for the polynomials x,x2}... we have the multiplication law xi>xv = xp+<*, the ring I\x] is obtained by introducing a new element x among the elements of I which can be multipUed commutatively by any element of I. We say that I[x] is obtained from I by adjoining the indeterminate x to I. I[x] is also called the ring of polynomials in x over I. These results apply when I is any commutative ring with unity. We now proceed to draw conclusions concerning the ring I[x] from the fact that I is an integral domain. Theorem I. If I is an integral domain, so is I[x\. We have already seen that I[x] is a commutative ring with unity. We must now show that I[x] has no divisors of zero. If f=a0 + a1x+...+amxm9 and g = b0 + b±x+ ... +bnxn are two non-zero polynomials, we may assume am #= 0, bn # 0. Then fg = c0 + cxx+ ...+cm+nx™+n, where cm+n = am6n=i=0, since I has no divisors of zero. Hence /gr#=0. In other words, if fg = 0, and /#0, we must have g = 0. Therefore I[x] has no divisors of zero.
22 I. RINGS AND FIELDS The process of adjoining an indeterminate to an integral domain to obtain another, more extensive, integral domain may be repeated. If J = I[xx], J[x2] is an integral domain. Its elements are polynomials , , , n where a€ = ai0 + aaxx+...+ aimx^ is in I\x^\y and m is the maximum of the degrees of a0,..., an. The elements a{i lie in I. Hence, the elements of J[x2] can be written in the form n m in which each term a{jxlx{ is the commutative product of aij9 x\, x{. If J' is the integral domain I[x2], the elements of Jf[x^\ are of the form m n and the integral domains J[x2] and /'[a;!] are easily proved to be equivalent. We denote this integral domain by I[xvx2]. Proceeding in this way, we can adjoin indeterminates xl9x2, ...,xt in turn to I. The resulting ring I[xv ...,xr] is independent of the order in which the indeterminates are adjoined. For the elements of this ring are of the form r fh s <!-<) nr ... S aiit ir-0 ,..<,*ll- the term written explicitly being the commutative product of aii...ir> xi> •••> x%- Also, I[xv ...,xr] is an integral domain. We have already proved this result when one indeterminate is adjoined to I. We assume it is true when r— 1 indeterminates are adjoined. The result is assumed true for J = I[xv ...,av_J, and by Th. I is therefore true for J[xr]. But J[xr] = I[xv ...,&r]. Hence we have Theorem II. If I is an integral domain, so is I[xv ...,xr]. We conclude this section by introducing an important operation on polynomials in I[xv ...,xr] known as specialisation. Let be any polynomial in our integral domain. We select any number j (0^j<r) of the indeterminates xly...,xry say, for example,
6. POLYNOMIAL RINGS 23 xv ..., Xj. Let a19..., o^ be any elements of I. Then if in / we replace xly..., Xj by ax,..., a^ respectively, / becomes /•- "£' ...S6i( ^{...^, where . 5» ** i i is in /. (a* is, of course, the product of i factors each equal to a.) Then / is said to specialise to a polynomial /* in I\xi+V ...,xr]. We use the notation to describe the specialisation. By using the fact that equations such as f+9 = h9 fg = k express relations between the coefl&cients of/, gy h and k which do not involve the ^determinates xv ..., #r, we see at once that if, for any specialisation of the indeterminates, /->/*> g->g*> h->h*y k->k*, then /* + gr* = A*, /*gr* = &*. The converse of this result is not, of course, true. Note. While we do not have to consider polynomials whose coefl&cients are in a non-commutative ring, it should be pointed out that a theory of such polynomials exists. The only result of this theory which we shall use is the following. Let f=a0xm + a^™-1 +... + am, where at is in a non-commutative ring K, x is indeterminate and is assumed to commute with every element of K. Then if a is any element of K, there exist polynomials p = box™"1 +...+ bm_l9 q = c^-1 +...+ cm_x, and elements 6m, cm of K such that /= (:^)(60^-1+...+6^^ = (Co*™-1 +...+ cm_x) (x - a) + cm. Indeed, we need only take b{ = a<a0 + a'-S +...+ ai9\
24 I. RINGS AND FIELDS 6. The division algorithm. In this section I denotes an integral domain, K a commutative field, and I[x]9 K[x] the rings obtained by adjoining the indeterminate x to Iy K respectively. Let / = a0 + a±x+ ...+amxmy g = 60 +&!#+...+ 6n#n (n^m) be two polynomials in I[x] of degrees my n respectively, and let us suppose, in the first place, that bn = 1. Then, clearly, is a polynomial in I[x] of degree m1<m. If n ^ mly and if then A=A-a'mlxmi-n9 = f-{amx™-n + a'mx™>-")g is a polynomial of degree m2<mv Proceeding in this way we arrive, in all cases, at a polynomial q of degree m — n such that is a polynomial of degree less than n, the degree of g. This result can be generalised to include the case in which bn is a regular element of I. For, if b~1bn = 1, 9' = b^g is a polynomial of degree n in which the term in xn has coefl&cient unity. Then there exists a polynomial q' such that r=/-jV=/-(jV)» is of degree less than n. We now show that the polynomials qy r obtained in this way are unique. For if q'y r' are any two polynomials such that r' —f—q'g is of degree less than n, we have the equation r~-/ = (q'-q)g. But the degree of r — r' is less than ny and the degree of (q' — q)g is at least ny unless q9 = q. Hence, unless q' = qy we have a polynomial of degree less than n equal to a polynomial of degree at least n. This is impossible, and therefore q = q\ and consequently r = r'.
6. THE DIVISION ALGORITHM 25 The process given above for determining q and r is called the division algorithm. A special case of the result just proved is the Remainder Theorem. If c is any element of I, take g = x — c. Then there exists a polynomial a such that * , x f-q{x-c) = r is in I. This equation remains true when we make the specialisation x->c. Hence £l x where/(c) is the result of replacing x by c in/. Returning to the division algorithm, it may happen that r = 0. We then have three non-zero polynomials/, g9 q in I[x] such that <7 and g are said to be divisors of/, and / is a multiple of ^ and of g. If the degrees of both g and q are greater than zero, both g and g are said to be proper divisors of/. We now suppose that I is a field K. Since every non-zero element of K is regular, the division algorithm can be applied to any pair of polynomials/0,/i of K[x], By repeated applications of the division algorithm we shall construct a sequence of polynomials/2,/3, .... We shall denote the degree of/^ by miy and suppose that m0^mv Our purpose is to obtain information on the common divisors of/0 a,nd/1. Applying the algorithm to /0,/1, we get the equation /0 = ^1/1+/2^ where m2<mv If/2 is not zero we apply the algorithm to /x,/a, obtaining the equation /1 = ¢2/2 +/3» and proceed thus until we have J8-1 = ?*/s+/*+i> where /8+1 is zero. We reach this stage after at most mx +1 applications of the algorithm. Now, let g be a divisor of/0 and of fv Then, if /0 = ho9> and /x = hxgf we have /2 = h2g, where h2 = h0 — gx Ax.
26 I. RINGS AND FIELDS Hence /3 = hzg, where hz = h± — q2h2, aM so on. Finally . , fa = M> where h8 = *^a - ga-1 h8_v Thus any common divisor of/0 and/^ is a divisor of the set of polynomials/2, ..., /8. Conversely, let g be a divisor of/8. Then since /• = hs9> we have f8„x = qsf8 = q8h8g = h^g, say, /8_2 = (?mjm + *«)? = V2^ say, and so on, until finally we obtain /1 = *ift and /0 = ^0gr. Thus any divisor of/8 is a common divisor of/0 and/^ Since any common divisor of/0 and/x is a divisor of/8, the degree of such a divisor g cannot exceed ra8, the degree of/8. On the other hand, f8 is itself a divisor of/8, and therefore of/0 and fv These polynomials therefore have a common divisor of degree ra8, and have no common divisor of higher degree. We call /8, or a/8, where a is any non-zero element of K, the highest common divisor of/0 and/^ The division algorithm enables us to express f8 in terms of/0, fx as follows. We have the equations /2 =/o"~3i/i> /3 = /1"~ #2/2 = -Qjo + (l+9i9*)fi> and by a simple induction we find that fe = *fo + bfi, where a, b are polynomials in K[x]. If/0, fx have no divisors of degree greater than zero, /0, f± are said to be relatively prime elements of K[x], Then ms = 0, and fs is a non-zero element of K. Let its inverse be g. Then 1 = f*g = <*9fo + bgfv that is, l**Af0 + Bfv where A, B are polynomials in K[x].
6. THE DIVISION ALGORITHM 27 A non-zero element of K[x] which has no proper divisors is called a prime element of K[x], or an irreducible polynomial. Our work above now enables us to prove Theorem I. Iff is an irreducible polynomial in K[x], and g, h are polynomials in K[x] such that f is not a divisor of g, butf is a divisor of gh, then f must be a divisor of h. Since/ is not a divisor of g, the degree of any common divisor of these two polynomials is less than the degree of/. But/is irreducible. Therefore the only divisors of/have degree zero. Hence/ and g are relatively prime, and so we can find polynomials a, 6 such that 1 =af+bg. Therefore h = afh + bgh. But/divides gh. Therefore, for some polynomial k9 gh=fk. Hence h = (ah + bk)f, showing that/ is a divisor of h. We note here that the term factor is often used in the same sense as divisor, so that we may refer to the highest common factor of two polynomials instead of to their highest common divisor. We conclude this section with a theorem of frequent application. Let K* be an extension of the field K, and suppose that /, g are polynomials in K[x], and therefore in K*[x]. If/, g are regarded as polynomials in K*[x], the process of the division algorithm, when applied to/and g, is carried out entirely in the subring K[x]. Hence the process of finding the highest common factor of / and g in K*[x] can be carried out entirely in K[x]. Thus the highest common factor of/ and g in K[x] is the highest common factor in K*[x], and, conversely, the highest common factor of/ and g in K*[x], when multiplied by a suitable element of jBl*, is in K[x], and is the highest common factor of/and g in this ring. Hence, in particular, we obtain Theorem II. If K* is an extension field of the field K, and iff, g are two polynomials in K[x] which are relatively prime, then f, g are still relatively prime when regarded as polynomials in K *[#]. 7. Factorisation in an integral domain. We have defined a unit or regular element of an integral domain I to be an element a which has an inverse a"1 in I, so that aa*1 = 1. We saw that the inverse is necessarily unique.
28 I. RINGS AND FIELDS Clearly, 1 and — 1 are always units of I. There exist integral domains, such as the ring of natural integers, with no other units. If, however, the integral domain is a field, every non-zero element m it is a unit. We shall make use of the following results in the sequel: Theorem I. The product of any finite number of units is a unit. If the product of any finite number of elements of I is a unit, each element in the product is a unit. Let ev ..., er be units, ef \ ..., e~x their respective inverses. Then ci ... c*.cJ ... cy. == ti c^ Co So ... Sr Sj, == A • Hence e±... er is a unit. If a±... ar = e is a unit, then a± ...ars-x = 1. That is, ai(a1... ^-1^^+1... ar) = 1. Therefore ai is a unit. We now define the term factor. If any non-zero element a of an integral domain I can be written as a = 6c, (1) where 6, c are in 7, these elements are said to be factors of a, and a is said to be a multiple of b and c. When a is written in the form (1) it is said to be decomposed into the factors b and c. If one of the factors is a unit, the decomposition is said to be trivial. Since every I has at least one unit ev every element of I has at least one trivial decomposition a = e^e^a). An element of I which is not a unit may or may not have a non- trivial decomposition. If it has only trivial decompositions it is called a prime element of I. (If I is the polynomial ring K[x], the prime elements are the irreducible polynomials of degree greater than zero.) We prove that if any product a1...arof elements in I is a prime, then one of the factors is a prime, and the rest are units. Forif 6 = ^...^, and each ai is a unit, so is b [Th. I]. This is contrary to the definition of a prime. Hence one ai at least, suppose it is av cannot be a unit. Then since b is prime, b = a^ _aj can only be a trivial decomposition. But ax is not a unit. It follows that a2...ar is a unit, and therefore, by Th. I, that a2,az,...,ar are all units.
7. FACTORISATION IN AN INTEGRAL DOMAIN 29 An element a of I is said to be completely factorised if it is expressed as the finite product a = ax... ar, where each ai is either a unit or a prime. By Th. I we can combine all the factors which are units into a single one. We then write this first, using a Greek letter to denote it, so that a — ouax ...ar implies that a is a unit, av ...,ar are primes. In particular cases, of course, a may be unity. Two complete factorisations of the same element a = ouax ... ar, a = ^bx ...b8 are said to be equivalent if r = s, and for a suitable reordering of the factors , /., x bi^e^t (i = 1,..., r), where ev ...,er are units. We now give the following definition: If an integral domain is such that every non-zero element can be completely factorised, and all complete factorisations of a given element are equivalent, it is called a unique factorisation domain. We shall abbreviate this to u.f.d. The ring of the ordinary integers is a well-known example of a u.f.d. But we now give an example of a ring, in which not all elements are units, which is not a u.f.d. Consider the set S of complex numbers which can be written in the form where a and /? are integers. It is immediately verified that this set is closed under the operations of addition and multiplication, and it is easily shown that it forms an integral domain. This integral domain is, of course, a subring of the ring of complex numbers. We define the norm N(a) of a by the rule N(a) is always a non-negative integer. The following properties are rapidly verifiable: (i) N(a) = 0 if and only if a = 0; N(a) = 1 if and only if a = ± 1; N(a) #= 2 for any a in 8; N(a) = 3 if and only if a = ±V-3; N(a) = 4 if and only ifa = ±2, ora = ±l± ^- 3.
30 I. RINGS AND FIELDS (ii) N(a)N(b) = N(ab). Since JV(1) == 1, we deduce that if aft = 1, N(a)N(b) = 1. Hence N(a) = N(b) = 1, that is, by (i), a = ± 1, 6 = +1. Therefore the only units of the ring are + 1 and — 1. Again, if ° a = a1...ar is a factorisation of a in which no factor is a unit, it follows from the fact that Tij/ \ vr/ \ tit/ x N{a) = NiaJ.^Nia,.) that r cannot exceed the number of prime factors of N(a). It follows from this that a can be completely factorised. (iii) The numbers ±2, ± 1 ± ^/- 3 are all primes. For if any one of them could be written as be, where neither b nor c is a unit, we should have ,T/LX ,T/ x N(b)N(c) = 4, while N(b)>l and N(c)>h Therefore N(b) = N(c) = 2, and we saw in (i) that there are no elements in the ring with norm equal to 2. The elements in question are therefore all primes. Now, since the only units are + 1 and — 1, we cannot have the equation l ± V-3 = e2> where e is a unit. Therefore the factorisations (2)(2) = 4 = (l + V-3)(l-V-3) are not equivalent factorisations of 4. The ring 8 is therefore not a u.f.d. We now return to the integral domain I, and assume that it is a u.f.d. We prove that any two elements a, 6 of I have a unique highest common factor. First of all let us call two primes p, q equivalent if there exists a unit e such that p = eq. Now let us examine the consequences of assuming that an element c is a factor of an element d, so that d = ce.
7. FACTORISATION IN AN INTEGRAL DOMAIN 31 Let c=*yc1...cri d = Sd1 ...d8, e = ee1... et x be complete factorisations of c, d, e in which, as we agreed, 7, 8, e denote units, all the remaining factors being primes. Then from the equation *, , ^ od1...d8 = 76¾...^¾ ...¾ we deduce that r +1 = 5, and that the primes cl9..., cr are equivalent to r distinct primes of the set dl9..., d8. Now let , ou , be the complete factorisations of two elements a, 6 of I, and suppose that, by suitably choosing the arrangement of the factors, and the unit multipliers of the primes, we can say that a< = 6i (*'= 1, ...,*), but ty and bk are not equivalent for any j and k greater than t. By what has been proved above, any factor of a is the product of a unit and a certain number of the aif and similarly, any factor of b is the product of a unit and a certain number of the b{. From the fact that I is a u.f.d. it follows that if d is a factor of both a and 6, d can be written , * a = oaix...aiv where 1 ^ ix < i2... < ii < t. Hence d is a factor of D = ax... ai9 which is itself a factor of a and b. This element D is called the highest common factor of a and 6, and is uniquely defined, save for a unit factor. We now prove Theorem II. Ifav...,an are any n elements of I, there is an element d, determined to within a unit factor, which is a factor of each ai9 and is such that any factor common to all the ai is a factor of d. We shall call d the highest common factor of av...9an. The theorem has already been proved when n = 2. Let us suppose that it is true for 71— 1 elements of 2", av ...,an-1, and let dn be their highest common factor. Let d be the highest common factor of dn and an. Then d is a factor of an, and of dny which is itself a factor of al9...,an_v Hence d is a factor of al9 ...yan_van. Conversely, if c is any factor of al9 ...yan_van, it is a factor of dn and an9 and
32 I. RINGS AND FIELDS hence of d. It only remains to prove that d is unique. Now, if d' is another element of I with the properties of dy df is a factor of a-i^..., an, and therefore of d. Therefore d = ad', where a is in I. Similarly ,, _ ., where 6 is in 7, and hence , ,* d = abdf giving ab = 1, so that a and b are units. We conclude this section with the important Theorem III. If K is any commutative field, K[x] is a u.f.d. We first prove that any polynomial in K[x] of degree n > 0 can be written as the product of a finite number of prime factors. We proceed by induction on n. If n = 1 there is nothing to prove, since, if/is of degree 1, and where gy h are of degrees ly m respectively, l + m = 1. Hence either g or h is of degree zero, and is therefore a unit, and the decomposition is a trivial one; that is,/is prime. We therefore suppose the result true for polynomials of degree less than n. Now let / be of degree n. If it is irreducible, then it is already expressed as a product of prime factors. If it is reducible, where gy h are of degrees ly m, say, and I + m = n (l<ny m<n). By hypothesis g and h can be expressed as products of a finite number of prime factors. Hence/ can be so expressed. We now prove that all factorisations of/ are equivalent. That is, if/ is a polynomial of degree n which can be expressed in two ways as the product of prime factors /s«/i4 / = ^i-..^ (^/?in K)y we prove that r = sy and after a rearrangement of the factors, 9i = 7ifi (*= 1,...,r).
7. FACTORISATION IN AN INTEGRAL DOMAIN 33 We again proceed by induction on n. For n = 1 the result is trivial, since all polynomials of degree 1 are irreducible. Assume that the result is true for polynomials of degree less than n. Since <tfi-/r = £& - 0.. fx is a factor of gx or of fig2... g8 [§ 6, Th. I], If it is a factor of the second polynomial, the same theorem tells us that it is a factor of g2 or of fig3...g8. Continuing thus, f± is either a factor of some gi or of/?. This last conclusion is untenable, since/x has degree greater than zero, while/? is of degree zero. Hence, byre-ordering the factors gi9 if necessary, we can say that 9i = Yifv Since gx (as well as fx) is irreducible, yx must be a unit. K[x] is an integral domain, and so we can cancel the factor/1? obtaining the equation - £ 0 between two polynomials of degree less than n. By our induction hypothesis , , ,, , . Jr r—1=5—1, that is r = s, and, after a suitable arrangement of the factors, 9i = 7ifi (i = 2,...,r), where the yi are units. This proves the theorem. 8. Factorisation in polynomial rings. The main theorem of this section, which we shall prove with the help of three lemmas, is Theorem I. If I is a u.f.d., so is I[x], Let/ be any non-zero polynomial in I[x], and suppose that d is the highest common factor of the coefficients of/. We can write f=*dg9 where g is a polynomial in I[x] whose coefficients have a unit as their highest common factor. Such a polynomial is called a primitive polynomial. It is clear that both g and d are uniquely determined, save for unit factors. The importance of primitive polynomials is brought out by the following lemma. Lemma 1. The product of two primitive polynomials is a primitive polynomial. HP I 3
34 I. RINGS AND FIELDS Let / = a0 + axx + ...+anxnt and g = 60+6!»+ ... + bmxm be two primitive polynomials, and suppose that the product fg is not primitive. Then the coefl&cients of this polynomial have a common factor d which is not a unit. If p is a prime factor of d, p must divide all the coefl&cients of fg. Now, / is primitive, and so not all the coefl&cients of/ can be divisible by p. Let ar be the first coefficient which does not have p as a factor. Similarly, let b8 be the first coefficient of g not divisible by p. Consider the coefficient of af+* in the product/</. It is cr+8 = arb8 + ar-A+l + ar+l V-l + -- + «06r+« + ar+$bQ = arb8+pc, where c is in I. By hypothesis cr+8 has p as a factor, and therefore the product arb8 must be divisible by p. The prime factors of arb8, since I is a u.f.d., are the prime factors of ar and of bs. Therefore either ar or b8 is divisible by p, contrary to hypothesis. Our assumption that fg is not primitive is therefore false. Let K be the quotient field of /. By § 7, Th. Ill, we know that K[x] is a u.f.d. We use this theorem, together with properties of primitive polynomials, to establish our theorem. Let <p be any polynomial in K[x]9 and suppose that 0 = a0 + a1x+..t+anxnf wnere a€ = ai\bi (ai9 bt in I). We now write the coefl&cients in 0 with a common denominator 6 = b0...bn. If Ci = V--6*-iaA+i--A> and /= ¢0 + ^3+. ..+cnxnf then $ = //6, the polynomial/lying in I[x], If now / = <*g, where g is primitive, we have expressed cf> in the form * = b9-
8. FACTORISATION IN POLYNOMIAL RINGS 35 Lemma 2. The primitive polynomial g is uniquely determined by the non-zero polynomial <j>> save for a factor which is a unit of I. CI c Suppose that $ = j-g and also 0 = jh, where g and h are both primitive in I[x]. Then adg = bch, and therefore ad is a factor of the coefficient of every term in bch. But since h is primitive, the highest common factor of the coefficients is be. Hence , , be = eaa, where e is in I. Similarly, since be is a factor of adf ad = e'fcc, where e' is in I. Therefore bead = (eer)bcad. Since <f> is not zero, and I is an integral domain, ee' = 1, and therefore e is a unit. Hence and the lemma is proved. If now the polynomial \Jr in K[x] determines the primitive poly- /* nomial h, and \lr = -h, then a By Lemma 1, gh is primitive. Hence, by Lemma 2, the primitive polynomial determined by ¢^ is gh. We therefore obtain Lemma 3. If <f> is an irreducible polynomial in K[x], the corresponding primitive polynomial in I[x] is irreducible, and conversely. With these preparations we now come to the proof of our theorem. Let / be any polynomial in I[x], and let d be the highest common factor of the coefficients. Write f=dF, and consider F as a polynomial in K[x], which is a u.f.d. Let the prime factors of F in K[x] be 0X,..., 0r. These are uniquely determined, save for units of K. If V 3-2 *<-?**
36 I. RINGS AND FIELDS where Fi is the corresponding primitive polynomial in I[x], then, by the above lemma, Fi is irreducible in I[x]. We have o1...or that is, bx ...brF = ax... arFx ...Fr. Since F,FX, ...,Fr are primitive polynomials, 6x...6f is a factor of ax... ar, and conversely. Hence ax...ar — ebx... brf where e is a unit of J, and therefore Writing ed = 8dx... d8> where dv ...,d8 are primes, the equation f=8d1... d8F1...Fr (8 a unit) expresses/ as a product of prime factors. Now let / = eex... es> Ox... 0^ (ea unit) be another representation of / as a product of irreducible factors in I[x\. By Lemma 3 the Fi9 Gj are irreducible in K[x], and therefore, since K[x] is a u.f.d., we must have r = r', and (JL = a^Ft (a^ a unit of K), when the factors have been suitably arranged. By Lemma 2 a* = ^, where ei is a unit of I. Hence, finally, 8dx... ds = ecx... e^^ ... er, where # and eex ... er are units, whilst the d{, e^ are not units. But I is a u.f.d. Therefore s = 5', and after a suitable rearrangement of the factors, , a m -± r T\ dt = 0iei (^ a unit of I). This proves that I[x] is a u.f.d. By a simple process of induction we then deduce Theorem II. If K[xv ..., xr] is the ring obtairied by adjoining the r indeterminate^ xl9..., xr to the commutative field K, then K[xl9..., xr] is a u.f.d.
9. EXAMPLES OP FIELDS 37 9. Examples of fields. In constructing the spaces whose geometry we propose to discuss, the first step is always the selection of a field K. It is therefore appropriate to end this chapter with a few examples of fields. I. We have already noted that the complex numbers form a commutative field under the usual laws of addition and multiplication. II. We know that the natural integers form an integral domain 7. By § 4, this integral domain can be embedded in the quotient field of 2", which is just the field of rational numbers. Both these fields are commutative and without characteristic. We now give an example of a field with finite characteristic. III. Let p be any prime integer. We saw in § 2 that the ordinary integers, reduced modulo p9 form a commutative ring with unity. The zero 0 represents the set of integers which are multiples of py and the unity 1 the set of integers equal to 1 (modulo p). We now show that every non-zero element of this ring has an inverse. If a is any integer not a multiple of p, a and p are mutually prime. Hence we can find integers 6, c such that ab—pc = 1. If now a' — a (modulop), b' = 6 (modulop), it follows that a'b' = 1 (modulop). The set of integers modulop therefore form a field. This field is clearly of characteristic p. This example also illustrates the case of a field with a finite number of elements. IV. Let I be any integral domain with elements a,6,.... The quotient field K consists of elements ajb (b #= 0). The polynomial ring I[x], consisting of polynomials aQ + axx+ ...+anxn, is an integral domain. There is also an integral domain K[x], consisting of polynomials £+£*+•...+£*» (6....6, + 0). We show that the quotient fields of I[x] and K[x] are isomorphic.
38 I. RINGS AND FIELDS To each element g b(> + b1x+...+bmxm of the quotient field of I[x] there corresponds the unique element of the quotient field of K[x] similarly represented. Iff'jg' is another representation of this element in the quotient field of I[x]y so that fgf = f'g, the elements//</ and f'jg' are equal in both quotient fields. Conversely, any polynomial in K[x] j aa °h tin M b0 bl bn can be written . _ c0 + c1x+... +cnxn _/ T c c where ct = bQ... b^a^^... bn9 c = b0... bn. If, similarly, . _ d0 + dxx+ ...+dmxm __ g *~ d ~d' we have, in the quotient ring of K[x], 0 _ d(cQ + cxx+... + cnxn) f ~~ c{d0 + dxx+...+dmxmY so that <f>/r/r determines an element of the quotient field of I[x]. If, similarly, and , _ g[ V " d" and <j>yjf' = <f>'ifrf then c'dfgf = cd'f'g, and therefore the element of the quotient field of I[x] corresponding to <f>jrjr is the same as that corresponding to $'/&'. I'he correspondence between the two fields is therefore one-to-one. It can be immediately verified that it is an isomorphism. The quotient field of K[x] is denoted by K(x). It is called the field of rational functions, over K, of the indeterminate x. By the above argument, the field of rational functions of x2 over the field
9. EXAMPLES OF FIELDS 39 K(xx) is isomorphic with (and can be identified with) the quotient field of K[xl9 x2]. It is denoted by K(xv x2). Proceeding, we see that K(xl9...9xr_1){xr) = K(xly...9xr) is isomorphic with the quotient field of K[xv ...,xr]. It is called the field of rational functions of xv ..., xr over K. It is evident that the characteristic of this field is equal to that of K. V. Finally, we give an example of a non-commutative field. Let K be the field of the real numbers. Defining matrices as in § 2, Example V, consider the set Q of matrices which are of the form oc 8 -/? -y\ -8 a y -0 f$ -y a -8 y /? 8 a where a, fiy yy 8 are in K. Clearly, the sum of two matrices of Q is a matrix of Q, and if a, b are matrices of Q, the equation a + x = b has the solution x = b — a in Q. Thus the set Q forms a group under the addition law. Next, if a= /a 8 —8 — y\ a* = the product F* = aa* = / t> s P —r q where p = aa* - /?/?* — 77* — 88*, q = /?a* + a/?*- *7* + 7(J*, r = ya* + fy?* + a7* -/?£*, 5= (Ja*-7/?* + /?7* + aeJ*, and therefore aa* is in Q. Since multiplication of matrices is associative, and distributive over addition, the set Q forms a subring of the ring of matrices. The
40 I. RINGS AND FIELDS zero of Q is the matrix with a = /? = y = eJ = 0, and the unity of Q is the matrix with a = 1, /? = y = 8 = 0. If, now, a#0, and a* is given by a*= a(a2+ ^ + 72 + ^)^, ^ = -/^2 + ^2 + ,),2 + ^)-^ 7* = _r (a2 + /?2 + y2+ ^2)-1^ ^ = -^2 + /^2 + ^ + ^2)-1^ we find that aa* = 1 = a*a. Therefore every non-zero element of the set Q is regular: Q is a field. Let us write 1, i,j, k for the matrices corresponding to <x=l, /? = y=(J=0; /?=1, y=£ = a = 0; y=l, (J = a = /? = 0; 8=1, a =/? = y = 0, respectively. Then the element a of Q written above is expressible in the form a = a + fii + yj + 8k, and the product (a + j3i + yj + 8k)(a* + 0*i + y*j + 8*k) can be evaluated by multiplying out and using the relations i2=j2 = p = -1> jk = -kj = i, ki = — ik = j, y = - j* = &. In particular, since ij and ji are unequal, the field Q is not commutative. This field is usually called the field of quaternions.
* CHAPTER n LINEAR ALGEBRA, MATRICES, DETERMINANTS I n this chapter we investigate properties of a set of elements which can be combined according to certain rules to form elements of the same set. These rules involve the use of the elements of a field. In the applications which have to be made in later chapters we shall usually assume that the field used is commutative, and without characteristic. But the main results of this chapter are independent of any such assumption, and, unless an explicit statement is made to the contrary, our results will be true for any field whatsoever. We begin our investigation by selecting a field, which we denote throughout by K. We shall often refer to K as the ground field. 1. Linear dependence. Having selected a ground field K, let L be a set of elements (not necessarily belonging to K) which have the following properties: (i) there is a law of composition in L, denoted by +, under which the elements form an additive group; the zero of the group is denoted by 0;, (ii) corresponding to each element of K there is an operation which transforms any element of L into an element of L. If a is the element of K9 u the element of L, the result of operating on u with a is an element of L which is written au; (iii) the operations described above obey the following laws: (a) \u = u, (b) a(u +v) = au + av, (c) (a + b)u = au + bu, (d) a(bu) = (ab) u. A set Shaving these properties is called a left-hand linear set over K. From the definition, certain properties of linear sets can be immediately deduced. First, u + Ou = lu + Ou = (1 + 0)m = lu = tc, for all u in L. Hence, for all u in L, 0w = 0.
42 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS Corollary. m+(-1)m= 1w+(-1)w = Om = 0. Hence (— 1) u is the negative of u as defined in the additive group. Secondly, if a is a non-zero element of K, u + aO = aa_1ii + a0 = a(a~1M + 0) = a(a-hi) Hence aO = 0. This result is also true if a = 0, by the first result. The converse of these results is immediate. If au = 0, and a 4= 0, we have the equations u = lw = (0-½) m = cr^at*) = a-H) = 0. Hence, the equation au _ q implies either a = 0 or ti = 0. Corollary. If aw = 6w and w + 0, then a = 6; and if au = at? and a 4= 0, then u = v. A right-hand linear set can be defined as above, except that all operations are performed on the right. We can deduce similar results to those proved above for left-hand Unear sets. In the case of a commutative field K we shall not distinguish between right- hand and left-hand Unear sets, and we shall write au = ua, and talk simply of linear sets. But in the foUowing work we shaU not assume that K is commutative. However, it will be sufficient to confine our attention to left-hand Unear sets, since from each property of these sets a corresponding property of right-hand Unear sets will be immediately deducible. Whenever there is no risk of confusion we shall simply call the left-hand Unear set L a linear set.
1. LINEAR DEPENDENCE 43 Any r members uv ..., ur of L are said to be linearly dependent over K if there exist elements av ..., ar of K, not all zero, such that a^ + .^+a^ = 0; otherwise they are linearly independent. We note that no finite set of elements of L which includes 0 can be linearly independent, since, if the other members of the set be uv ...,ur, we always have the equation a0 + 0ut + ... + 0ur = 0 (a + 0). If uv ..., ur are linearly independent, then any subset u^,..., Ujs consists of linearly independent elements. For the linear dependence of the subset would involve the linear dependence of the whole set. Any element v of L which can be expressed in the form v = a{ut + ...+arur (at in K) is said to be linearly dependent on uv ..., ur. We shall denote the set of elements of L which are linearly dependent on ut,..., ur by the symbol L(uv ...,ur). The zero 0 is always in the set. We now prove Theorem I. L(uv ..., ur) is a linear set. If u, v are in the set, u = axux +...+ arurf and v == bxut + ...+ brur. Then u+v = (^ + 6^^ + ... + (^ + 6,.)11,.. Hence u+v is in the set. Also au = aaxux +... +aarury and therefore au lies in the set. Since the negative of u in the additive group is equal to (— 1) u, we see that the set L(ut,..., ur) is an additive group and is closed under multiplication by elements of K. It is therefore a linear set. Theorem II. Ifuv...,ur are linearly independent, the expression v = axut + ...+arur for any element of L{ut,..., ur) is unique. Conversely, if the expression is unique, uv...,ur are linearly independent.
44 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS If any element v of L(uv ..., ur) can be written in two ways as v = axut + ...+arur, and v = bxut + ... + brur, we see that 0 = (¾ - bx) ut +... + (ar - br) ur. Since uv ...,wr are linearly independent, it follows that «i = ^ (»= 1,...,^). Conversely, suppose that the representation is unique, but that axux + ...+arur = 0. Then v = b±ut + ...+ brur can also be written as v = (^+^)^ + ... + (6,. + ^)11^. From fy + a, = 6^ (i = 1,...,r), we deduce that a* = 0 (i = 1,...,r). This proves that tf|,...,ur are Unearly independent. Theorem III. Ifvv...9va lie in L(uv...,ur)9 and wv...,wt lie in L(vv ...,vs), ihenwv ...,wt lie in L(ut, ...,nr). For, from the equations v{ = o^M, + ... +airur (i = 1,. and W{ = bnvt + ... + biavs (i = 1,. we deduce that W| = caMi + -- + cirMr (*=!»• Whefe , V- A „ ..,*), ..,0. ..,<). Theorem IV. /w any finite set uv...,ur, not all 0, we can find a subset tc^, ...,1¾ , which is linearly independent, and such that L{uv ...,ur) = £(wtV ...,Mtg. We prove this by induction on r. The theorem is evident if r = 1, since, by hypothesis, ut =# 0, and is therefore linearly independent. Suppose now that the theorem is true for r — 1 elements, and consider the set uv ...,ur. If these elements are linearly independent
1. LINEAR DEPENDENCE 45 we choose them as the subset, and there is nothing to prove. Assume then that they are dependent, and that a1ut+. ..+arur = 0, where at least one ai 4= 0. If ak =f= 0, we see that uk = - ak ^i^i -...- afcW-i^-i - aklak+iuk+i -...- a>k 1a>rur. Not all the elements uv ..., uk_v uk+v ...,ur can be 0, otherwise Uk = 0 also. We can therefore apply the induction hypothesis to L(uv ...,ti£_j, Uk+V ...,wr), and we deduce that there exists a set of linearly independent elements u^9..., u^ such that L(uv...,uk_v uk+v...yur) = L(u{l>...9uis). Clearly L(U£V ...,u^ is contained in L(uv ...,ur). Conversely, any element v of L(uv ..., ur) can be written in the form v = 6^ + ...+6^+...+6,.11,. = (h ~ M*1^) ui + ••• + (bk -bkVk1® k) ^+... + {br-bka^ar)ur, and is therefore in L(uv ...,uk_vuk+v ...,ur) = L(uii9...,uis). Hence L(uv ..., ur) = L(u^f..., w^), and the theorem is proved. The elements vv ..., vs, whether linearly independent or not, are said to form a basis for the linear set L(uv ...,ur) if L(vv...,va) = L{uv...,ur). If the elements are linearly independent the basis is described as a minimal basis. We shall prove that the number of elements in a minimal basis for L(uv ...,ur) is the same for all minimal bases, and this number will be called the dimension of L(uv ...,ur). The principal step in proving that the number of elements in a minimal basis is always the same is the result known as theExchange Theorem. This has many applications in algebra. Theorem V. If vu ...,vs are s linearly independent elements of L(uv ...,ur)> there is a subset Ujt,...,Ujs of the basis uu ...,ur, such that if we exchange this subset for vt,...,vs the new set of r elements is a basis for L(uv ..., ur).
46 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS We prove this theorem by induction on 8. Assume that the theorem is true for vv ...,vs_v and suppose that these elements have been exchanged with ^,...,11^^, where ily ..., v^,...,ir is a derangement of 1,...,r. Then, by hypothesis, L(uu ...,ur) = L{vt9 ...,tV-n uia> .--.^), and contains v8. Hence vs = a1vt +...+ eVxtVi + a8uis +...+ aruir. (1) Not all the elements as,...,ar of K can be zero, since this equation would then imply the linear dependence of vv ...,vs. We may, without loss of generality, assume that the derangement iv ...,ir has been chosen so that a8 ^ 0. We can then write uia = -07^1^1-----^^-1^-1 + ^1^ -^^+1^.+1- - -a7larUir. (2) Now, ^1,...,1^,1^,11^,...,1¾. lie in L{vt,..., vs_u uis,... ,1¾). and therefore, by Th. Ill, L(vly...9vs_vvs,uia+iy...9uir)^L(vly...yvs^u uis,...,uir). But conversely, equation (2) shows that 1¾ lies in L(vv ...,vs_uvsy uis+l>...,uir), and therefore L(vt9 ...9vs_l9 uia, ...,uir)^L{vly ...yvs_ly vs, uia+i>...,uir). Hence L(uv ...,ur) = L(vt9...,vs_t9 uisy...,uir) = L(vly...yvs_vvsy uis+l> ...yuir). The same reasoning can be applied to the case 5=1, when there has been no exchange. Hence the theorem is proved. Corollaries (1) Ifuv ...,ur is a minimal basis for L(uly ...,tir), and vv ...,vr are r linearly independent elements of this sety then L(uv...,ur) = L(vv...,vr). (2) Ifvv ...,t?r+| are any r+1 elements of L(uv ...,ur)y they are linearly dependent.
L LINEAR DEPENDENCE 47 If any r of the elements Vf are linearly dependent, the corollary needs no proof. Suppose, then, that vx,..., vr are linearly independent. Then vr+l is in L(ul9 ...,ur)9 which, by the Exchange Theorem, is the same as L(vv...9vr). Therefore vr+i =ai»i + ...+arVr> and therefore vv .. .,tv+i are Unearly dependent. Note that this proof does not assume that ul9...9ur form a minimal basis for L(uv ...,ur). Theorem VI. The number of dements in a minimal basis for L(ul9..., ur) is the same for all bases. Let vt,..., vn> and wt,..., wm be two minimal bases. Then L(uv ...9ur) = L{vv ...yvn) = L(wv ...,wm). If n>my Corollary (2) states thati?j, ...,t?n are linearly dependent, and similarly, if m > ny wv ..., wrn are linearly dependent. But, by definition, the elements of a minimal basis are linearly independent. Therefore m = n. As stated above, this number is called the dimension of the linear set. As an evident corollary to the above theorem we have (1) If Ly M are linear sets of finite dimensions Z, m respectively, and L^M, then l^m. If L^M and I = m, then L = M. We now give an important example of a left-hand linear set over a field K. We consider the set of n-tuples (av ...,an) consisting of n elements ai oiK. Two n-tuples (av...,an) and (blf...,6n) are equal if and only if «< = &< (i = l,...,tt). * If u = (al9...,an)9 and v = {bl9 ...,bn) are two n-tuples, we define u + v by the equation u + v= (a1 + bv...9an + bn)f and au (a in K) by the equation au = (aal9..., aan).
48 II. LINEAR ALGEBJtA, MATRICES, DETERMINANTS With these definitions it is evident that n-tuples form a left-hand linear set. If ^ = (1,0,0,...,0), ^2 = (0,1,0,...,0), • • • ^=(0,0,0,...,1), then uv...,un are linearly independent, and any n-tuple u = (al9..., an) is given by the equation u = axux+...+anun. Hence the set of n-tuples forms a linear set of finite dimension n, for which uv ..., un are a minimal basis. The elements of this linear set are usually called left-hand vectors, or left-hand n-vectors when we wish to emphasise the dimension of the set. We say that the set of vectors forms a left-hand vector manifold. In the same way we can define right-hand vectors, and, when the ground-field K is commutative, simply vectors. Theorem VII. Any left-hand linear set L(uv...yun) of finite dimension n is isomorphic with the set of left-hand n-vectors. Since the linear set is of dimension ny uly...yun are linearly independent. Hence [Th. II] any element u of the set can be written in a unique way as tc = a1ul+...-\-anun. (3) There corresponds to u, therefore, the unique vector (aly...9an). Conversely, the w-vector above determines the unique element of the linear set given by (3). The proof that this one-to-one correspondence is an isomorphism is trivial. It will be noted that the correspondence between the linear set and the set of n-vectors depends on the choice of a minimal basis in the linear set. In the remainder of this chapter, and in the applications to be made later, we shall have to consider simultaneously the properties of a finite number of subsets, each of finite dimension, belonging to some linear set L. Now, if we choose a minimal basis for each of these subsets, we have an aggregate which consists of a finite number of elements of L. These elements define a subset U of L, of finite
1. LINEAR DEPENDENCE 49 dimension, which contains each of the given subsets. All our operations will take place in L'. We may always assume, therefore, that we are dealing with a linear set U of finite dimension. By Th. VII, U is isomorphic with a vector manifold. For this reason we shall usually speak of the elements of U as vectors, and in this sense the discussion in the following paragraphs is related to vectors. We shall not state the dimension of the manifold of vectors explicitly, since our results will not depend on it; but on occasion it may be necessary to assume that the dimension is 'sufficiently high'. 2. Matrices. A matrix of p rows and q columns over a field K (called, for brevity, &pxq matrix) is an array A = /au a12 . alq\ = (ar8) I «21 a22 • a2q 1 \ ' ' I \Ct>pl dp2 - apq/ ofpq elements of K, arranged in rectangular form. In § 2 of Chapter I we met, by way of example, the special case in which K is the field of complex numbers, and p = q. The theory of matrices is closely related to the theory of vectors, and we shall use the results of the preceding section to develop properties of matrices. The connection with vectors is established in this way. We consider in some vector manifold of finite but sufficiently high dimension, which need not be specified, a set of q left-hand vectors ul9...,uq. By means of the matrix A we can determine from these q vectors p vectors vl9...9vp where ^ = %i% + ...+Va (*' = 1,...,!>). (1) Then L(vl9 ...9vp)^L(ul9 ...9uQ). (2) The individual vectors vi are uniquely determined from the vectors ut by means of the matrix A. The set vv ..., vp is called the transform of the set ul9..., uQ by A. We notice, on the other hand, that if we are given condition (2) the matrix A is uniquely determined if and only if uv ...,uq are linearly independent. This is an immediate consequence of § 1, Th. II. Sets of equations such as (1) will occur very often in this book, and it is convenient to have an abbreviated notation for them. If we generalise the notion of a matrix so that the elements may be
50 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS vectors instead of elements of K, and let U, V be respectively the qx 1 andpxl matrices we write (1) in the form V = AU. (3) We shall see later that in writing A U for the p x 1 matrix ^11^1+... + ¾^ V«pi% +...+apquq/ we are using a special case of the law of multiplication of matrices. Finally, in writing ajxl matrix, such as ?7, in full, it is sometimes typographically convenient to write U = [uv ...,uq], the square brackets indicating that the elements enclosed should be read as a matrix of one column. Now let B be any otherpxq matrix over K, and use it to define p vectors wl9 ...,wp, where [101,...,1¾] = BU. Then Vi + w^ (aa + ba)u1 + ... + (a{q + b{q)ua (i = 1,...,^). These equations can be abbreviated to the equation V+W = (A + B)U, if the sum of two pxq matrices is defined by the equation ( an «12 • «iA + Ai &ia • blQy «21 «22 • a2q I I ^21 ^22 • ^2q «pl «j>2 • apq/ \>pl bp2 . bpq/ = /aii + fcn «12 + ^12 • «1<? + 61(?> «21 + ^21 «22 + ^22 • a2q^~^2q • • • • *pl + &pl «i>2 + 6p2 • «Pfl + 'W
2. MATRICES 61 With this definition of addition for p x q matrices, we see that p x q matrices over K form an additive group. The zero of the group which, without risk of confusion, may be denoted by 0, is the zero p x q matrix, that is, the p x q matrix whose elements are all zero. Now let A be a p x q matrix over K, and let B be a q x r matrix over K. Let ul9 ...,ur be r vectors and vv...,vq the vectors defined by the equation v — FiTT where now U = [ul9 ...yur] and V = [vl9 ...,t?J. Similarly, consider the set of vectors W = [wl9...9wp] defined by the equation W = AV. Then L(wl9 ...,wp)cL(vl9 ...9vq)c L(ul9 ...9ur)9 and hence there must be a p x r matrix G such that W=CU. As in § 1, Th. Ill, we may take G = (c„), Q where ctj = ^aikbki (i = l,...,p;j= 1,..., r). (4) If we formally eliminate V from the equations W = AV9 V=BU by writing W = 4(£t7) = 4J3C7, this suggests that we define the product AB of A and B to be the p x r matrix (7, whose elements are given by means of (4). We see that the product of a p x q matrix and an r x s matrix is only defined when q = r9 and that the result is &p x s matrix. If AB and BA are both defined they are not necessarily equal. We note here that AB is said to be obtained from A by post- multiplication by B9 and that AB is obtained from B by pre- multiplication by A. Matrices do not form a ring, since the sum of two matrices is only defined when each is a p x q matrix, and the product is only defined when the number of columns in the first matrix is equal to the number of rows in the second matrix. If we restrict ourselves to
52 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS square nxn matrices we do obtain a ring, as we shall see in the next section; but certain associative, distributive and commutative properties also hold for all rectangular matrices. Let 4, Av 42 denote pxq matrices, let B, Bv B2 denote qxr matrices, and let C denote anrxs matrix. If BC = D, AD = E, r then d<,= :¾¾ (i= l,...,ff;j= 1,...,*), and therefore Q q r = 2 S aiu^utCtj r = S factj> where F = (/^) = 42?. Hence and the product can be written as ABC without ambiguity. This is the associative law for the multiplication of matrices. By direct calculation we may easily prove that 4(^ + 52) = 4^ + 4^, and (4x + 42) B = AXB + A2B. These are the distributive laws for matrices. If 4 is a pxq matrix, and il9...,ip,, jv ...,jq, are any integers satisfying the relations l^i1<i2...<ip><:p, we can obtain from 4 a new matrix of type p' x q' by striking out all the elements which do not lie in the i^th, i2 th, ..., or ip.th rows, and in the ^th, j2th, ..., or j^th columns. The matrix we obtain is called a submatrix of 4.
2. MATRICES 53 Now let pv ..., p8, ql9..., qt be s +1 positive integers such that 8 t 1 1 Let the submatrix of A whose elements are in the ( S Pk + 1W, ( S Pk +2W, ..., ( S ^W rows, and the ( 2 #fc + l)th, ( 20* + 2)th, ..., ( £ g^lth columns U-i 1 U-i / U-i / be denoted by 4^. The submatrix Ay is a ^ x ¢^ matrix, and in an obvious notation we can write A = /An . Alt\ 181 ' 4/ The matrix A may then be described as divided into blocks, corresponding to the numbers pv...,p8,ql9...,q(. If B is a q x r matrix, let r1? ...,rw be a set of positive integers u such that 2 ri = r. We may then divide B into blocks corresponding l to the numbers qv ...,qt\ rv ...,ru, and write 1? = /-Bn • -Bi 5a • Btu Now, ^4lA. is a Pi*qk matrix, and Bkj is a (foxr, matrix. Hence AikBki is defined, and is a pi x r, matrix. We can therefore define aPi^Tj matrix Gij9 given by the equation (7,,= 2 4¾. (5) Consider the element in the Ith row and mth column of Ctj. It is 2 ahnbnk> n-l where & = Z#a + ^ *= 2 ^a + w* Thus /Cn is simply the division of (7 = -45 into blocks corresponding to the
54 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS numbers pv ..., pa, rv ...,ru. We may therefore regard A and B as 8xt and txu matrices whose elements are matrices, and their product is defined by (5). This is a definition similar to that given in (4). But it must be remembered that (5) is only possible when the submatrices of A and of B are chosen in the way we have indicated. Similar, but simpler, reasoning shows that if A and B are two p x q matrices divided into blocks corresponding to the same set of numbers, thus A=/An . Au\, B=/Bn . Blt\ \A8l . AJ \Bs1 . BJ then A + B = /An + Bn . Alt + Blt\ W + -B.1 • A*+BJ The device of considering matrices divided into blocks is often used. The operations discussed above are the only ones which will be employed, and the matrices concerned will always be matrices with elements in a field. 3, Square matrices. A matrix in which the number of rows is equal to the number of columns is said to be a square matrix. If we consider only nxn matrices, we see that two such matrices can always be added or multiplied, and the result is an n x n matrix. The properties demonstrated in §2 enable us to say that nxn matrices over Kform a ring. If we refer to I, § 2, Example V we see that the ring is non-commutative, and has divisors of zero. But it has unity, namely, the unit matrix . o ^0 0 . 1/ where <*« = 0 (**.?")> and <*«=! (•-1,..., *)• The symbol Sy defined in this way is frequently used in algebra. It is called the Kronecker delta.
3. SQUARE MATRICES 55 In accordance with the terminology of I, § 3, an n x n matrix A is said to be regular if it has an inverse A~l such that AA-1 = A~*A = In. From the theory of rings developed in the previous chapter, we know that a matrix cannot have two distinct inverses. We also see that if both A and B are regular, having inverses A*1, B"1 respectively, then AB is regular, with inverse B^A"1. For {AB) {B^A-1) = ABB-tA-1 = AInA~* = AA'1 = /n, by the associative law. Similarly, (B-iA-i)(AB) = In. We now obtain two theorems giving necessary and sufficient conditions for anwxw matrix A to be regular. Theorem I. An nxn matrix is regular if and only if it transforms a set of n linearly independent left-hand vectors into a set of n linearly independent vectors. (1) Let uv ..., un be a set of n linearly independent vectors. Suppose that the n vectors vi = ailu1 + ...+ainun (i= l,...,n) are linearly independent. By § 1, Th. V, Cor. (1) L(vv...,vn) = L(uv...,un). Hence ^ = 6,^ + ...+6<nt;n (i = l,...,n), where by is in K. Then n n Vi = S ^,6,^ +... + 2 aqbjnvn9 n n and ^ = 2 &#%^i f...+ 2 &«a* A- Since both the sets u1?..., wn and ^,..., t;n are sets of linearly independent vectors, each of the two equations above must be an identity. Therefore n n 2 a>ijbjk = #iA;, and 2 &#%* = #**• In matrix notation this is simply AB = /n = £A Therefore ^4 = (a{j) has an inverse 5, and is regular.
56 n. LINEAR ALGEBRA, MATRICES, DETERMINANTS (2) Now suppose that A is regular, and B is its inverse. Let uv...,unbesb set of independent vectors, and let vi^ailu1 + ...-^ainun (i= 1,...,»); then L(vl9...9vn)sL(uv...9un). Now 6^+...+ binvn n n = S b{ianul +... + S 6«aM«H Hence and therefore -*<i«i + ... + *<»t«n -««. L(uv...,un)^L(vv...,vn), L(u1}...,un) = L(vu...,vn). Since uv...,un are linearly independent, it follows from § 1, Th. VI, Cor. 1, that vv ...,vn are linearly independent. This completes the proof of the theorem. If G is any pxq matrix n . x we may regard the p rows of C (civ ...,c^a) as £) left-hand g-vectors. We speak of these as the q-vectors given by the rows of C, and we shall always regard them as left-hand vectors. Our second criterion for a regular matrix is given by Theorem II. An nxn matrix A is regular if and only if the n-veetors given by the rows are linearly independent. As in the previous theorem we denote n linearly independent vectors by uv ..., un, and the transforms of these by A by vv ..., vn. (1) Suppose that the n-vectors (aiv...,ain) are linearly independent. Then if n i —1 we must have bi = 0 (i = 1,...,n). Now if vl9..., vn satisfy a relation &!*>!+... +bnvn = 0 (bi in K)y this implies that n n E 6^1^1+ -. + S biainun = °- i-1 i-1
3. SQUARE MATRICES 67 Since uv ..., un are linearly independent, n 26<atf = 0 0'= It ...tn)f <-i and therefore, by the above, bi = 0 (i = l,...,n). Therefore t^,..., tfn are hnearly independent, and, by Th. I, A is regular. (2) Suppose that A is regular. Then vv ..., vn are linearly independent. Now, if the n- vectors (aiv ..., ain) are linearly dependent, there exist elements bv..., bn of K, not all zero, such that n 2^a =0 (j= 1,..., rc). Then 6i^i+... + 6nvn = 2 Mii^i + ... + 2 biainun = 0. This implies the linear dependence of t^,..., vn. From this contradiction we deduce that the n-vectors given by the rows of A are linearly independent. The methods used above also give us Theorem III. If A is a pxn matrix, and the left-hand n-vectors determined by the rows of A determine a vector manifold of dimension r, the manifold L(vv ..., vp), where vi = aixux+...+ainiin is also of dimension r, provided that uv ...,un are linearly independent. If the vectors defined by the rows of A be then ai9...,aig, the vectors of any subset of the vectors av ...,ap are hnearly dependent or independent according as the corresponding vectors vix,..., vig are linearly dependent or independent, and conversely. For iif...,ii Uf..,U i-l if and only if _ , ^ / • , \ since wx,..., wn are linearly independent.
58 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS It follows that if the maximum number of independent vectors in the set av...,ap is r, the maximum number of independent vectors in the set vv ..., vp is also r. This proves the theorem. Theorem IV. If A is a pxn matrix, and the p left-hand vectors determined by the rows of A are linearly independent, then p^n, and there exists an (n—p)xn matrix B such that thenxn matrix @ is regular. We have seen that a minimal basis for the manifold of n-vectors is given by the n vectors (ta>->*<n) (»-l,...,n). It follows that the maximum number of hnearly independent n-vectors is n. Therefore p^n. Now let uv...,un be n linearly independent vectors and let Then, as in the above theorem, vv...,vp are hnearly independent. Applying the Exchange Theorem [§ 1, Th. V], we deduce that there is a derangement il9..., in of the numbers 1,..., n such that if then vv...,vn form a basis for L(ulf..., un). This basis is necessarily minimal. Hence, by Th. I, *>i = cilul+ -~+Cinun (i = !f —»*)» where C = (ctj) is regular. But where B is the (n—p)xn matrix whose ^th row is This establishes the theorem. 4. Transformations of a matrix. Now let A be a p x q matrix. If uv ...,uq are q linearly independent vectors, A transforms them into p vectors, vv ..., vp9 say, and L(vv ...,vp)^L(uly ...yuq). If we are concerned with the vector manifolds L(vv...fvp) and
4. TRANSFORMATIONS OF A MATRIX 59 L(uv ..., uq), we are at liberty to change the bases in these manifolds without altering the manifolds themselves. Since uv...,uq are assumed to be linearly independent, a new set of vectors uf,..., u* forms a minimal basis for L(ul9..., uq) if and only if U=QU*, (1) where, in the notation of § 1, U = [uv ...,uq], ?7* = [uf,..., w*], and Q is a regular g x # matrix [§ 3, Th. I]. Ifvv ...,vp are linearly independent, a similar argument applies to L(vv ...,vp). The set v$,..., v* is then a minimal basis for L^i,..., vp) if and only if ym = p^ where P is a regular pxp matrix. This result requires modification if vv ..., vp are not linearly independent. But in the case we are considering P is a regular pxp matrix, and if F* = PF, then V = P^F*, and therefore L(vf,...,v*)g:L(vv...,vp)^L(vf,...,v*), so that Zr(t?f, ...,**) = £K, ...,¾). Hence, in order that the equation (2) should represent a change of basis in L{vl9..., vp), it is a sufficient condition that P should be regular. Simple examples show that it is not a necessary condition when vv ..., vp are not linearly independent. Equations such as (1) or (2), where Q and P are regular matrices, are said to define allowable transformations of the bases uv ...,uq and vv...,vp of L(uv...,uq) and L(vv...,vp) respectively. These allowable transformations replace the matrix equation V = AU (which connects vl9...,vp with uv ..., uq) by V* = A*U*> where A* = PAQ. We note that (i) A = IpAIq, (ii) if 4* = PAQ (P and £ regular), then A = P-M*^1; (hi) if 4* = P-4Q and A** = PM*#*, then 4** = (P*PM(#Q*).
60 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS Thus the relation between A and A* is reflexive, symmetric and transitive. It is therefore an equivalence relation [I, § 2]. Our object in this section is to determine a set of canonical forms for p x q matrices. In other words, we shall obtain a set of p x q matrices with the properties (1) No two matrices of the set are equivalent, in the above sense. (2) There is a matrix of the set equivalent to any pxq matrix. We first establish the existence of a numerical character which is invariant under our equivalence relation. Let rA denote the maximum number of q-vectors given by the rows of A which are linearly independent. If these j-vectors are all zero, rA = 0. By § 3, Th. Ill, we see that rA is also the maximum number of linearly independent vectors of V. That is, the dimension of L(vv ..., vp) is rA. But since L(vv...,vp) = L(v*,...,vl), it follows that rA is also the number of linearly independent vectors in the set v*,..., v*. Therefore rA =, rA*. We notice that the equations V* = PAU, V* = PAQU* determine the same vectors of L{uv...,uQ) = !«...,<). Therefore the number of independent rows of PA is equal to the number of independent rows of A * = PAQ, and both these are equal to the number of independent vectors in the set v*,..., v*, which is rA. Hence, calling the number rA the row-rank of A, we have Theorem I. The row-rank of a matrix A is equal to the row-rank of PAQ, where P and Q are regular matrices. This includes the cases in which Q = Iq, P = Ip, when we have, respectively, the matrices PA, and AQ. Pre-multiplication or post- multiplication of A by a regular matrix does not alter its row-rank. Now, the replacement of the vectors vv...yvp by the same vectors vixy..., v* in a different order is a transformation Vx = W where Px is a regular matrix. For, since every permutation of the numbers !,...,# has an inverse permutation, Px must have an inverse.
4. TRANSFORMATIONS OF A MATRIX 61 We choose Px so that the first rA permuted vectors are linearly independent. Let us denote the permuted vectors by VU 0'= !> •••>?). Then since the vectors vi,i (.7 = ^+1,...,^) are linearly dependent on the first rA, we have the equations Let Vitf = It *>jkvltk Zc-l ^= /KA+11 \*,x U m • • = ^+1,. bU+lrA hrA J ..,p) | 1' and let P2 = I Ir Since IPJ[-B IpJ the matrix P2 is regular. If we now put H = AK> calling the transformed vectors v2l9 ...9v2p9 then of these, the Hi (* = 1,...,^) are linearly independent, but the rest are zero; that is, v2j = ° U = rA+ly...,p). Since V2 = P2K = P2PX V = P^^tf = ^2^, say, and uv...yuq are linearly independent, the last p — rA rows of A2 must be zero. We therefore write 0' where C is a r^ x q matrix.
62 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS By the Exchange Theorem we can replace rA of the vectors uly...yuq by the linearly independent vectors v21, ...9v2rA, and thus obtain a new basis for L(ul9 ...,uq). Before doing this we make a transformation tt — n tt which merely permutes the vectors ul9 ...yuq, so that those to be replaced are the first rA vectors. The exchange is then represented by another transformation where Q2 is regular. Hence, finally, we have V2 = A2U = A2Q1QiU2 ICQiQ*\ that is, V2 = (D\ U2> "■©■ say, where D = (d{j) = CQ1Q2. Now consider the first rA of the p equations given by this last matrix equation. Since Q we have u2i = ^d^u^ (i = 1, ...,rA). Since u21> ...yu2q are linearly independent, we must have dij = 8a (i = 1,...,^)- Hence /D\ = /IrA 0\ U/ \0 0/' That is, P^AQ^^ /IrA 0\ Since Pv P2, Ql9 Q2 are regular matrices, P = P2P± and Q = QXQ2 are regular. We have therefore proved Theorem II. IfAisapxq matrix of row-rank rA> A is equivalent to the matrix PAQ = (IrA 0\ U oj-
4. TRANSFORMATIONS OF A MATRIX 63 It is sometimes convenient to effect the transformation (or reduction) of A to this canonical form by a series of elementary steps, or elementary transformations. We define these as the following transformations of the basis wly..., wr (not assumed to be minimal) of a vector manifold L(wv ..., wr): (i) the interchange of two elements wi9 w^ (ii) the multiplication of an element wi of the basis by a non-zero element of K; (iii) the addition to any element of the basis of any vector linearly dependent on the remaining r — 1 elements of the basis. Each of these transformations of the basis is called an elementary transformation, and is represented by the equation W* = RW. The matrix R is regular, for the inverse of an elementary transformation is easily constructed, and is itself an elementary transformation. It follows that the result of any number of elementary changes of basis replaces the basis by an equivalent basis, and is reversible. We now consider the equation V = AU, Q that is, vi = yElaijuj (i = 1,...,p). j —l Unless rA = 0, there is at least one element of A, say aijf which is not zero. By performing an elementary change of basis of type (i) on the bases uv ..., uq and vl9..., vp, we obtain new bases, which we denote by , ultl,...,ultQ and v1>v ...,vltP, and a relation Vx = A^ (A± = (ahij))9 where ain = <%4=0. By a transformation of type (ii) applied to uhl we may arrange that altll = 1. We then perform the transformation of type (iii) on ulv ..., ulq which is given by the equations Q i-2 so that we now have the matrix equation Vi — A^U^y
64 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS where A2 = IIX 0 \ [c b}> Bind C is a (p — 1) x 1 matrix, <7=(cw) (1 = 2,...,^=1). Now perform the transformation on vltl,..., vliP. This is a series of transformations of type (iii). We obtain the equation V2 = A3U2, where Az = /Jx 0^ (o' i)- Since the number of independent rows of A3 is still rAi it follows at once that B is of row-rank rA — 1. We now show by induction that the matrix A can be reduced to canonical form by elementary transformations on the bases of L(vv ..., vp) and L(ul9..., uq). There is nothing to prove when rA = 0, since A must then be a zero matrix. Let us suppose that the theorem is true for any matrix whose row-rank is less than rA. By what has been proved above we can reduce A, by elementary transformations, to the form .j Q. (o B)> where B is of rank rA—\. By the induction hypothesis we can reduce B to canonical form by elementary transformations applied to v2 a> • • • > v2,p an(l u2,2> • • • > u2t <r But these transformations can also be regarded as elementary transformations applied to v2Vv229..., v2p and w2,i> u2t 2,..., u2q. This proves the theorem. As a corollary we have Theorem III. Any allowable transformation of the basis of a vector manifold L(uv ...,uq) of dimension q is the result of a finite succession of elementary transformations. For we can apply the result proved above to the set of vectors t*f,...,u* and to the set uv...,uq, and reduce the matrix P to canonical form. Suppose that a succession of elementary transformations, corresponding to matrices Qv ..., Qa, on the vectors
4. TRANSFORMATIONS OF A MATRIX 65 uf,...,u*, together with a set of elementary transformations, corresponding to matrices Rv...,Rb, on the vectors ul9...9uq9 transforms the first set of vectors into vf,..., v*9 and the second set into vl9..., vq. Then if the matrix expressing the relation between the vf and the Vj is in canonical form, we have the equations v* = v{ (i= l,...,rP), But since P is regular, rP = q. Therefore the transform of U* = PU is V* = IQV. But V* = QaQa_1...Q1U*9 and 7 = ¾¾^...¾^. Hence QaQa^... ^P^1... ^1 = Iq> that is, P = Qr1... Qa1 Rb ... Rv Since (¾ is the matrix corresponding to an elementary transformation, so is Qj1. This proves the theorem. 5. The rank of a matrix. Until now, our theory of matrices has been based on the consideration of left-hand vectors, and in particular we have been led to consider the rows of a p x q matrix as left-hand ^-vectors. We now consider how our results would be affected if we based our investigation on right-hand vectors. The theory of right-hand vectors is immediately deducible from the results of § 1. If xl9..., xp are p such vectors, they are said to be linearly independent over K if the equation x1a1 + ...+xpap = 0 (fyin K) implies the equations ^ = 0 (i = 1,...,p). Also, if y is said to be linearly dependent on xl9..., xp. We then prove the analogues of § 1, Th. I-VII, noting in Th. VII that a set of right-hand vectors of dimension p is isomorphic with the set of right-hand ^-tuples in which multiplication on the right is the only kind admitted.
66 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS A p x q matrix A transforms the p right-hand vectors xv ..., xp into theq vectorsyl9...,yqy where p Proceeding from this equation we can deduce results similar to those obtained in §§ 2-4. The most important one from our point of view is that if x-±,..., Xp are linearly independent, the number of linearly independent vectors in the set yly...,yq is the number of linearly independent columns of A, regarded as right-hand vectors. We call this the column-rank of A. Just as in §4 we show that if A is of column rank sA, there exist regular matrices P and Q such that But we saw in § 4 that the row-rank of A is equal to the row-rank of PAQ, which in this case is clearly sA. Hence we have Theorem I. In any pxq matrix A the number of linearly independent rows, regarded as left-hand vectors, is equal to the number of linearly independent columns, regarded as right-hand vectors. This number we now call the rank of A. We recall that by §3, Th. II, an n x n matrix is regular if and only if its rank is n. The rows of A may be considered as right-hand vectors, and the columns as left-hand vectors if we interchange rows and columns. We define the transpose A' of A as the qxp matrix where a^ = ajt {i = 1,...,g; j= 1,...,#). Clearly, the number of linearly independent rows of A1, regarded as left-hand vectors, is equal to the number of linearly independent columns of A, regarded as left-hand vectors. The number of linearly independent columns of A', regarded as right-hand vectors, is equal to the number of linearly independent rows of A, regarded as right-hand vectors. By the theorem above these numbers are equal, and give the rank of A1. The consideration of the case when K is commutative, and left-hand and right-hand vectors are identical, gives us Theorem II. If the ground field K is commutative, the rank of a matrix A is equal to the rank of its transpose A'.
5. THE RANK OF A MATRIX 67 On the other hand, if K is not commutative, this theorem is not necessarily true. For K contains at least two elements a, b such that ab + ba. Neither a nor b is zero. Consider the 2x2 matrix A = /l a\ \b ab)' If P(l,a) + q(b9ab) = 0, then p + qb = Qf and pa + qab = 0. Hence — qba + qab = 0, that is, q(ab - 6a) = 0. Since K is a field, and ab # ba, we deduce that ¢ = 0. Therefore p = — qb = 0. Hence the number of independent rows of A is 2, and its rank is 2. On the other hand, the equation {l,a)p + (byab)q = 0 is satisfied by p = 6, q = — 1, and so the rank of A' is 1. Our final theorem in this section is often useful: Theorem III. If A is a pxq matrix of rank r, A contains a sub- matrix of type rxr which is of rank r. No submatrix of A is of rank greater than r. If A is of rank r it has r linearly independent rows. Every other row of A is linearly dependent on these. If the submatrix defined by the r independent rows be denoted by J5, B is an r x q matrix of rank r. Recalling the proof of § 3, Th. IV, we see that there exists &> (q-r)xq matrix C such that the matrix is regular, where the hth row of G is (h> ---jiq) being a derangement of (1,...,q). Since D is regular, its rank is ¢, and hence its columns are linearly independent, regarded as right-hand vectors. In particular, the ixth,...,irth columns are
68 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS linearly independent. These columns contain only zeros in the last (q — r) rows of D. It follows that the columns of the rxr submatrix E, which consists of the ^th,..., irth columns of B, are independent. That is, E is of rank r. Hence the first part of the theorem is proved. If J?7 is a submatrix of rank s > r, it contains s independent rows. Let these lie in the jxth, ..., ^th rows of A. Since 8>r, there exist elements bi of K> not all zero, such that 2¾¾. = 0 (i = jv ...,j8; j = 1, ...,g). But since this summation includes values of j corresponding to the columns of F, it follows that the rows of F assumed to be independent are linearly dependent. From this contradiction we deduce the second part of our theorem. 6. Homogeneous linear equations. Let i beayxg matrix over the field K. Consider the j-tuples (xv ...9xq) which satisfy the equa ions ailx1+...+aiqxq^ 0 (i = 1,...,p). (1) If y = (Vv •••></(?)> and z = (Zi, ...,¾) are two ^-tuples satisfying (1), then . , . , v yb + zc = (y1b + z1c,...,yqb + zQc) also satisfies (1), for any b and c in K. It follows that the g-tuples (xv ...9xQ) which satisfy (1) form a right-hand vector manifold. Any vector of this manifold is called a solution of the left-hand linear homogeneous equations (1), in which (xv...9 xq) is called the unknown. We write the equations (1) in matrix notation as Ax = 0, (2) where x = [xv ..., xJ. If A is of rank r there exist regular matrices P and Q such that PAQ=lIr 0\ \o or Let y — \yv ...,yq] be a solution of (2), and let z = Q~xy. Then 0 = PAy = PAQz = 1^,...,^,0,...,0]. Hence zf = 0 (* = 1, ...,r).
6. HOMOGENEOUS LINEAR EQUATIONS 69 On the other hand, let zr+1,..., zq be any elements in any extension of K, and let rA 1 z= [0,...,0, zr+1,...,zj, and let y = Qz. Then Ay = P^PAQz = P^(Ir 0\z \0 0) = 0. Hence we obtain any solution of (2) as follows: let z = [0,..., 0,zr+1,...,za], where zr+1,..., zQ are indeterminate over K. Any solution y of (2) is obtained by a suitable speciahsation of the indeterminates in Qz. A basis for the manifold of solutions is obtained by taking z = [0,..., 0, 8hr+1,..., Shq] (h = r+ 1,...,q). Hence the manifold of solutions has dimension q — r. Let us call this manifold, which consists of right-hand vectors, R, and let L denote the manifold of left-hand vectors defined by the rows of the matrix A. If a = (av...9aq) is any vector of L, we can write p If y is any vector of J?, axyx +...+aqyQ p p The two vectors a and y are said to be apolar. The theory of right-hand linear homogeneous equations £*A7 = ° 0' =1,...,3) (3) can be developed as above, and we can prove that if the matrix B = (b„)
70 n. LINEAR ALGEBRA, MATRICES, DETERMINANTS is of rank t, the solutions form a left-hand vector manifold of dimension q — t. In particular, consider the case when the columns of B are right-hand (/-vectors forming a basis for the manifold R of Solutions of (2). Then and the solutions of (3) form a left-hand vector manifold of dimension q — t = r. But any vector of L is then a solution of (3), and it therefore follows that L is the manifold of solutions of (3). Hence the manifold L and the manifold R are in a dual relationship, each defining the other uniquely. We conclude this section by using the foregoing results to prove Theorem I. If A is a pxq matrix of rank rA, B a qxr matrix of rank rB, then C = AB is apxr matrix of rank rc, where rA + rB-q^rc^mm(rA,rB). (i) Since B is of rank rB there exist regular matrices Q (of type q x q) and R (of type r x r) such that B = QIITB 0\R. (4) U 0/ Also, by §4, Th. I, C* = GBr\ A* = AQ (6) are of ranks rc, rA respectively. We write A* = {Af Af), where A* is a p x rB matrix, and A% a p x (q — rB) matrix. Now A* has rA linearly independent columns, regarded as right-hand vectors. It follows that Af has at least linearly independent columns, since any set of rA columns of A contains at least this number of columns of A%. Hence, the rank of the matrix M?0) is at least rA + rB-q'
6. HOMOGENEOUS LINEAR EQUATIONS 71 But (Af 0) = (Af A!)/Irs 0\ -t :°) ° = ABR~\ from (4), = CR~l = C*, from (5). Since C* is of rank ra, it follows that rA + rB-q^ra. (ii) The solutions of the equations Gx = 0, (6) form a right-hand vector manifold R of dimension r — rc. If y is any solution of the equations Bx = 0, (7) we have Gy = A By = 0. Hence, any solution of (7) is a solution of (6). But the solutions of (7) form a vector manifold S of dimension r — rB. Since it follows that r—rB < r — r0; that is ra ^ rB. If instead of considering equations (6) and (7) we considered the equations xC = 0, and xA = 0, we should obtain the relation Finally, then rc < min (rA, rB), and the proof is completed. 1. Matrices over a commutative field. The remainder of this chapter is devoted to the case in which the ground field K is commutative. Most of the results we shall obtain will be independent of the characteristic of the field, but some will be proved only in the case of a field without characteristic. Unless the contrary is stated,
72 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS K is assumed to be commutative, and no restriction is imposed on its characteristic. The rank of a matrix A is now equal to the rank of its transpose *A' [§5, Th. II]. Hence, in particular, if A is a regular nxn matrix, A' is also a regular nxn matrix. An nxn matrix of rank less than n is said to be singulary and the terms 'non-singular' and 'regular' are synonymous. We consider how to obtain the transpose of the product AB of a pxq matrix A and a,qxr matrix B. Let AB=C, where IfC'is the transpose aikbkj ofC, < (i= 1,. ..,p 1 = cii Q = li^jkhi fc-i Q «= S akjKk = S bikakj, k~\ since iT is commutative. Hence C" = B'A'. Now let .4 be a non-singular nxn matrix. A' is also non-singular, and therefore has an inverse B = {A')-1- We then have A'B = In = 54', and hence (4'-B)' = i; = In = (54')'. But (A')' = A, and therefore B'A = In = AB', that is ^ = 4-1. or B = (A~1)'. Hence, for any non-singular matrix A, (A')-1 = (4-1)'. We shall usually write this matrix as A, and call it the complement of A.
7. MATRICES OVER A COMMUTATIVE FIELD 73 We now have three methods for obtaining nxn matrices from a non-singular nxn matrix A. These are: forming the inverse, the transpose, and the complement. If we denote by 8 the operation of forming the inverse, T that of forming the transpose, and U that of forming the complement, we have TV = 8 = UTy US = T = SU, ST=U = TS, and S2 = 272 = *72 = I, where I is the identity operation which leaves A unaltered. The four operations I, S, T, U form an Abelian group. Other properties of these operations which can be verified immediately are: (A + B)' = A' + B', (AB)-1 = B-*A-\ (AB) = IB, where A, B are nxn matrices, which need not be regular for the first property to hold. It is only in the last case that K must be taken as commutative. The first two properties hold in any field. 8. Determinants. When the ground field K is commutative we are able to define determinants exactly as in the case in which K is the field of complex numbers. Determinants over a commutative field K have many properties which are similar to those of determinants defined over the field of complex numbers. It will therefore be sufl&cient to give a brief outline of the main results which we shall need in the sequel. We begin by defining certain symbols which are frequently used. (i) The symbol eH-in (OT€ilmmmin) is zero unless il9...9ini& some derangement of the first n natural numbers. If the derangement is an even permutation [I, § 1], ^-^ = ^...^=1. If the derangement is an odd permutation, eh-tn = e^ t_ = — 1.
74 n. LINEAR ALGEBRA, MATRICES, DETERMINANTS Both notations will be used as may be convenient. In tensor theory there is an important difference to be observed in the uses of the tjro symbols, but this difference does not concern us. (ii) The symbol (n) is zero unless il9..., ip and jx,..., jp are derangements of the same set of p distinct integers, each one of which lies between 1 and n (inclusive). When the derangements are of like parity, <n) and when they are of unlike parity, Evidently **:::fc--i. (n) (n) The (n) beneath the ^-symbol is usually omitted when there is no doubt about the value intended. Now let A = (dy) be an n x n matrix. We define its determinant to be det^ = «12 a22 nl U/n2 n n -in aln a2n «v.i -"Obi From this definition it follows at once that if B = (btj) is a p x q matrix, and A is the nxn submatrix whose elements are in the ixth,..., inth rows and jjth,...,jnth columns of B, then detA=i i ftr.th^-b^. Theorem I. If A is an nxn matrix, A' its transpose, det4 =detu4'. In the expression for det.4 consider any term ^M,i"flW'V
8. DETERMINANTS 75 in which e*!—*» is not zero. Then il9...,in is a derangement of 1,...,7i, defining a substitution [I, § 1] (1 2 • n\. Let the inverse substitution be /1 2 . n\ Then these substitutions have the same parity, and therefore If we rearrange the factors of ailX... ainn, we obtain the equation ahl---ainn =zalj1---anjn> and eh-in a{il...ainTt = eh-Ualh...anJn. Conversely, any term eh-lnalh...anjn in which the e-symbol is not zero determines uniquely a term Therefore eH-Haiil...ainn. det^4= S ... iev»W.1(„fl{ = S ... S^-^aiii...alin = S ••• £ ^.-^^^...^^ = det^'. Theorem II (Laplace's expansion): an al2 ^21 ^22 anl an2 a In a 2n a, iih a a, iph hJp lipjp a, ip+iip+i inJp+i a i>p+1 in linin where jl9..., jn is any fixed derangement of I,...,n, and the summation is over all possible divisions of 1,..., n into two sets ix < i2... < ip and ip+l <... <in
76 n. LINEAR ALGEBRA, MATRICES, DETERMINANTS (so that there are (%) terms) and pi = +1 or — 1 according as iv ...,in and jl9 ..., jn are like or unlike derangements of 1,..., n. By rearranging the factors ati in the expression for det^4 we find that n n &QtA = 2 ... 2 6{l-{nail...ainn Since n n det^l = e'i-..*n S ... S e*i-*»flfeft...aWll. In the summation on the right consider the p\ x (n —p)\ terms in which kl9...9kp are derangements of il9...9ip9 and A?p+1, ...,A;n are derangements of ip+l9...,in. These terms sum to ull...tn **!*! %il ^ *ii* a, and hence detii = £p< a nil ipip a, a, ip+lfp+l a a, inip+i a iph hip aiPip a ip+i1p+i ain1p+i ip+lfn ainin ainin This result is usually referred to as the Laplace expansion in terms of the determinants of the jrfh,... 9jp th columns. Using Th. I we obtain a similar expression for det^4 in terms of determinants in p selected rows. Theorem III. If two columns (rows) of a determinant are the same9 the determinant is zero. We first prove the theorem for the case n = 2. By direct evaluation a a b b == ab — ab = 0.
8. DETERMINANTS 77 Now suppose that in a determinant of n rows and columns the Ath and &th columns are identical. Expanding the determinant in terms of these two columns we have *ii a In a nl a, = 2± ahh aixk aizh aizk ahU a< ahin a* where h, k, jz,, jn is a derangement of 1,..., n. Since ai2h ai2k = 0 for all values of iv i2, the theorem, as applied to columns, follows. The result for rows follows from Th. I. If we expand det A in terms of the ^th column, det^ = 2(-l)m^ a li nh-i a lh+l a a, <-ii a a, a i+n *nl i~l h-1 "i-1 h+1 ai+l h-1 ai+l h+1 an h-1 an h+1 In H-ln a i+ln a* If we write the coefficient of aih in this expression as Ahi, we have n S%iw = deti. i-l On the other hand, if h 4= &, n Sa^w = det5, i=i where B is the matrix obtained from A by replacing the hth column by the kth column. By Th. Ill, det B = 0, if h*k. Hence n £^*4w = <^detA i-i Expanding in terms of rows we get, similarly, the equation n <-i We call A^ the signed minor or cofactor of a^ in det A.
78 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS We now prove Theorem IV. If A, B, C are nxn matrices, and C = AB, then detC = detAdetB. If G Hc,) = (s^O, detC= £ ... S^-^^i-V = £•••£ S-.^'-Sa^"'^.^ = S-. S^-^ndet^^!...^) = det^4detJ5. Theorem V. // J. is any matrix with n columns, B any matrix with n rows, any t-rowed determinant of the matrix AB is equal to a sum of terms, each of which is the product of a t-rowed determinant of A by a t-rowed determinant of B. Let A = (a^) be a p x n matrix, B = (by) smnxq matrix, so that 0 = (ci}) = AB is a p x q matrix. Any Crowed determinant D extracted from G is of the form /) = Chkx • %kt ' • • Ciikt • citkt 1 that is = £ ... £ £ - £ WKi.-^)^-6*^ s... s ii-i ii-i a, hh a Uh aixU (bh*i"'bit*)*
8. DETERMINANTS 79 Now a. hh "hit a, ith a, Uh 2«' a "HOI aUax a hot a, itOt where the summation 2 is over the I J sets of distinct gv ...,gt which can be chosen from the set (1,...,n). Hence Z> = 2 2 -. S a. not aitoi a, hot a U at *to?<W»W 0 a. hOi a hat a, itOi a, itOt Jgxki JOikt "ffikt . k Otkt We have proved then that bhh Ciikt Hxkt %kt a HQ\ a. ilOt a> HQ\ a not ^ifci Jatki J0ikt Jotkt and this is the form in which we shall subsequently use this theorem. The values of certain determinants will be needed in the next section, and we state these values here. They can be obtained at once, expanding the determinants by rows or columns. (i) ax 0 0 a2 • • 0 0 0 0 *1«>2 ' This determinant has non-zero elements only along the principal diagonal {au = at, atj = 0 (*#.?))• (ii) «1 0 0 0 0 0 Oa • • 0 . ah • • a . ak , • • • • an = ajflg...^.
80 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS Non-zero elements occur only along the principal diagonal and in one other place {au = ai9 ahk = a for a particular hy k). (iii) If A = (a^), where then detu4 = e*i •••*». With these preliminaries, we consider certain operations on a matrix A which correspond to the elementary transformations discussed in § 4. The operations are: (a) interchanging two rows of A, say the hth and kth rows; (6) multiplying the elements of the hth row by an element a of K\ (c) adding to the elements of the hth row a linear combination of the corresponding elements of the other rows. The operation (a) is equivalent to premultiplying A by the matrix B=(Shij) (»,j=l,...,n), where (hv ...,hn) = (1, ...,^- 1,k,h+ 1,...,k— l,h9k + 1,...,n). That is, B is the unit matrix with the hth and &th rows interchanged. By (iii) above, detU = -1, and therefore If the two rows of a determinant are interchanged the sign of the determinant is changed. Operation (b) is equivalent to premultiplying A by JB= A 0 . . 0 / 0 b2 . . 0 \ 0 . . 6^ 0 \0 . . . bn where bi = 1 if i + h, bh = a. By (i), det B = a, and therefore If the elements of a row of a determinant are multiplied by a constant a, the determinant is multiplied by a. Operation (c) can be split up into a set of operations, each of which consists of adding to the hth row a multiple bk of the corresponding element of the kth row, k taking the values (1,...,h— 1, h +1,...,n). For a given k each operation is equivalent to premultiplying A by a matrix B of the type considered in (ii) above. B has 1 in each place on the principal diagonal, bk in the ^th row and fcth column, and zeros in all other places. But det -5=1, and so
8. DETERMINANTS 81 Adding to the dements of any row of A a linear combination of the corresponding elements of the remaining rows does not alter the value of det A. All these operations may be performed on columns instead of rows. If a given operation on rows is carried out by premultiplying A by a matrix B, postmultiplying A by B' gives the corresponding operation for columns. Since deti5 = detJB', the results stated above remain true when 'column' is read for 'row'. We now discuss some properties of the cofactors Ait of the elements a{j of a determinant. If A = (atj)9 the matrix A* = (Atj) is called the adjugate of A. We saw that n n i-l i-l Hence AA* = (det A) In = A*A. If, therefore, det A 4= 0, the matrix \detii/ is the inverse A"1 of A. The determinant of this matrix must be (det A )-1, for if A is regular, so that A{A~i) = /n, then det A det (A'1) = det 4 = 1, and therefore det (A-1) = (det ^4)-1. On the other hand, if A has an inverse U, we have det A det J5 = 1, and hence det A =f= 0. We have thus proved Theorem VI. An nxn matrix A is non-singular if and only if det A is not zero. Let us now examine the case in which detu4 = 0. Then n S%\ = 0 (h,P= 1,...,n).
82 n. LINEAR ALGEBRA, MATRICES, DETERMINANTS Now the cofactor Aiq is itself a determinant, and contains all the rows and columns of A except the qth row and the jth column. If we multiply the above equation by the cofactor of a^ in Ajq, and sum over the following values of h: h= 1, ...,0-1,0+1, ...,n, we obtain the equation iAj§ 4.-0. where AjA, I denotes the determinant obtained from Ajq by replacing the element a^ by %, for each value of I. Since AjQ\k) " A*9 and otherwise the equation Ai<Kk) """^, a4)=u> reduces to the equation Hence any two rows (or columns) of the adjugate matrix A* are proportional, and we have Theorem VII. If detA = 0, and p>l, the determinant of any pxp 8ubmatrix of A* is zero. We now prove Theorem VIII (JacobVs Theorem): • • • a jp+iip+i a 'jp+ltn a inip+l a 'in in (det^)*-1, where iv...,in and j1}... fjn are any derangements of the set {!,... ,n).
8. DETERMINANTS 83 Let B be the matrix obtained from A by interchanging the ith and jth rows of A. If, with the usual notation, the elements of the adjugate matrix jB* are represented by Br8, it is easily seen that and Brj = — Ar{. Assuming Jacobi's Theorem to be true for the matrix B, we readily prove its truth for the matrix A, and conversely. A similar result is obtained by interchanging any two columns of A. Hence, if we prove the theorem when the theorem as stated above follows. If p = l the theorem is true, by definition of Aiih. If p > 1 and det A = 0, it follows from Theorem VII. Assume that det ^4 + 0. Then An . Apl , 0 0 ^ip A pp 0 0 • 1 0 A •"nn A -^pn 0 l a li «n\ det A 0 0 det A 0 a, 'p+ii Ln\ det A a. and therefore p+ip anp a m a. nn 0 a p+ln Lll 1P1 LlP ''PP det A = a p+ip+i a a np+1 p+ln (det4)* Since det A + 0, the theorem follows. As a special case we have the result det;l*-(det4)»-*. We have already shown [Th. VI] that emnxn matrix A is non- singular, and therefore of rank n, if and only if det A is different
84 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS from zero. We now obtain further results connecting determinants with the rank of any matrix. If A is a p x q matrix of rank r, there exists at least one rxr submatrix of A which is of rank r, and any 8X8 submatrix of A (s> r) is of rank not greater than r, and therefore less than 8 [§5, Th. III]. It follows from Th. VI that the determinant of any 8X8 submatrix must be zero. Hence we have Theorem IX. A pxq matrix A is of rank r if there exists in it an rxr submatrix with determinant different from zero, and if every 8X8 submatrix (s > r) has zero determinant. If A = (a^.), and the matrix a Hh *<rh a. Hil • "ixq h1r %ir a* *0, \a, a. is of rank r. Hence the vectors given by the ixth,...,irth rows of A are linearly independent. But since A is of rank r, the vectors determined by the rows of A form a manifold of dimension r, and therefore any row is linearly dependent on the ixth,...,irth rows. It follows from this reasoning that to solve the equations ailx1+...+aiqxq^ 0 (i = 1,...,^), (1) we have only to solve the equations for i = iv ...,ir. To simplify the notation let us write and consider the solution of the equations &ti*i+- ^ + biqxq = 0 (i= 1,..., r). (2) Then Jih 9rh JUr *0. By renaming the xi we may suppose that jp = p (p = 1,..., r). Now fb€j 6 = 16, let Bji be the cofactor of btj in ^n "n
8. DETERMINANTS 85 If (xv ..., xq) is any solution of the set of equations bax1+...+biqxq = 0 (i= 1,..., r), (3) then 2 £ BnbyXi = 0 {h = 1, ...,r), that is, 3 ^u • ^1/1-1 ^ii blh+1 . 6lr ^ = 0. (4) br\ - Kh-i bri brhJrl . br Conversely, let xr+1, ...,xq be any q — r elements of K, and let xl9...,xr be defined by the above equations (4). By direct substitution we see that (xv ...,¾) is then a solution of (3). Thus we see how to construct explicitly all the solutions of (1). Theorem X. A necessary and sufficient condition that the non- homogeneous equations Q 2 ai]xj^ci (i= 1,...,11) have a solution is that the matrices . alq\ and /an . alq \apl • apq/ \apl • apq have the same rank. Let us refer to the two matrices in question as the matrix of coefficients and the augmented matrix. The rank of the augmented matrix cannot be less than that of the matrix of coefficients. If the equations have a solution, the last column of the augmented matrix is a linear combination of the remaining columns of the matrix. The rank of the augmented matrix is therefore equal to the rank of the matrix of coefficients. On the other hand, if we are given the equality of the ranks, we deduce that the last column of the augmented matrix is linearly dependent on the first q columns, and if the equations 9 c<= 2¾¾ (* = !> •••»!>) express this linear dependence, (£x,..., £q) is a solution of the given equations. None of the results proved so far in this section requires any restriction on the characteristic of K. We conclude with a theorem
86 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS which is only true if K is not of characteristic 2. First of all, an nxn matrix A is said to be symmetric if A = A', skew-symmetric if A — —A'. If A is skew-symmetric, detA = det(-A)' = det(-/Jdet.4' = (-l)~detA Hence, if n is odd, 2det4 = 0, and therefore, if K is not o/ characteristic 2, detA = 0. On the other hand, a , x \ A= a h g\ [h b f) \g f c/ is skew if K is of characteristic 2, since any element I of K satisfies the relation ? — _ 7 But det A = abc + a/2 + bg2 + ch2, which is not zero for all a, 6, c,/, gr, ^ in jfiT. In fact, if a = 6 = c= 1; /=gr = & = 0, det^l = 1. This brief account of determinantal theory establishes the main results which we shall use in later chapters., We have shown that nearly all the results proved in textbooks for determinants whose elements are real or complex numbers extend to determinants over any commutative field. The many theorems on special methods of expansion, and on the value of determinants of special type (e.g. Vandermonde's determinant) which can be found in textbooks can easily be deduced from the above theorems for determinants over any commutative field. When necessary, such special results will be assumed in the sequel. 9. A-matrices. Let K be a commutative field, and let A be an indeterminate over K. In the following chapters we shall often have to consider matrices whose elements are polynomials in A, and it is convenient to obtain here a number of results which will frequently be used.
9. A-MATRICES 87 Strictly speaking, we have so far only defined matrices over a field. We can, however, deduce the basic properties of matrices whose elements are in an integral domain /, say K[X]y simply by regarding these matrices as forming a subset of the set of matrices whose elements are in the quotient field of the integral domain. Addition and multiplication of matrices of the subset give further matrices of the subset. A matrix whose elements are in the integral domain K[X] is called a \-matrix. An nxn A-matrix which has an inverse which is also a A-matrix is called a regular A-matrix, of order n. Theorem I. A necessary and sufficient condition for the nxn X-matrix M to be a regular \-matrix is that det M is a non-zero element of K. (i) Suppose that M is regular, and let N be its inverse. Since the elements of M and N are polynomials in A, the determinants of these matrices are polynomials in A. Let det M = /(A), det N = g(\), where/(A) is of degree r, g(X) of degree s. Then since MN = /n, detM det# = 1, that is, /(A)flr(A) = 1. But one side of this equation is a polynomial of degree r + s, while the other side is a polynomial of degree 0. Hence r = s = 0, and/(A), g( A) are non-zero elements of K. The condition is therefore necessary. (ii) Suppose that detM = m, where m is a non-zero element of K. Let M = (m^), and write n^. = (-)^01-1 m ii mn-i m li+l %-n m>j-u-i mj-n+i m i+n m 'i+i i-i mj+i i+i • • • • mn\ • mni-\ mni+l Then N = (n{j) is a A-matrix, and [§ 8, Th. VI] m m m, j-m m i+ln MN = In.
88 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS We now define equivalent A-matrices. Two p x q A-matrices A and B are said to be equivalent if there exist regular A-matrices M $nd JV, of orders p and q respectively, such that A = M BN. (1) This relation is reflexive, since A = IpAIr From (1) we have B = M-1AN^ and Jkf~\ N'1 are regular A-matrices, since M, iV are regular. The relation is therefore symmetric. Again, if besides (1) we also have B = PCQ, where P and Q are regular, then A = (MP)0(QN)9 where MP, QN are regular A-matrices. The relation is therefore transitive, and it is thus an equivalence relation. We wish to find a set of canonical forms for all p x q A-matrices. As in § 4 we say that a set of p x q A-matrices is a set of canonical forms for the set of allpxq A-matrices if (a) no two matrices of the set are equivalent; (b) there is a matrix of the set equivalent to any given p x q A-matrix. Clearly, if we can construct such a canonical set, and also find a method for determining the canonical form equivalent to any pxq A-matrix, we shall have a method for determining whether two given A-matrices are equivalent. Our first step is to consider certain simple operations on a A- matrix A. These are known as elementary transformations [cf. §4], and produce a A-matrix B equivalent to A. (i) Let kl9..., kp be any derangement of the integers 1,...,#, and let P be thepxp matrix Then detP = ± 1, and the matrix B = PA is the matrix obtained from A by performing the permutation II 2 . p\ on the rows. Hence, the elementary transformation of A which
9. A-MATRICES 89 consists of a permutation of the rows of A transforms A into an equivalent matrix. (ii) A permutation of the columns of A transforms A into an equivalent matrix. The proof is similar to that in (i), the new matrix being of the form AQ. (iii) Let R be the pxp matrix (r^) where Ui = $a (***)• rhj=fj&) (j= 1>- • •> *- 1.A+ !>•• •> P)f where/^(A) is a polynomial in K[X], and rhh — fhft) = a (a^n K> and not zero). Then det R = a, and R is therefore a regular A-matrix. We obtain a matrix B = RA equivalent to A by multiplying the ^th row by a and adding to it/^A) times the first row, /2(A) times the second row,..., A-i(A) times the (h— l)th row, fh+1(h) times the (h+l)th row, ..., /P(A) times the pth row. (iv) If a column of A is multiplied by a non-zero element of Ky and to it we add a multiple (by a polynomial) of any other column, we obtain a matrix, equivalent to A, of the form AS. The proof is similar to that in (iii). We now show how a matrix can be simplified by a series of elementary transformations until it is reduced to a canonical form. Let A be a p x q A-matrix. If it is the zero matrix, any matrix equivalent to it is clearly zero, and no reduction is possible. We therefore assume that at least one element of A is not zero, and by a permutation of rows and columns we bring a non-zero element of A to the top left-hand corner. Let this element an be of degree p in A. We show that unless aiv regarded as a polynomial in A, is a factor of every element ai;j of A, we can transform A into an equivalent matrix C, where cn is not zero and is of lower degree than au. Our transformation will be a series of elementary transformations. (i) If, for some j, an is not a factor of aip then aii = an6(A) + c(A), by the division algorithm, where c(A) is not zero, and is of lower degree than alv From the jth column we subtract 6(A) times the first column, and then interchange the first and jth columns. The resulting matrix is equivalent to Ay and has c(A) in the top left-hand corner.
90 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS (ii) If, for some i, an is xiot a factor of ail9 we operate on rows instead of columns. (iii) Suppose that an is a factor of all non-zero elements in the first row and column, but is not a factor of ahk (h,k>l). Let aix = 6ia11, and axi — fy^n- Then if we subtract from the ith row bi times the first row, for each i> 1, and then subtract from the jth column c^ times the first column, for each j>l, we obtain a matrix B equivalent to A in which 611 = «n> bu = 0 (j> 1), ba = 0 (i > 1), while bhk is still not divisible by blv Now add the hth row of B to the first row, and apply method (i). We obtain a matrix C equivalent to A9 with cn # 0, and of lower degree than alv Since there is a finite upper bound to the degree in A of the elements of the matrix A, the processes described above cannot be repeated more than a finite number of times. We eventually obtain a matrix J5, say, equivalent to ^4, in which 6n is a factor of every element by. By (iii) we can arrange that 6n = 0 (»>1), 6y = 0 (i>l). Thus a series of elementary transformations has produced from A the equivalent matrix PAQ= /bn 0 0 0 622 . 0 . . 0 0 bp2 . where 6n is a factor of all the elements 6 If the submatrix D ,, ^1 = /&22 • \bpz . is not zero, we can repeat the process given above, until we obtain an equivalent matrix \o 0J9 each element of Gt containing c22 as a factor. Then /1 0\PAQ(1 0\=/6n 0 0\ [0 pj [0 Qj 0 c22 0 , \o 0 cj
9. A-MATRIOES 01 so that the elementary transformations on the submatrix Bx can be considered as elementary transformations on A. Proceeding in this way we can transform A into an equivalent matrix in which the only non-zero elements are on the principal diagonal, and each element on the diagonal is a factor of all those following it. Finally, multiplying each row of this matrix by a non-zero constant, A is equivalent to the matrix MAN = '^(A) 0 0 ES{A) 0 0 EM) 01 0 Oj (2) where each 2^(A) is a polynomial in A with + 1 as the coefficient of the highest power of A, and 2^(A) is a factor of 2?m(A). Since equivalent A-matrices have the same rank, and the rank of the matrix MAN is evidently r, it follows that r is the rank of A. Again, writing MAN = E = (Cy)> an evident extension of § 8, Th. V gives us the theorem *iih *<th = 22 A /* m. ix\x m it Ai ™*iA* mU*t a*i Pi a AtPi aAx^ a\tHt n >i/i n nit n Hh n Mt it It follows that if Dt(\) is the highest common factor of the determinants of order t formed from A,Dt(\) is also a factor of each determinant of order t extracted from E. But since if, N are regular, the equation can be written as MAN = E A = M^EN-\ and therefore the highest common factor of determinants of order t formed from E divides each determinant of order t formed from A. Hence, save for a factor in K, this highest common factor is Dt(A). But from the properties of the polynomials 2^(A), Dt\)-E1(A)E%(\)...EA\)9
|j8 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS provided that the coefficient of the highest power of A in Dt(X) is taken to be +1. Therefore Et{X) = DtMID^iX) (i =l,...,r), defining D0(X) = 1. This last result shows us that the JE?i( A) are uniquely defined by means of the highest common factors of the determinants of different orders extracted from A. Calling the E{( A) the invariant factors of A we have now proved Theorem II. If Aisapxq X-matrix of rank r, and E^X)*..., Er(X) are its invariant factors, there exist regular X-matrices M, N such that MAN = /J^(A) 0 . 0 ^»(A) 0 EM) °i Since the inverse of a regular A-matrix is a regular A-matrix we also have Theorem III. Two pxq X-matrices A and B are equivalent if and only if they have the same invariant factors. Hence the set of matrices (2) is a set of canonical forms for pxq A-matrices. If the field K is such that every polynomial /(A) in K[X] can be written as a product of linear factors, it is sometimes convenient to use elementary divisors in place of invariant factors. We define these as follows. Let E^X) = (A-aiMA--a2)«i2... (A-a8)«u, E2(X) = (A - a^i (A - a2)***... (A - as)*«, Er(X) = (A - o^fn (A - a,)*.... (A - a,)*«, where each etj^0, and A — ax,...,X — a8 are the distinct linear factors of Er(X)> that *s> erj > ® (J = 1> •••>*)• Then those factors (A-a^i, (A-a^i, ..., (A-a2)«u, ..., (A-asK for which the indices etj are greater than zero are called the elementary divisors of A. By definition, these are determined by the invariant factors.
9. A-MATRIOES 93 Suppose, conversely, that we are given the elementary divisors of A, which is known to be of rank r. Let them be (A-a)0!, (A-a)°2, ..., (A-a)°< K^a2^ ... >af), (A-*)*, (A-*)*, ..., (rf^rf2^...). Each invariant factor JE^(A) must be a product of these factors. Since Er(\) contains 2^(A) as a factor (i < r), Er(\) must contain the highest power of A — a, the highest power of A — /?, etc. Hence ^(A) = (A - a)«i (A - /?)*i... (A - S)dK Similarly, since -6^( A) is divisible by E^X) (i < r — 1), ^U(A) = (A-a)«2 (A-/?)**... (A-eJ)d», and so on. Once all the elementary divisors which are powers of A —a are used up, the remaining E^X) do not contain A —a as a factor. Similarly for A—/?,.... When all the elementary divisors have been used up, the remaining invariant factors (if any) are Et(X) = 1. An equivalent form of Th. Ill, is therefore Theorem IV. A necessary and sufficient condition that two pxq X-matrices be equivalent is that they have the same rank and the same elementary divisors. Our next theorem is often useful in subsequent chapters. Theorem V. IfA, B are two nxn X-matrices which can be written A = A^-A^ B = BXX-B2, where Av A2, Bv B2 are nxn matrices over K, and Bx is non-singular, then the equivalence of A and B implies the existence of non-singular matrices P, Q over K such that PAQ = B. Since A and B are equivalent, there exist regular A-matrices M, N such that MAJjf = A Let X = B2Bi\ Then if M = JMi + JMiA+...+JH;Ar, where the Mt are matrices over K, and if p = MrX~i + {Mr_1 + XMr)\'-*+. .. + (M1 + XMz+... + X'~1Mr),
94 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS and P = M0 + XM1+...+X'M„ then M = (A7n - X)p + P = (A/n - B^r1) B1 B^p + P, that is M=BP1 + P, where Px = B^p. Similarly, let 7 = B^B2, and write q = N8M-i+{Ns_1 + Ns Y)\'-*+... + (Nt + Nt Y+... +N8 7-»), Q = N0+N1Y+...+NaY: Then we have N = q(\In-Y) + Q = Q1B+Q, where Qt = qB^1. Now PAQ-B = (M-BP1)A(N-Q1B)-B = MAN-MAQXB- BPXAN+BP^AQ^- B = BPXAQXB- BN-^ B - BPX M~XB (since MAN = B). Thus PAQ-B = BRB, where 22 = P^Q^N-^-P^M-1, and since M and jV are regular A-matrices, R is a A-matrix. Suppose that R is not zero, say R = RQ + R1X+...+R(\t (Rt*0,t>0). Then, from the above, PAQ-B = BRB, where the left-hand side is of degree 1 at most in A, whilst the right- hand side is of degree This contradiction shows that R = 0, and therefore PAQ = B. Since P, Q are independent of A, PA1Q = BV PAiQ = Bt. Since Bt is not singular, P, Q are not singular, and the theorem is proved.
10. MISCELLANEOUS THEOREMS 95 10. Miscellaneous theorems. We conclude this chapter with a number of results which will be needed later. Some of them require the use of theorems which are more appropriately dealt with in later chapters, and it will avoid repetition if we borrow these theorems without proof at the present stage. If A is any nxn matrix over K, the A-matrix A-Mn is called the characteristic matrix of A, and its determinant /(A) = det(,4-A4) is called the characteristic function of A. The roots of the character- istic equation -,. x A / (A) = u are called the latent roots of A. Let us write /(A) = a0 + a±\+ ...+anAn. Form the adjugate matrix of A — Mn. It is a A-matrix whose elements are at most of degree n — 1. We may therefore write it as B0 + B1A+... + Bn_1 A"-1, where the Bi are nxn matrices over K. Remembering that {A - A/n) adjugate (4 - AIJ = det (A - AJJ Jn, we have the equation (A-Mj(B0 + B1\+... + Bn_1A«-i)=f(A)In. Equating like powers of A on both sides of this equation, we have AB0 = a0In, 4^-^0 = ^4, AB2 — B1 = a2In, Premultiply these equations by Jn, A, A2,..., An respectively, and add. We obtain the matrix equation 0 = a0In + a1A+azA2+...+anAn. This result is usually stated in the form Theorem I. A square matrix satisfies its own characteristic equation.
96 H. LINEAR ALGEBRA, MATRICES, DETERMINANTS We note that since the characteristic function /(A) = a0 + at A +... + an Xn = det (A - A/n), /(0) = a0 = det A. Hence A = 0 is a root of the characteristic equation if and only if A is singular. We next introduce the idea of the square root of a matrix. Assuming the ground field K to be not only commutative but algebraically closed, so that every polynomial <j>(x) in K[x] is expressible as the product of linear factors, we shall prove the following theorem in the next chapter [III, § 8, Th. II]: If(j){x) is any polynomial in K[x] such that <f>(0) ^10, then there exists a polynomial g(x) such that [g(x)]r — xis divisible by <p(x), r being any given positive integer. We use the case r = 2 of this theorem and take $(x) to be the characteristic function f(x) of a non-singular matrix A. Then we can find g(x) such that \g(x)]*-x=f(x)h(x). Since f(A) = 0, \g(A)]*-A = 0, that is, [g{A)f = A. Hence we have Theorem II. If A is a non-singular matrix over an algebraically closed field K, there exists a matrix X such that X* = A. Finally, we introduce the idea of the minimum function of a matrix A. If ^(A) = A"»+«4 A"*-1 +...+am is the polynomial of least degree for which f{A) = 0, we call ^(A) the minimum function of the matrix A. Let the invariant factors of A — XIn which are of degree > 0 be E8+lW> •••> EnW-
10. MISCELLANEOUS THEOREMS We shall see in VIII, § 2, that there exists a matrix B, where 97 IB. 8+1 0 0 B S+2 0 . . Bn, such that (a) B = P-XAP for some non-singular P over K9 (6) det(J3<-A2) = ±^(A) (i = $+1,...,n). In fact, we need only take 3,= 0 1 where Et( A) = Ar< + an A'*-1 +, By Theorem I, ^^ = 0 We observe that 2^(A) is the minimum function for the matrix Bt. To see this, we first note that when r < ri the rth power of Bi has the form (BJ = (A In.r\ \B C )' where A is the zero (^ — ^xr matrix, Ir._r is the unit matrix of ri — r rows and columns, and B and C are, respectively, rxr and rx{ri — r) matrices whose elements are polynomials in the a{j. Now let /(A) = &0 + M+-+M8 (&s*0) be a polynomial of degree s over K, where s < r{. Then it follows that f{Bt) is a ri x ri matrix having the (s + l)th element in the first row equal to bs. Hence/(2¾ cannot be zero. Again, if g(X) is any polynomial over K such that g(B{) = 0, we can write ,^ /-*\ rw>\ , t/*\ gr(A) = a(A)i^(A) + 6(A), where 6(A) is of degree less than rv Then 6(5,)-^)-^)^(¾) = 0, and by the result just proved it follows that 6(A) » 0. Thus g(X) must have Et(\) as a factor. HP I 7
98 II. LINEAR ALGEBRA, MATRICES, DETERMINANTS Now suppose that F{\) is any polynomial over K such that F(A) = 0. We have F(B) = /F(Ba+1) 0 . 0 0 F(Ba+2) . 0 • • • • o . . F(Bn)/ = FiP-iAP) = P-*F{A) P = 0, and hence 1^(1^) = 0 {i = s+ 1, ...,7&). Thus i*\A) must have ^s+i(A),^+2(A), ...,JSJn(A) as factors, and conversely, if J7(A) contains these polynomials as factors, F(B) = 0. Since E^A) is a factor of Ej(\) for i <j we obtain Theorem III. If A is an nxn matrix over K, and A — XIn has the invariant factors ^(A),..., En(\), En(X) is the minimum function of A, and a necessary and sufficient condition that a polynomial f'(A) over K be such that f {A) is zero is thatf{\) has En(\) as a factor. *
* CHAPTER HI ALGEBRAIC DEPENDENCE In this chapter it is assumed that all fields are commutative. 1 • Simple algebraic extensions. In I, § 9, we discussed a method of extending a commutative field K by the adjunction of an indeterminate x. We now describe another method by which we can extend a commutative field K. Let f(x) be an irreducible polynomial of degree n in K[x]. We use f(x) to define an equivalence relation in K[x]. Two polynomials a(x) and b(x) of K[x] are to be equivalent if a{x)-b{x) = c(x)f{x), where c(x) is in K[x]. This relation is clearly reflexive, symmetric, and transitive, and it is therefore a proper equivalence relation [I, §2]. The polynomials of K[x] are divided into classes by this relation. We denote a class of equivalent polynomials by {a(x)}9 where a(x) is any member of the class. If a(x), a'(x) are equivalent polynomials, and b(x), b'(x) are also equivalent, a'{x) = a(x) + c(x)f(x)y and b'(x) = b{x) + d(x)f(x). Hence a'(x) + b'{x) = a(x) + b(x) + [c{x) + d(x)] f(x), and a'{x)b\x) = a(x)b(x) + g(x)f(x), where g{x) = a(x)d(x) + b(x)c(x) + c(x)d(x)f(x). Hence we may define addition of classes of equivalent polynomials unambiguously by the equation {a(x)} + {b(x)} = {a(x) + b(x)}y and multiplication by the equation {a(x)}{b(x)} = {a(x)b(x)}. It may be verified without difficulty that the classes form an Abelian group under addition, and that multiplication is associative, commutative, and distributive over addition. Hence the classes form a commutative ring with the given laws of composition. The zero 7-2
100 m. ALGEBRAIC DEPENDENCE of the ring is the class {0}, that is, the class of polynomials divisible by/(#). There is also a unity, represented by {1}. Now let {g(x)} be any non-zero class. Then g(x) is not divisible by f(x). Since f(x) is irreducible it follows that the highest common factor of f(x) and g(x) lies in if, and therefore polynomials a(x), b(x) can be found so that a{x)g{x) + b{x)f(x) = 1. We deduce that Wx)}{g{z)} = {l}. Thus every non-zero element of the ring of equivalent classes of polynomials has an inverse. That is, the ring is a field, which we denote by if'. If a, J3 are any elements of K, {«} = w if and only if a —ft is divisible by f(x); that is, if and only if a = /?. Hence the classes {a} form a subfield of K' isomorphic with K. By I, § 2, Th. I, we deduce the existence of an extension if* of K which is isomorphic with if'. The field if* is uniquely defined to within an isomorphism in which the elements of K are self-corresponding. Let £ denote the element of if* which corresponds to {x} in if'. Then, if g(x) is any polynomial of K[x], g(£) corresponds to {g(x)}. By the division algorithm we can write g{x) = h(x)f(x) + gx(x)y where gx{x) is a polynomial of degree less than n. If g(x) itself is of degree less than n9 we take h(x) = 0, g(x) = gx{x). Since {g(v)} = {0i(z)}> we have g(£) = ^(g). It follows that any element tj of K * can be written in the form V = <*0 + axi + ... + ctn^-1 (¾ inK)y (1) and, of course, the right-hand side of this equation, for any <*o> •jtfn-i in K» represents an element of if*. Now the element /(g) of if* corresponds to the element {/(#)} of if'. Since M) = o.
1. SIMPLE ALGEBRAIC EXTENSIONS 101 Let F(x) be any polynomial over K such that m = o. Then {F(x)} = {0}, and hence F(x) is divisible by f(x). F(x) is therefore zero or of degree n at least, and it follows that the representation (1) of the elements of K* is unique. Equation (1) thus gives a very convenient means of representing the elements of If*. Using this representation we have (*0 + x1£+... + 0^^) + (^ + ^+. -+/^-1^-1) ==(^0+^0) + (^1 + ^1)^+... + (^-1+^1)^^- To construct the product we have (^+^5+... + ^.^^)(^0+^+...+^1^-1) = cc0J30 + KA + ccj0) I +... + an_Jn^ £2""2, and we then use the relations 0 = ?M) = y0£r+yi£r+1+...+7n£n+r (r = n-2,^-3,...,0) to express £2n~2 in terms of £n~2,..., £2n-3, and then £2n-3 in terms of £n~3,..., Pn_4, and finally to eliminate powers of £ of degree higher than n—l from the product. A simple example will serve to show that for some fields K a true extension is constructed in this way. If K is the field of rational numbers, and /(0:) = ^-2, K* is the field of numbers where a and b are rational. K* is not equal to K. If, on the other hand, K is the field of complex numbers, every polynomial of degree greater than one is reducible, and no proper extension of K can be constructed. The method of extension described above is called simple algebraic extension. 2. Extensions of a commutative field. We now consider some general properties of extensions of a commutative field Ky first defining equivalent extensions. Extensions Kl9 K2,... of K are said to be equivalent if they are isomorphic in such a way that those elements of Kv K2,... which lie in K are self-corresponding.
102 m. ALGEBRAIC DEPENDENCE Let K* be any extension of K, and let £ be any element of K*. We say that £ is algebraic over K if there exists a non-zero polynomial f(x) in K[x] such that M) = o. If £ is not algebraic, it is transcendental over K. Clearly, any element a of K is algebraic over K; for we merely have to take £l x f(x) = x~a. Theorem I. If £ is algebraic over K> there exists an irreducible polynomial f(x) in K[x], unique save for a non-zero factor in K> such that j>/r. - /(§) = o. Since £ is algebraic over K there exists a non-zero polynomial g(x) in K[x] such that .*. _ 0 Let g(x) = gi(x)g*(x)---9^) represent the factorisation of g(x) into irreducible factors. Then g^(£) is in Z*, and ^)^).-^) = ^) = 0. Since K* is a field, and therefore has no divisors of zero, for some integer i (1 < i < r) we have 9M) = o. Take f(x) = ?,(*). If h(x) is any irreducible polynomial such that *(£) = o, then either ^(#) = a/(#) (a in i£), or ^(#) and f(x) are relatively prime. In the latter case there exist polynomials a(x) and b(x) such that a(x)f(x) + b(x)h(x) = 1. But this gives us the equation a(i)f{g) + b(i)h{g) = l, which, since the left-hand side is zero, is a contradiction. Hence f(x) is uniquely defined, save for a non-zero factor in K. It is called the characteristic polynomial of £ with respect to K.
2. EXTENSIONS OF A COMMUTATIVE FIELD 103 Now let £ denote any element of an extension K * of K. We consider all the elements of Z* which can be obtained by performing the operations of addition, subtraction and multiplication (a finite number of times) on £ and the elements of K. Such elements can be written - «., , . ,rx ,, v <*o + ai£+---+^ (otiinK). (1) Clearly these elements form a commutative ring with unity, the unity being that of K. If ?/, £ are two elements of this ring, the product ?/£ is equal to the product of these elements regarded as elements of K*y and therefore t/£ = 0 only if either r\ or £ is zero. The ring is therefore an integral domain. We denote it by K[g], noting that K[g] = K if and only if £ is in K. Since K[£] is an integral domain, it can be embedded in a field [I, §4]. We denote this quotient field by K(£)y and refer to it as the field obtained by adjoining £to K. Case (i). Let £ be algebraic over K, and let/(#) be its characteristic polynomial, of degree n. If 9(£) = ^ + ^^+...+^^, we can wri ^ = ^ ^ + ^^ where gx{x) is of degree less than n. Then and hence every element of K[£] can be expressed in the form (1), where r <n — 1. This expression is unique, for if any element could be so expressed in two ways, there would exist a non-zero polynomial g(x) of degree less than n such that QiZ) = 0, and this would contradict Th. I. It follows that the integral domain K[£] is equivalent to the simple algebraic extension of K defined, as in § 1, byf(x). Hence, in particular, K[£] is a field, and As an important corollary we have Theorem II. If £x and £2 are two elements of an extension of K which are algebraic over K and have the same characteristic polynomial, then If (£x) and K(£2) are equivalent extensions of K.
104 in. ALGEBRAIC DEPENDENCE Case (ii). Now suppose that £ is transcendental over K. Then ^ + ^+...+^ = /?0+A£+."+A£* (<*r>A*0), if and only if r = s and at = fi{ (i = 0,..., r). It follows that Z[£] is equivalent to the integral domain K[x], and therefore that K(£) is equivalent to the field of rational functions of the indeterminate x over K. The field J5T (£) is called a transcendental extension of K. In either case it is clear that K(E) is the smallest extension of K which contains £. There may, of course, be extensions of K which are proper subfields of !£(£), but these will not contain £. For instance, if £ is transcendental over K, Ze*(g») <=*(£). The ideas introduced above can be extended by considering the smallest extension of K which contains a finite set £x, ...,£n of elements of K*. The elements of this extension are those elements of K* which are obtained by applying the operations of addition, subtraction, multiplication and division (except by zero) to £lf..., £n and to the elements of K. The resulting field is denoted by K{£v •••y£n)> an(i evidently does not depend on the order of the elements £x, ..., £n. We can define the field K(£l9..., £n) inductively as follows: K(£v...,izr) is the field obtained by adjoining £r to #(£i,. -, £^)(^ = 2,...,^). Some of the £t- may be algebraic over K, and we shall usually find it convenient to order the £^ in the set gv ..., £n to satisfy certain conditions. If each §rf is algebraic, there is no restriction on the order. If some of the £^ are transcendental over K, we take one of them as £r If £if, ...,£*, are transcendental over K(£x) (and hence transcendental over K), we take one as £2, and continue thus. Eventually we can say that in K(£v ..., £n) the element £x is transcendental over K, the element £s is transcendental over K(£lf..., £a_j) (s = 2, ...,r), and £r+1,...,£n are algebraic over Jf(glf ...,£r). With this arrangement, #(£^...,£r) is equivalent to the field of rational functions ofxv...,xr with coefficients in K [I, § 9]. We have proved this result when r = 1, and we now assume it is true for r = s — 1. Then #(^,..., £*-i) is equivalent to the field of rational functions K(xv .-.,¾^). Now -KX£i> • • • > £*-i> £5) = ^(£i>..., £^) (£J, and since £s is transcendental over #(£^..., £^), #(£^..., £^) (£J is equivalent to #(£^ ...,^) (#s), and hence to ^(^1,...,^-1)(^) = ^(^,...,¾).
2. EXTENSIONS OF A COMMUTATIVE FIELD 105 The elements £v ..., £f are said to be independent indeterminates over K. In this book we shall only be concerned with the extensions of a field K obtained by adjoining a finite number of elements £x,..., £n toK. 3. Extensions of finite degree. An extension K* of K is said to be of finite degree if there exists a finite number of elements £i> •••> £n °f ^* such *ha* any element r\ of K* is expressible in the form r y ^ = <*l£l+...+an£n> the av ..., an lying in if. The elements £x,..., £n are said to form a basis for K* over K. This basis is a minimal basis if the equation 0^+...+0^ = 0 (¾ in if) implies that o^ = 0 (i = 1,..., w). The elements of if* clearly form a linear set £(ii,...,fj over K [II, § 1], and therefore if K* is of finite degree over K we deduce the existence of a minimal basis, the number of elements in the basis being the same for all minimal bases [II, § 1, Th. VI]. This number is called the degree of the extension K* of K. Theorem I. If if* is an extension of K of finite degree n, every element of K* is algebraic over K, and has a characteristic polynomial of degree not greater than n. Let £l9..., £n be a minimal basis for if* over if. Then, if tj is any element of if*, rj^ is an element of if* and Vii = <*ii£i+--- + ^-^ (*' = !> —>n). The linear equations Ki-7/)^+...+alnzn = 0, otnlx1+... + (ctnn-ri)xn = 09 have a solution in if*, namely, (£x, ...,£n), which is not trivial. Hence = 0. «11- a21 <*nl 7 «12 <*n-V «n2 • • . am a2»t «nn-7 i
106 m. ALGEBRAIC DEPENDENCE The polynomial F(x) = CC-t i *™~ X ^12 a In ^21 ^22 ^ • ^2 a, nl a /i2 a„ is not zero, the term of highest degree being ( — l)nxn. Since tj satisfies the equation ^,, H F{rj) = 0, tj is algebraic over jRl. Moreover, its characteristic polynomial is a factor of F(x); hence its degree cannot exceed n. An extension K* of K in which every element is algebraic over K is called an algebraic extension of K. We have proved that every extension of K of finite degree is algebraic. We now prove Theorem II. If the elements £l9 ...y£nofan extension K* of K are algebraic over K, then K(£l9..., £n) is an algebraic extension of K. We first prove that any element rj ofK(£l9..., £n) can be expressed in the form r_1 x the coefficients aix #An lying in 1£, and rl9..., rn being integers which are independent of the element tj. We prove this result by induction, assuming the truth of our statement for K(£l9..., £n_i). Since £n is algebraic over K, it is also algebraic over K(£v ..., £n_i). Let rn be the degree of its characteristic polynomial with respect to this field. Then, since K(£l9 ..., £n) is obtained by adjoining £n to K(E,l9..., £n_i), any element tj of K(£v ...,£n) can be written in the form 3-00 + 01^+- + ^-1^-1. the 0^ lying in K(£l9..., ^n_i). But by the hypothesis of induction we can write 0i=rii-..'*?: ^...in-^i1-"fir-i (* = o,...frm-i), where the integers r1?..., rn_i are independent of 0i9 and hence of r/, and the coefficients <x,il%inli lie in Z. Hence This proof is valid when n = 1 and therefore our result is proved.
3. EXTENSIONS OF FINITE DEGREE 107 It follows that the r1r2...rn elements &...& (^ = 0,...,/1-^ ...;»tt = 0,...,rn-l) form a basis for K(£l9..., £n). Hence, by Th. I, any element of ^£(£i> • • • > £n) iQ algebraic over If. By definition, therefore, K(£l9..., £n) is an algebraic extension of if, and the theorem is proved. Suppose now that of the elements £x,..., £n the first r, £x,..., £r, are independent indeterminates over K, but that £r+1,..., £n are algebraically dependent on^K(^v ..., £r). Then K (£1?..., £n) is an algebraic extension of If (£x, ..., £r). It is called an algebraic function field, and its elements are algebraic functions of the indeterminates £x,..., £r. Theorem III. If K* is any extension of K, the elements of K* which are algebraic over Kform afield K'y where If £ and r\ are elements of If* which are algebraic over K so is every element of If (£,?/), by the previous theorem. Hence, the elements £ —?/ and £?/, lying in If (£,?/), are algebraic over K. It follows that the elements of If* which are algebraic over K form an integral domain which contains K as a subring. Further, if £ is algebraic over K, but not zero, £-1 lies in if (£) and is therefore algebraic over K. Hence the set of algebraic elements of K* form a field K!. Clearly if c if'c if*. Theorem IV. If if* is any extension of K, and if' a subfield of K* whose elements are algebraic over if, then any dement of K* which is algebraic over if' is also algebraic over K. We do not assume that K' contains K. Let £ be any element of if* which is algebraic over if', and let xn + 7\xxn~x +...+i]n (^in if') be its characteristic polynomial with respect to if'. Consider the field K(rjv ...,?/„,£). Any element can be written as S6g<-* (^mK(Vl,...yVn)). But for each i we can write £<= £ .../£ 4...^^-1...^1 «...in inif), and therefore K(7/l9 ...,?/„,£) has a finite basis over if, and hence, by Th. I, it is algebraic. It follows that £ is algebraic over K.
108 III. ALGEBRAIC DEPENDENCE We now introduce the notion of algebraic dependence. A set of elements gx,..., £r of an extension K* of K is said to be algebraically dependent over K if there exists a non-zero polynomial f(xv ...,xr) in K[xl9 ...,xr] such that /(^,...,^) = 0. If not dependent the elements are algebraically independent over K. In the case r = 1, £2 is algebraically dependent over K if and only if it is algebraic over K. A proof similar to that of § 2, Th. I, enables us to show that if £x,..., £r are algebraically dependent over K there exists an irreducible polynomial f(xv ..., xr) such that /(£1(...,U = 0. If £x,..., £r are algebraically independent, §i> •••>§r are independent indeterminates over K, and therefore if(£x, ...,£r) is the field of rational functions of the indeterminates £l9..., £r. We have already remarked that there exists in K(£l9..., £n) a set of indeterminates, which we may take as ^,...,^,., which are algebraically independent over if, and which are such that any element of K(£l9...,£n) is algebraic over K(£l9 ...,£r). More generally, there may exist in any extension if* of if a finite set of elements £i> • • • > £r which are algebraically independent over If, and such that any element r/ of K* is algebraic over K(£ly...9£r). When this is the case we say that K* has & finite algebraic basis with respect to if, and since gx,...,£r are algebraically independent over if, these elements are said to form a minimal algebraic basis. Certain properties of extensions K* of K with finite algebraic bases are similar to properties of linear sets of finite dimension [II, § 1]. We now prove the generalised form of the Exchange Theorem. Theorem V. If 0/i,....,%,) is any algebraic basis for an extension K* of K9 and £l9..., (^ are t algebraically independent elements of if*, then JO, and for a suitable rearrangement of ?/i,...,%, the set (£i> •••>£/, Vt+i> • • > Vs) w a basis for if*. We consider the case £ = 1 first of all. Since £x is in if*, it is algebraically dependent over K(t/19...,%). Therefore there exists a non-zero polynomial/^,...,x8+1) such that /(?/i,...,^,£i) = 0. If s = 0, this implies that £x is algebraic over K9 which is contrary to hypothesis. Hence s ^ 1.
3. EXTENSIONS OF FINITE DEGREE 109 Again, and for the same reason, one at least of the indeterminates xv ..., x8 must be present inf(xv ..., x8+1). We may suppose that the rji are rearranged so that xx is present. Then it follows that ijx is algebraically dependent over K(£l9ijZ9...,¾). By hypothesis, any element £ of K* is algebraically dependent over K(ijlf7i2f ...,¾). But we now know that ijv tj2, ..., rj8 are algebraic over K(£l91/2,---, Vs)- Hence, by Th. IV, we deduce that £ is algebraically dependent over #(£1,^2,...,¾). This proves the theorem when ¢=1. We now proceed by induction on t. It is clear from the definition of algebraic dependence that if £i> •••,£< are algebraically independent over K, any subset is also algebraically independent. In particular, £^..., £^ are independent. Hence, if we assume that our theorem is true for t — 1 elements £,, we have the relation A _ ^ t-1^8, and we may suppose that the ijv ..., rj8 are arranged so that the set (£1,...,^-1,^,...,¾) is an algebraic basis for K* over K. Since £, is in K*> it is algebraic over J^(£i, ...,£/_i,%, ...,¾). That is, there exists a non-zero polynomial/^, ...,x8+1) such that /(£i,...,^-i,^,...,^a,£/) = 0. If t — 1 = s, it would follow that £!,...,£* are algebraically dependent, contrary to hypothesis. Hence t—1<5, that is t^8. Again, one at least of the indeterminates xt, ...,xs must be present hi f(xi,.-., xa+i)> otherwise £x,..., £, would be algebraically dependent over K. We may suppose that the rii are rearranged so that xt is present. Then it follows that ijt is algebraically dependent over #(£1, -..,£<-i,£/,Vt+i> ..-,^). By hypothesis, any element £ of K* is algebraically dependent over #(£1, • • •, £<-i, Vt> • • •, Vs)- But we now know that £1» • • •, &-i. Vt>—>V* are algebraic over #(£1,...,^,^+1,...,¾). Hence, by Th. IV, we deduce that £ is algebraically dependent over K(£x,..., £„ ^/+1, ...,¾). Hence (£1,...,^,^+1,...,¾) is an algebraic basis for if*, and the theorem is proved. We deduce from the first part of this theorem that any set of t elements, where t>8, must be algebraically dependent. Now
110 HI. ALGEBRAIC DEPENDENCE suppose that the set (ijl9 ..., rj8) is a minimal basis for K*9 and let (£i> •••>£«') be another minimal basis. Then s' ^ s and s < a', and therefore 5 = 5'. Therefore the number of elements in a minimal algebraic basis for K* with respect to K is independent of the minimal basis chosen. We call this number the dimension of K* with respect to K. Theorem VI. If £v ...,£r is a minimal algebraic basis for an extension K* of K9 and £ is any element of K*, there exists a non-zero irreducible polynomial f(xl9..., xr+1) ^n K[xi> •••j^r+i] suc^ $**& /(^,...,^,0 = 0. If 9(xi> • • • 9 xr+i) i8 any other polynomial such that (7(^,...,^,0 = 0, then g(xl9..., xr+1) = f(xl9..., xr+1) h(xl9..., xr+1)9 where h(xl9...9xr+1) ^s ^n K[xi> •••j^r+J- Since £ is algebraic over K(£l9...9£r) there exists a non-zero irreducible polynomial/^,...,#r+1) in K[xv ...,#r+1] such that /&,...,&,£) = <>. Now let g(xv ...,xr+1) be any polynomial such that <7(£i,...X,0 = o. Since f(xl9...,#r+1), regarded as a polynomial in K(xl9 ...,xr) [#r+i]> is irreducible, g(xl9 ..., #r+1), likewise regarded as a polynomial in K(xl9...,xr)[xr+1]9 either contains/(^,..., xr+1) as a factor, or the two polynomials are relatively prime. Case (i). Suppose that y v /./ \^\xl9 •••*Xr+l) g(xl9...,xr+1) —j[xl9...,xr+1) -^- ^ . Then h(xl9 ...,xr)g(xl9 ...,zr+i) =f(xl9 ...,xr+i)Hxi> — »av+i). Since K[xl9...,xr+1] is a unique factorisation domain, and f(xv ...,#r+1) is irreducible, and contains terms in &r+1, the polynomial k(xl9...,xr) must be a factor of h(xl9 ..., #r+1). Hence, after division by k(xv ..., xr)9 g(xl9...,xr+1) = f(xl9...,xr+1)h'(xl9...,xr+t).
3. EXTENSIONS OP FINITE DEGREE 111 Case (ii). Iff(xl9 ..., #r+1) and g(xv ..., #r+1) are relatively prime as polynomials in K(xv • ..,&,.) frr+iL there exist elements a(xv ...,xr+1) and b(xlf...,xr+1) of K(xl9 ...,xr) [xr+1] such that a(xl9 ...9xrJhl)f(xl9 ...,xr+1) + b(xl9 ...,xr+1)g(xx, ..., zr+1) = 1. Let ^-^1)-1¾^¾ «d 6<* *»>- fe^y- where -4(&i, ...,&r+i) an(i ^(#i> •--^+1) lie ^11 ^T#i>---^r+iL and 41(a;1,..., xr) and .6^:½..., xr) lie in K[xv ..., xr]. Then ^ll^l* • • • y xr) A (Xl9 ..., #,.+1)/(^1, •.., #r+l) + 41(iC1,..., xr) B(xl9..., xr+1) g(xl9..., #r+i) = A^x^y..., #r) ^o1(ic1,..., xr)9 and therefore 4^,...,^)1^,...,^) -54»,&)i(Si,..X0/(Si,»X0 + A1(Zv...9£,)B(Zv...9L9Qg(tv...9£„Q = 0. Hence §!,...,§,. are algebraically dependent over K9 contrary to hypothesis. The only possible case is therefore the first, and the theorem is therefore established. 4. Factorisation of polynomials. Let f(x) be any irreducible polynomial in K[x], and let us use it, as in § 1, to construct an algebraic extension K(£) of K. Now consider f(x) as a polynomial in K(£) [x]. By the Remainder Theorem [I, § 6], /(*)M*-£)A(*), since /(£) = 0, f^x) lying in K{£) [x]. Thus, when we extend K to K(£)9 f(x) becomes reducible. More generally, let F(x) be any polynomial of degree n in K[x], and suppose that F(x) = Fl(x)F2(x)...Fk(x)9 where the factor Ft(x) is irreducible and of degree n{ > 0. Then k 1
112 m. ALGEBRAIC DEPENDENCE If, for some i, n< > 1, we can construct an algebraic extension of K in which F^x) is reducible; that is, we can find an algebraic extension R' of K over which F(x) has at least k+l factors. Proceeding in this way, we reach a field K*, which is an algebraic extension of K, over which F(x) is equal to the product of n factors, each of the first degree. When a polynomial with coefficients in a given field can be expressed as the product of linear factors, it is said to be completely reducible over the field. Hence we have Theorem I. Given any polynomial f(x) in K[x], there exists an algebraic extension of K over which f(x) is completely reducible. This theorem takes the place in our present work of the ' Fundamental Theorem of Algebra', which asserts that when K is the field of complex numbers every polynomial in K[x] is completely reducible over K. A field K which is such that every polynomial in K[x] is completely reducible over K is said to be algebraically closed. The field of complex numbers is one example of such a field. In later chapters we shall often begin by considering a field K which is algebraically closed. But we cannot restrict ourselves to such fields, for the adjunction of an indeterminate to K gives a field which is not algebraically closed. For instance, if K is the field of complex numbers, and x, y are indeterminates, x2 + y2—l9 regarded as a polynomial in y, is irreducible over K(x). We shall therefore investigate fields which are not algebraically closed, merely stating the important Theorem II. Given any field K there exists an algebraic extension K* of K which is algebraically closed. K* is uniquely defined by K to within equivalent extensions. This field K* is called the algebraic closure ofK. We shall not use this theorem, and therefore omit the proof, which is difficult. Let us return to the commutative field K in which the polynomial F(z) = Fx(x)...Fk(z) is of degree n, the F^x) being irreducible polynomials in K[x], Let K% and JfiT* he any extensions of K over each of which F(x) is completely reducible. If F(x) = al\ (a-£<), and F(x) = a U {x - Vi) are the respective factorisations of F(x) in Kf, K%, we examine the
4. FACTORISATION OF POLYNOMIALS 113 relation between the extensions K(l;l9...9£n) and K(ijl9..., rjn). These fields are subfields of Kf and K£ respectively. Since we shall not go outside these subfields, we may suppose that Z* = Z(^,...,U and K* = K(Vv...iVn). Then Kf and K$ are both algebraic extensions of K [§3, Th. II]. We prove that they are equivalent extensions of K. The proof is by induction on n. If n = 1 there is nothing to prove, since K% = K = K%. We therefore assume that the theorem is true for polynomials F(x) of degree less than n. If the polynomial we are considering has a linear factor x — a in K[x], K(a, £2,..., £n) = K(£2,..., £n), and the theorem follows by the induction hypothesis. We now assume that no irreducible factor F^x) of F(x) is of degree 1, so that none of §!,..., £n and ^,..., r/n is in J5T. Since Kf is a field, Kf [a;] is a unique factorisation domain, and hence the factors of F±(x) are certain of the factors x — gi# By renaming the factors, if necessary, we may assume that x — £± is a factor of jFi(#), so that 2^(^) = 0. Similarly, we may suppose that J^O/x) = 0. By § 2, Th. II, the fields if (£x) and #(^) are equivalent. It follows that the rings K(^) [x] and K(^) [#] are equivalent, so that if is the factorisation of F{x) in 1^(^) [x], the polynomials Ff'^x) being irreducible polynomials with coefficients in K^), then also ^)=:^(^)...^(0:) is an equivalent factorisation of F(x) in K(^) [a], the polynomial F^x)(x) being the isomorph of Ffl)(x). But i^(o;) has now a linear factor # -^ in 1^(^)[&], ora;-^ in #(^)[#], so that, by the induction hypothesis, we may assume that the extension obtained by adjoining £2,..., £n to K{ix) is isomorphic with the extension obtained by adjoining 9/2,..., ?/n to lf(?/i), the isomorphism being such that any element of K(£±) is the isomorph of the corresponding element of the equivalent field K(rj^). Hence, K(£v ..., £J and K(tjv ...,7jn) are equivalent extensions of K. If ^(#) is any polynomial of degree n in !£"[#], and, if over some extension of K n W-«11 (*-&), HPI 8
114 HI. ALGEBRAIC DEPENDENCE the field K(£v ..., £n) is called a root field of F(x). We have therefore proved Theorem III. The root field of a polynomial F(x) in K[x] is uniquely defined to within equivalence over K. We may use these theorems to illustrate an important observation, namely that, given two extensions Kx and K2 of a field K, there may be no extension K* of K which contains Kx and K2 as subfields. To see this, let F(x) be an irreducible polynomial over K, of degree n(n> 1), and take Kx and K2 to be equivalent, but not identical, root fields of F(x) such that Kx and K2 intersect only in K. Suppose that gx,...,£n are the roots of F(x) in Kl9 and that r/l9...,r/n are its roots in K2. Then ^4=^ for any choice of i> j. If J5T* were a field containing both Kx and K2i we should have, over X*, n F(x) = aU (*-&), and J^(^) = 0. Hence II (^-^) = 0, and therefore r\i = gA for some choice of k, contrary to hypothesis. If F(x) is any non-zero polynomial in K[x]9 any element £ in K, or in an extension of Ky such that F{E) = 0, is called a root or zero of -F(#). We now prove Theorem IV. A necessary and sufficient condition that two poly- nomials in K[x] should have a common root is that their highest common factor should be of degree greater than zero. Let F(x) and G(x) be the polynomials, d(x) their highest common factor. Then there exist polynomials a(x), b(x),f(x), g{x)y all in K[x], such that ^x) F^ + b(x) Q(x) = d^ (1) and F(x) = d(x)f(x), 0(x) = d(x) g(x). (2) Let If* be an extension of K such that F(x) and G(x) are completely reducible in K*[x], Then d(x) is also completely reducible. If £ is a common root of F(x) and 0(x) then, from (1), £ is a root of d(x), and if £ is a root of d(x) we see, from (2), that £ is a root both of F(x) and 0(x). This proves the theorem. We note that the highest common factor of F(x) and 0(x) can be found by operations in K[x] [I, § 6], but the common root of F(x)
4. FACTORISATION OP POLYNOMIALS 115 and 0(x), if there is one, may lie in some algebraic extension of K, defined by d(x). The theory of root fields is a large one, but most of it lies outside the scope of this book. We conclude this section with an account of the symmetric functions of the roots of a polynomial. Let xl9...yxn be n independent indeterminates over K9 and let c1 = x1+...+xn, C2 = #i#2 + #13¾ + • • • + ^n~l^n> • • • • • Cn ~ Xl • • • Xn' We call cvc2, ...,cn the elementary symmetric functions of xl9 ..., xn. They are evidently unchanged by the n! permutations of the indeterminates xv ...,xn. Any polynomial in K[xl9 ...9xn] is said to be a symmetric integral function ofxv ..., xn if it is unaltered by the n\ permutations of xl9 ...,xn. We prove that such a polynomial is expressible in terms of the elementary symmetric functions, or, more precisely, we prove Theorem V. If f(xv...,xn) is a symmetric integral function of xv ...,xn, it lies in K[cv ...,cJ. n We define the total degree of an expression x^xl*... xr* to be 2 ri9 i = l and the total degree oif(xv ...,xn) to be the maximum of the total degrees of its terms xrfxr2% ...xr£ for which the coefficient aTi fa... fn 4s 0. We prove the theorem by induction, first on n, then on the total degree of f{xv...f x J. In the case n = 1, c± = xv and hence K[x^\ == K[c±], and the theorem is trivial. On the other hand, if f(xv ..., xn) is of total degree 1, £l K ,, ,, Interchange xi and x^ Since f(xly..., xn) is symmetric, we must have ai = a^ We deduce that /(*!>•.M*n) =«(*!+-+^) + 6 = ac± + b. It therefore lies in K[cv ..., cn]. Having established our theorem when n = 1, we assume that it is true for symmetric integral functions of xl9...,xn_v It is true for symmetric integral functions of xl9..., xn of total degree 1; and so we assume it is true when the total 8-2
116 III. ALGEBRAIC DEPENDENCE degree is k-~ 1, and prove it is true for the symmetric polynomial f(xv..., xn) of total degree k. n 6 /(*i, ..., Xn) ==/o+/A+... +fkxt where/0, ...,/^ are in K[xv ..., #n_i]. If we consider the permutations of xv*>xn which leave xn unchanged, f(xv ...,¾) is unchanged, and therefore /0,...,fk are unchanged by the permutations of %v • • • 9 xn-v I* follows that/0,... ,/j. are symmetric integral functions of xv ...9xn_v By the induction hypothesis each of these polynomials, and in particular/^ is expressible in terms of the elementary symmetric functions of xv ..., xn_v Writing d0= 1, d1 = x1 + ... + %n-v #2 = X± #2 "I" %i %3 ~^" • • • "I" xn—2 Xn—1> we have /0 = ^(d^,..., dn^)9 and we note that ct = di + Xnd^ (i = 1,...,n). (3) Now /(»!,..., xn) = /0 + zn i^,..., xn), and, by (3), <f>o(Cl> • • • > <V-l) = <f>0(dl> • • • > dn-l) + *n&o(Xl> • • • > *n)- Therefore /(¾ ...,^)-^(¾ -»Vl) = ^[%. ...,^)-¾¾ -^n)]- (4) The total degree of <fi0(cv ...,cn-i)> regarded as polynomial in #i> • • • > ^n* iQ equal to that of$Q(dl9..., dn_1), which is that of/0. Hence, since this cannot exceed k, the total degree of the right-hand side of (4) cannot exceed k. Since the left-hand side of (4) is a symmetric integral function of xv...,xn9 the right-hand side must contain, besides xn9 the factor We can therefore write f(xv ...,xn)-fo(cv ...,cn_x) = x1...xnO(xv ...,xn) = cnO(xv...,xn),
4. FACTORISATION OF POLYNOMIALS 117 where 0(xv ...,xn) must be symmetric in xv ..., xn9 and is of total degree < k — n. By hypothesis, we can write and finally and thus/(^,... ,#n) lies in #[/^,...,cn]. This proves the theorem. This theorem is mainly used to connect the coefficients of a polynomial with the roots of the polynomial. We adjoin cv...,cn to K9 obtaining a field K* which is a subfield of K(xl9...,xn). We consider the polynomial in K*[x], f(x) = xn-c1xn-1 + c2xtl-2 -... + (- l)ncn. In an extension of K*[x] this equation can be written as f(x) = (x-xt) (x-x2)... (x-xn), and Th. V can be regarded as a property of the root field of/(#). Again, consider the polynomial F(x) = ocP-v1xn-1+... + (-l)nvn in K[x], and let §i> •••> §n be its roots, in some extension K* of K. Any element of the root field K(£v...9jzn) can be written as a polynomial in E,l9...9£,n with coefficients in K [§3, Th. II]. Let a(£v ...,£n) denote such a polynomial. It is said to be a symmetric function of the roots gv ...,£n if a^,..., #n) is a symmetric integral function of the indeterminates xv ...9xn. By Th. V, a(xl9...,xn) = b(cv...,cn)9 where b(cl9...9cn) is a polynomial with coefficients in K. Hence, by specialising xi-^^i (i = 1,..., n)9 we find that and we have proved Theorem VI. // jP(o;) ia any polynomial in K[x]9 any symmetric function of the roots of F(x) lies in K. On the other hand, it may happen that a{E,v ..., £n) is in K9 although a(xl9...,xn) is not symmetric. Theorem VII. If rj is any element in the root field of F(x)9 the characteristic polynomial of ij over K is completely reducible in the root field.
118 m. ALGEBRAIC DEPENDENCE As above, let £v ..., £n be the roots of F(x), and let us write wEere a(xl9..., xn) lies in K[xv ..., xn]. Now let a*{xv...9xn) (i = 1,...,n,!) denote the result of performing a permutation on the indeterminates xv...,xn in a(a;1,...,xn). Consider the polynomial n! 11^-^1,...,^)3 = ^^61^^+... + 6,,,. The coefl&cients bi are clearly symmetric integral functions of #1, • • •, #n> and therefore if ¢(#) lies in J5l[#]. But (j>{rj) = 0, and therefore the characteristic polynomial of tj is a factor of <f>(x). Since ¢(#) is completely reducible in the root field, so is the characteristic polynomial. We conclude this section with Theorem VIII. The elementary symmetric functions cl9...9cn of n independent indeterminates over K are themselves algebraically independent over K. Let the indeterminates be xv ...,xn. If n = 1, cx = xl9 and the result is evident. We prove the theorem by induction on n9 using the notation of Th. V. By hypothesis, then, dl9..., dn_± are algebraically independent over K. Now, if cv..., cn are algebraically dependent, there is a relation #a>...>0 = 0o + 0icn + -..+&<4 = 0, where the fa lie in K[cv ...,cn_1], and are not all zero. Putting xn = 0 in this identity, we find that <f>(dv ...,dn„v0) = 0o(clf ...,Cn-l)«^0 = 0o(^l> — i^n-i) = 0. By hypothesis, $0(dv ..., dn_1) = 0 implies that <fi0 is identically zero. Hence, if <j>{cl9 ...,cj = 0, we have #a>...,O = cnfr(cl9 ..., cj = 0, and therefore tlr{cly..., cn) = ^ +... + <ptcf~x = 0. The theorem now follows. For we prove, in succession, that 0i,02> •*•>& are aU identically zero.
5. DIFFERENTIATION OF POLYNOMIALS 119 5. Differentiation of polynomials. Let x be an indeterminate over Ky and let f(x) = a0 + axx + oc2x2 + ... + ocnxn be any polynomial in K[x]. Then, if h is any new indeterminate, f(x + h)-f(x) = a0-a0 + a1[(x + h)-x] + ...+aJ{x + h)n-xn] = ^1 + a2[x + h + x] + ...+01^ + 11,^-1 + x(x + h)n~2 +...+ a;71-1]}, and therefore f(x + h)—f(x) -„ v 7 , 7V - ^-^ =f'{x) + 1ig{x9h)9 where g(x, h) is a polynomial in K[x, h], and f'(x) = ax + 2oc2x + 3a3a;2 + ...+ na^"1. The polynomial/'(#) is called the derivative off(x) with respect to x, J 7/» and is often written -?-/(#) or -p. Let us consider in what circumstances f'(x) = 0. We must have 0^ = 0, 2a2 = 0, ..., nocn = 0. If K is without characteristic, this implies that ax = a2= ... = an = 0, and therefore f(x) = a0, and so lies in K. Conversely, if f(x) is in K we find that f'(x) = 0. If K has characteristic p, the equation f'(x) = 0 implies that the coefficient at = 0 only when i is not a multiple of p. In this case, then, f(x) lies in K[xp], Conversely, if f(x) is in K[xp], the derivative is zero. This difference between fields with and without characteristic has many important consequences, and will shortly lead us to impose the restriction that K be without characteristic. Let us first note some formal consequences of the law of differentiation. If / V /> /> . O ™ we see at once that U(*) +9(*)Y = («i + A) + 2(Oi+A)* + ...
120 HI. ALGEBRAIC DEPENDENCE and that = [<*0 A) + («1A + a0 fil) * + • • • + (<*r#> + a,-i A + • • • + ao A) *" + ...+ anfim**+»]' = (a^0 + a0ft) + ...+ r(ary?0 +...+ a0&) x'-1 + ... + (m + n) anpmxm+n-x = {ax + 2a2a; + ...+ wa^-1) (/?0 + /?xa; + ... + /?mz"») + (a0 + a^ + ... + anxn) {fit + 2fi2x +...+ mpmxm-x) = f'(x)g(x)+f(x)g'(x). As an immediate corollary, we have, for any positive integer r, {U(x)ry = rfr(x)[f(x)Y-\ We extend the definition of a derivative to rational functions of x as follows. so that yg(x)=f(x), we define t/', formally, by the equation f'(x) = y'g(x) + yg'(x), and therefore we have the equation „' = IM JMIW = f'(*)9(x)-f(x)g'(x) y g(x) [g(x)f [g{x)f In the particular case in which g(x) is a factor off(x), y is a polynomial ^(#), and we can verify that y' = h'{x). The operation of differentiation may be repeated any finite number of times. If f(8)(x) denotes the 5th derivative of f(x), the following results are immediate: (i) [fW(x)r=f(r+s)(x); (ii) [xn]W = n(n- 1)... (n-r + l)xn~r (r^n); (Hi) /W(») = 0 if/(#) is a polynomial of degree less than r; (iv) if K is of characteristic p, and r ^ #, /(r)(*) = 0. This follows from (ii), since, if r^p, one at least of the integers n,n— 1,...,n — r+1 is a multiple of p.
5. DIFFERENTIATION OF POLYNOMIALS 121 (v) if J5T* is an extension of K, and f(x) is in K[x], the derivative oif(x) in K*[x] is equal to its derivative in K[x], Iff(xv ...,xn) is a polynomial in K[xl9..., xn], we may consider it as a polynomial in K[xv ...,xr_vxr+l9...,xn] [xr], and so define its derivative with respect to xr. We call this the partial derivative with respect to xri and write it ^-f(xv...9xr). When r=l, Now =— (oafl... a;^... »J») = mrax^... ^¾^-½¾^... x%», and therefore, if r # 5, 32 3¾. 9¾ - (Mft... xrnn) = ^_ j~JL (<l ... aj.) J = mrm8ax?1... x^l^x?'"1 ... xfi^x^'"1... #™* and this relation is obviously true when r = s. Similarly, we show that 3_ dxm ki^1"*'"^)= ^.(4^-^)» and we can write either of these expressions as 33 dxrdx8dxt [axp...x%*]. Since the polynomial/^,...,xn) is a sum of terms like ax?l... x%», we obtain the commutative and associative laws 32 32 and 4(^^- -^) = 3^(^/(^ -^)
122 HI. ALGEBRAIC DEPENDENCE Our next result is valid without restriction only when the field K is without characteristic. As a preUminary we note that the usual elementary proof of the multinomial theorem (^+...+3^ = 27 ' tag...aff, ' 1 • • • •' n • where the summation is over all ordered partitions (rv ...,rn) of r, is valid over any field K without characteristic. Now let We can write this symbolically as ^5»/(«i.-.*») = (^9^+--+^^-) f(x»->xn)> We prove Taylor's Theorem. If f(xv...,xn) is any polynomial in K[xv...,xn] of total degree N, then N I f(x1 + y1,...,xn + yn)= S ^|^J»/(*i.—.*»). Let us first establish this theorem when n = 1. If /(«) = a0 + a1a;+... + aJva^r, then f{x+y) = 0^ + 0^(3:+2/) + ...+0^(3:+2/)^ = a0+ <*!«+...+ (½^ + y[ax + 2a2 a; + ... + NctN x"-1] + ^-[2.1a2 + 3.2a3x+...+N(N-l)aNxN-i] +W\NlaN .. , 1 df(x) 1 232/(a:) 1 „dNf(x) = §o^*/(*).
6. DIFFERENTIATION OF POLYNOMIALS 123 We now proceed by the method of induction, assuming that the result is true for any polynomial in less than n indeterminates. If ^-(^+--+^5¾^ we verify by direct calculation that =AC) ^¾¾ [Dv*sf{xi> --^- Now, by the induction hypothesis, N I f(xl + Vl> • • • i *n-l + Vn-V xn + 2/n) = S 77 Dyx/fal* • •> ^n-l> ^n + 2/n)> and, applying Taylor's Theorem for the case w = 1, we have N-a l yt ^yxJ\xV • "9Xn-Vxn + Vn) = 2j 7i 2/n ^Zd •^yxJ\xH '">xn)' Therefore tf 1 ^-« 1 d* The summations on the right are over a lattice of points on the boundary of and within an isosceles right-angled triangle of side N. We may exhaust all the points of the lattice by considering those on lines parallel to the hypotenuse. In other words, put s + t = r (0^<r), and let r vary from 0 to N. Then 'khMyk^'^ *-> N l r = 0' l by the result stated above. This proves the theorem. We now consider certain applications of the process of differentiation to the theory of the roots of a polynomial.
124 m. ALGEBRAIC DEPENDENCE Theorem I. Iff{x) is any polynomial in K[x] of degree n greater than zero, a necessary and sufficient condition that (in an extension K* of K in which it is completely reducible) f(x) should have less than n distinct roots is that the highest common factor of f(x) and f'(x) be of degree greater than zero, or else thatfr(x) = 0. The latter case only arises in the case of a field of finite characteristic. (i) The condition is necessary. Suppose that f(x) is completely reducible in K*[x] and that f(x) = a(x - gj-t...(»- £r)"r (m< > 0), where £x,..., £r are distinct. Then f'{x) = a £ m<(* - ^)-4...(«- UY*-1 •••(*- &)"*• If%, ...,m5> 1, ms+1, ...,mr = 1, then or else (x - £)1¾-1...(#- Q™*""1 is a common factor in !£*[&] oif(x) and /'(#). Since/(#) and/'(a;) are not relatively prime in !£*[#], they cannot be relatively prime in#[>][I,§6,Th. II]. (ii) The condition is sufficient. Consider, first, the case in which f'(x) is zero. Then, as we saw above, p. 119, K must be of finite characteristic, say py and/(#) is in K[xp]. Let where <f>(y) is in K\y\. There exists an extension K' of K such that <f>(y) is completely reducible in K'[y]\ r <f>(y) = aU(y + Vi) (rp = n). There also exists an extension K* of K' containing elements £,...,5, such that # = V< (i=l,...,r). Now, since jfif* is of characteristic p, a simple application of the binomial theorem shows that (x + £i)p = xp + £f = xp + V+
5. DIFFERENTIATION OF POLYNOMIALS 125 Hence in K*[x], r f(x) = all(xP + Vi) t-1 = a (ft (* + &)}*. Hence f(x) is completely reducible in K*[x] and has only r distinct roots, £x,...,£r. Next, suppose that f'{x) is not zero, and that f(x) and f\x) have a common factor g(x) of degree greater than zero. Since f'(x) is of degree n — 1, g(x) is of degree m<n. We have /(a) = 0r(a;)A(aO, and hence f'(x) = gr(a:) A'(oj) + g'{x) h(x). Since (/(a:) is a factor of /'(#) it is a factor of !/'(#) ^(#), or elge this product is zero. Since h(x) is not zero, the product can only vanish if g'(x) is zero. The proof given above shows that g(x) is then the pth power of a polynomial in some extension K'[x] of K[x], and hence in this extension f(x) has a multiple factor. On the other hand, if g'(x) is not zero and g(x) is a factor of gf(x)h(x)i g(x) and h(x) must have a common factor, since g'(x) is of lower degree than g(x). Hence once again f(x) contains a multiple factor. Hence in both cases we have, for a suitable extension K', /(*) = n[/i(*)n where some mi > 1. Making a further extension K* of K' so that each/i(a;) is completely reducible in K*[x], we obtain the required theorem. We conclude this section with Theorem II. If K is without characteristic, the roots of an irreducible polynomial f(x) are all distinct. In the root field we may write f(x) = aU(x-U 1-1 and we must prove that E,v ..., £n are all distinct. If they are not, the previous theorem tells us that f(x) and f'{x) have a common factor. Since/'(#) is of lower degree than/(#), and f(x) is irreducible, there is no coitnmon factor. K being without characteristic, we have excluded the possibility oif'{x) being identically zero.
126 in. ALGEBRAIC DEPENDENCE 6. Primitive elements of algebraic extensions. In § 3, Th. I, we saw that any extension of K of finite degree was an algebraic extension of K. On the other hand, we saw in § 1 that a simple algebraic extension of K is an extension of finite degree. We now strengthen this relationship between extensions of finite degree and simple algebraic extensions by proving Theorem I. If K is a commutative field without characteristic, any extension J5l* of K which is of finite degree is a simple algebraic extension. Since K* is an extension of finite degree we can write K* = K(Tl9...,Tn)9 where rl9...9Tn are algebraic over K. We prove the theorem by induction on n. When n = 1, K{r^) is the algebraic extension of K defined by the characteristic polynomial of r± [§2, Th. I]. We therefore assume that the theorem is true for K(rl9 ...,rn_^). By hypothesis, there exists an element £ of K(t19 ..., 7,^) such that K(Q = 1^,...,7^). We must show that !£*(£, 7n) is a simple algebraic extension of K. Thus we need only consider the case n = 2; that is, we need only prove the theorem for the field K(^9r/)9 and the theorem will then be proved by induction. Let f(x) and g(x) be the characteristic polynomials of £ and y with respect to K9 and suppose that in the extension K' of K f(x) = (s-gi)... (z-£J, g{x) = (a-flx)... (x-yj, where we may suppose that £ = £x and r/ = ijv The result we are trying to prove is trivial if either n or m is equal to 1, so we assume that n > 1, and m > 1. Since K is without characteristic and f(x)9 g(x) are both irreducible, the roots £v ..., gn are distinct, and so are the roots Vv^Vm [§ 5, Th. II]. If k 4= 1, then rjk =J= r)9 and therefore the equation £* + *% = £ + *?/ (1) has, for all values of i9 a unique solution x = cik in K'. The field K is without characteristic, and therefore contains an infinite number
6. PRIMITIVE ELEMENTS OF ALGEBRAIC EXTENSIONS 127 of distinct elements. We can therefore find in K an element c different from any of the solutions cik (i= l,...,n; k = 2,...,m) of the equation (1) which happen to be in K. For this value of c, £i + cVk * £ + cV (i = !> •••>**; & = 2> ...,^). Let 0 be the element in K(£, 9/) defined by 0 = £ + q/. Since 0 is in 2£(£, ?/), !£(#) lies in lf(£, rj). We prove, conversely, that K(£, ?/) lies in if(0), and therefore that the two fields are the same. We know that , x 9(V) = 0, and /(^)=/(0-^) = 0. Writing f(6-cx) = h{x), the polynomial ^(#) Ues in K{6) [x]. Also, the polynomials h(x) and </(#) have the common factor x — i) in -£'[&]. This is also the highest common factor, for if x — rjk (k > 1) were a common factor, %*)=/(0-<%) = o, and for some value of i we should have 0-cyk = £i> and therefore 0 = ^ + cr/k, contrary to the definition of c. The polynomials g(x), h(x) oiK(6) [x] therefore have x — rj as their highest common factor in K'[x]. Their highest common factor in K(6) [x] is therefore ax + b = a(x — r/) (a and b in K(6)). It follows that rj is in K(0). But £ = 0-<"/> and therefore £ is also in K(6). Hence K(^r/) is in K(6), and the theorem follows. An element £ of an extension K* of If is called a primitive element of K if 2£* = If (£). Any extension of finite degree of a field without characteristic therefore contains primitive elements.
128 HI. ALGEBRAIC DEPENDENCE 7. Differentiation of algebraic functions. In § 5 we defined the partial derivatives of a polynomial in K[xv ...yxn] with respect to xl9 .♦.,xn. Assuming that the ground field K is without characteristic we now show how to extend the definition so that we can define the partial derivatives of an algebraic function rj of xl9...,xn. We shall then obtain an important criterion for the algebraic independence of a given set of elements of a given function field. Let JSl* be an algebraic extension of K[xl9...,xn], and let rj be any element of it. As in § 3, r/ is said to be an algebraic function of the indeterminates xv...9xn. By § 3, Th. VI, there exists an irreducible polynomialf(xv...,xn,y), unique save for a non-zero factor in K, such that », x /(^,...,^,^) = 0. Any polynomial g(xv ..., xn, y) such that g(xl9...9xn,ri)=rO contains/(&!,..., xn,y) as a factor. Now, xl9 ...,xn are algebraically independent over K. The polynomial f(xl9..., xn,y) therefore contains y, and, since K is without characteristic, We write 3 *, v for the result of replacing y by i) in w-f(xl9...,xn,y). This latter polynomial cannot contain f(xl9..., xn, y) as a factor, and therefore With these prehminaries, we define the partial derivatives drijdxi by means of the equations ^-/(^,...,^,7/)^ + ^-/(^,...,^,^) = 0 (» = 1,...,n). We note that if F(xv ...,xn,y) is any polynomial such that F(xl}...,xn,V) = 0, then F{xv ...,xn,y) = 0(xv ...,xn,y)f(x1,...,xn,y),
7. DIFFERENTIATION OF ALGEBRAIC FUNCTIONS 129 and therefore = /(*!> -,xn,y) |^0{xlt...,xn,y)^ + ^- G(xv ...,xn,y)\ + 0{x1,...,xn,y)^d-f(xl,...,xn,y)^ + ^f(x1,...,xn,y)]. It follows that ^¾....¾^)^¾¾....¾^^. In the particular case in which 7] lies in #(^,..., xn), our definition of 3^//3¾ coincides with that given in § 5. For if g(Sl,-.-,*n) h(xlf...,zn)' where both g(xv ...,a?n) and /1(0^,...,a?n) are polynomials without common factor, the characteristic polynomial of rj is f(x1,...,xn,y) = h(x1,...,xn)y-g(xl9...,xn), and our present definition leads to g|- = I^,...,0^—0(0;^ Since the dimension of i£* is n, a minimal basis is formed by any n algebraically independent elements of K*. Let £lf...,£n be 7& such elements. Then, by § 3, Th. VI, there exists a unique irreducible polynomial F(xl9...,xn,y) such that ^,...,^) = 0. As above, we show that ^(^,...,^) + 0, and define 3?//3£t- by the equations We wish to estabhsh a relation between the partial derivatives dr/jdxi and 3^//3^, but we must first prove a lemma.
130 m. ALGEBRAIC DEPENDENCE Let £l9..., £r be r elements of an algebraic extension of a field K without characteristic, and letf^x) be the characteristic polynomial of & with respect to K. We prove that If F(xv ...,xr)is any polynomial in K[xv ...yxr] such that 2^,...,^) = 0, then there exist polynomials A(xly..., xr), B^x^ ...yxr) in K[xly...,xr], where A(£ly..., £r) 4= 0, such that r A(xly ...,xr)F(xv ...,o;r) = S B^x^ ...,zr)/t(^). i-i This result has been proved already when r = 1, being implicit in §2, Th. I, and we proceed by induction, assuming its truth when r = s — 1. In place of K, we consider the field K(£,8), which is a simple algebraic extension of K. Over this new field the characteristic polynomials f^x) (i = 1,...,s— 1) of £lf...,g9_x may be reducible. But since 2£(£8), with K, is without characteristic, only one irreducible factor of each characteristic polynomial f^x) can become zero for the specialisation o;->^. We write where <f>i(x8,x) and ^(o;8,o;) are polynomials in K[x8,x], <f>i(£,8,x) is irreducible in K(g8) [x]y and &(£*>£*) = o, Mk£<)*o (i-i,...,*-i). By hypothesis, there exist polynomials a(xv ..., x8_ly £8) and 6^, ..., x8_ly £8) in K{Q [xly..., z8_i] such that a(£l9..., £8_x, £J 4= 0, and 8-1 ^i,...,Vi)W%-»vi»y = SM«i.•••>xs-i>£8)$i(£8>xi)- (i) ^=i Since K(£8) is a simple algebraic extension of K we may assume [§ 1, (1)] that a(xv ...yx8) and b^x^ ...yx8) are in K[xly ...yx8]. Now multiply (1) by appropriate factors, so that each term &(£8,#t) is replaced by/^). Then (1) becomes { II lM£* *<)l a(xi> • • • > *s-i> &) ^(*i> • •>*#-i. £.) = S%,..,Vi,Q/i(4 (2) where i-1 Bi(xly..., a^, £,) = bi{xl9..., ^, £8) n frj(£8> xj) i (j = 1,..., i-i.i + i, ...y8-l).
7. DIFFERENTIATION OF ALGEBRAIC FUNCTIONS 131 If we write 8-i A(xly...,x8) = a(xl9 ., x8) n ifa(x8>xi)> we have = nV«(k&) + 0. Now, the polynomial 8-1 .4(^,...,^)^(^,... S^<(*i,...,*•)/<(*<) (3) vanishes (by (2)) for the specialisation This must be true for the coefficient of each power-product in xl9 ...yx8__v Hence, the coefficient of each power-product must be of the form 0(x8) f8(x8). Therefore the polynomial (3) must be equal to B8(xv ...9x8)f8{x8)9 where B8(xv ...9x8) lies in K[xv ...,aj. Therefore a A(xv...,x8)F(xv...,x8) = %Bi(xl9...9x8)fi(xi)9 *=i and our lemma is proved. Now let rfv...,?inbe n elements of an algebraic extension K* of K(xv ...,xn) which are algebraically independent over K, and let £ be any other element of K*. By §3, Th. VI, there exist unique irreducible polynomials/^, ...,xn,z), g(yv ...,yn,z) such that /(^1,...,^0 = 0 and ^i,...,^n,0 = 0, and also unique irreducible polynomials fi(xl9 ...,#n,2/J such that fi(*i> ~-,xn>Vi) = 0 (» = 1, ...,n). In the lemma proved above, take the field K(xv ..., xn) in place of K, and the field K* in place of the extension field of K. Then we can find polynomials A(x,y,z) = A(xl9 ...9xn9 yv ...,yn, z), Bi(x,y9z) = Bfa, ...,xn, yv ...,«/n, z), C(x,y,z) = G{xv ...,xn, yv ...,«/n, z) such that A(x,y,z)g(yv...,yn,z) n = S -B«fo«/,2;)Uxv ...,xn,yt) + C(x,y,z) f(xv ...,xn,z), <-i where A(x,ri,£)^0. 9.2
132 m. ALGEBRAIC DEPENDENCE Differentiating this equation with respect to xi9 y^ z, and substituting ?fy for yp £ for z, we obtain the equations 0 = £ Bfa v, Q £-ft{xy Vt) + 0(x9 7/, £) £-f(z9 a W *-i -^./1^.71/ ■ ^X-'V.W^ 1 4(x, i/, £) g- 0fo, 0 = Bj(x, V, £) g-/,(«, V,), (5) A(x,V,Q^g(V,Q = C(x,V,Q^f(x,Q. (6) Since -4(#,?/,£)4:0 and ^(/(^,0 + 0, it follows that C(x, 7}, £) 4= 0. We know already that ^/(^, £) 4= 0. Now, since f(x, £) = 0, » 3/3 ^5[¾W¾*,•0][¾/^/¾/^,J]■ using (4), (6), and (5) respectively. But this last expression is t-id7itdx{' Hence we have proved Theorem I. If yv ..., r/n is any minimal basis for an algebraic extension K* of K(xl9..., xn), and £ is any element of K*, then ^ »9£9% Since the particular basis xv...,xn for K* has no special significance, this theorem gives a rule for determining the partial
7. DIFFERENTIATION OF ALGEBRAIC FUNCTIONS 133 derivatives of £ with respect to any base ijv ..., tyn when they are known for any other base xl9...,xn. In particular Hence we have a Corollary. The matrices are inverses of one another. The partial derivatives~(i = 1,...,n) are defined only when the minimal basis ijv ...>rjn is given. If r/v-'-yVr are anV elements of K* and £ is any other element of K* algebraically dependent on them, the irreducible polynomial/(^,...,«/r,z) such that may not be unique. If ^ = ~~f(y,z) \yimmVi is not zero, and if we define ^- formally by the equation the proof of Th. I, which does not make use of the fact that Vi> "-yVn are algebraically independent, shows that dx{ j-idrijdx^ This result has some importance in practical problems. Now let xl9...fxn be any minimal basis for an algebraic function field, and let r/v ..., rjr be r elements of the function field. The rxn matrix is called the Jacobian matrix of the functions /rjly...yrjr with respect to xv ...,xn. When r = n the determinant of the Jacobian matrix is called, simply, the Jacobian of ^, ...,7jn with respect to xv ...,xn, and is denoted by d(9l>->9n) d(xv...9xn)m
134 HI. ALGEBRAIC DEPENDENCE Theorem II. If (xv ...,xn) is a minimal basis for an algebraic function field If*, and ?li,...,Vn are n efewewte of K*, a necessary and sufficient condition for ?/i,...,?/n to be algebraically independent is d(xv...,xn) (i) Suppose that yi,...,yn are algebraically dependent. Let f(yv ..., yn) be a non-zero polynomial such that f(Vi> -»7«) = 0. At least one of the rjly...yrjn is algebraically dependent on the rest. After rearrangement, let this be ijn. Then we may assume that f{Vi> •••> Vn-v Vn) ^8 a non-zero irreducible polynomial in and since K(tjv ...,rjn^ is without characteristic, Also, f(r/v ..., Vn-v Vn) ^8 *^e zero °^ ^*> and therefore ^-/(^1, • • •» Vn) = ° (» - 1, • •., »)• But • VdJk_dl_ and, since not all 3//3^. are zero, it follows that det (^ = ^^ = 0. \dXfJ d(xu...,xn) (ii) Now suppose that riv ..., rfn are algebraically independent. Then we may take them to be a minimal basis for K*, and dxJdTfj is uniquely defined. By the corollary to Th. I, d(Vl>-,Vn) d(Xy.:,Xn) = j d{x1,...,xn)d{tIl,...,r}n) and therefore d(V»->Vn) + 0 d{xl,...,xn)
7. DIFFERENTIATION OF ALGEBRAIC FUNCTIONS 135 More generally, we can prove Theorem III. If (xl9...,xn) is a minimal basis for an algebraic function field K*> and 7jv ..., rjr are r elements of K*> a necessary and sufficient condition that yi,...,yr be algebraically independent is that the Jacobian matrix ofr/l9...9 rjr with respect to xl9..., xn be of rank r. Suppose that the Jacobian matrix is of rank r. By II, § 3, Th. IV, this matrix can be augmented to a regular nxn matrix, the elements in the part added being 8ipi (/> = ^+1,...,7^ = 1,...,^), where (ir+1,...,in) is a derangement of the set (r+1,...,n). Hence, by the above theorem, the elements Vl9--->Vr9 xir+x>'->xin are algebraically independent. Therefore the subset yi,...,yr ^8 a set of algebraically independent elements. To prove that the condition is necessary, we assume that ^,..., %■ are algebraically independent. Then r^n, and by the Exchange Theorem for algebraic dependence [§ 3, Th. V], we can find n — r of the x's, say , , x Vp = xiP (p = r+l,...,n), such that the set (ijl9...,rjn) is a minimal basis for K*. Then 9(^1,...,^) ' and the rows of the matrix must be linearly independent. In particular, the first r rows are linearly independent, and the rank of the Jacobian matrix of Vi>--->Vr wi*h respect to xv ...,xn is r. This completes the proof. Finally, we note the following theorem, which is an immediate consequence of the results proved in Th. I: Theorem IV. If (zv...9zn), (yl9...,yn)9 («i,...,««) are three minimal bases of an algebraic function field K*, the Jacobians satisfy the relations dfci* — >gn) = d(si>—>sn) d(yi>—>yn) d(xv...,xn) d(yl9...,yn) d(xl9...9xn)9 9(Zl,...,Zn)%l>--->2/n) -1.
136 HI. ALGEBRAIC DEPENDENCE 8. Some useful theorems. We conclude this chapter with the proof of some theorems which are of use in later chapters. Theorem I. If K is any commutative field containing an infinite number of elements, and if 4>i(xl9...,zn) (t = 1,..., r) are r non-zero polynomials in K[xv ...9xn], there exist elements cl9..., cn in K such that 0<(ci»—»Cfi)*° (* = !> •••>*•). We need only consider the case r = 1, for if r #(x19...,xn) = n&(si,...,*«)t then ^1,...,0 + 0 implies &(Ci,...,cn)*0 (* = ^---^)- We prove the theorem for a single polynomial ${xl9...9xn) by induction onw. If n = 1, let <f>(xx) be of degree m (m > 0), that is, <f>(x1) = d0x? + d1x?~1+...+dm (d0#=0). By § 4, there is an extension K* of K in which ^{xx) is completely reducible That is, 0(#l) = d0(*l-£i) — (*l-£m)> where £1? ...,£m are uniquely determined elements in K*. Since K^K*, and If contains an infinite number of elements, there is an element c± in K not equal to any one of the set £1?..., £m. Hence since cx # £*, and there are no divisors of zero. We now assume that the theorem is true when n — 1 indeterminates are adjoined to K. Let be a non-zero polynomial in K[xv ...,xn]. We may suppose that ^0(^1^---^71-1) + 0. By hypothesis we can therefore find elements cv ..., cn_! of if such that <f>0(cl9 ...,^^)4=0. Now apply the proof for the case n = 1 to the polynomial <f>(cv ...,cn_vxn), and the result follows.
8. SOME USEFUL THEOREMS 137 Corollary. The theorem is true if K is any commutative field without characteristic. For K then contains the field of rational numbers, and therefore has an infinite number of elements. Theorem II. If K is a commutative field without characteristic which is algebraically closed, <j){x) any polynomial in K[x] such that ¢(0) + 0, and if r is a given positive integer, then there exists a polynomial f(x) such that [f(x)]r — x is divisible by <f>(x). Since K is algebraically closed we can write 0(z) = an (*-&)•< (£,inZ), where ^,...,^ are assumed to be distinct. Let polynomials g^x),...,gk(x) be defined by the equations (l-r)(l-ar)...(l-ylfr) 1 ^sr'K-i)! ( w J' where f */r denotes an element of K such that (£J/r)r = ^. Such elements exist, since K is algebraically closed. By direct calculation it may be seen that [gt(x)]r—x is divisible by (x—E^fK Now write (j>{x) = (x-Qoifaix), so that <pi(x) does not contain x — E,t as a factor. By the division algorithm, there exists a pair of polynomials a^x), b^x) such that Then, using this identity, gjps) _ &(*K(«0&(*) , giWbjWix-gjY* ■ + ■ (j){x) 4>{x) 4>{x) We may write ft(«0«i(«0 9i(x)bj(x) gi(x)ai(x) = h(x)+Ai(x) {x-itfi tv ' {x-Zi)'*' Ai(x) being a polynomial of degree less than s€, and then 9i{*)b{(x) _ h Bi(x)
138 m. ALGEBRAIC DEPENDENCE Hence g{(x) = A4(x) fa{x) + Bt{x) (x - £<)% and since [gi{x)]r — x contains (x — gj)8* as a factor, it follows that [Ai(x)(f>i(x)]r — x contains (& — §<)** as a factor. Now let /(*)-2U<(*)&(*). Since Afflcfr^x) contains (x — £J8* as a factor for all J=H, and [Ai(x)<j!>i(x)]r — x contains (x — ^)s* as a factor, it follows that [f(x)]r — x contains (x — ^)s* as a factor. By symmetry, this holds for all values of i from 1 to ky and therefore [f(x)]r — x contains k n(#--£i)8' as a factor. But a is in K, and not zero. Therefore l [f(x)]r — x contains (j){x) as a factor. This proves the theorem. It will be noted that the degree off(x) is less than that of ¢(#). But f(x) is not uniquely defined by the requirements of the theorem. We could, for instance, take f(x) = 2^4^)^) + ^)^), where el9...,ek are k elements of K such that €5=1, and r/r(x) is any polynomial. .. *
* CHAPTER IV ALGEBRAIC EQUATIONS 1. Introduction. In this chapter we shall be dealing with polynomials in several indeterminates, with coefficients in a field K. This field will be assumed to be commutative, and to contain an infinite number of elements. Let f(xl9..., xn), g(xv...,xn),... be a set of polynomials in K[xv ..., xn]. A set of elements (£x,..., £n) of an extension If* of the ground field K is called a solution of the equations f(xv...,xn) = 0,\ 0(^,...,^) = o,r (1) if M»-,£n) = 0, It will be observed that xl9..., xn in (1) are not indeterminates, but unknowns, otherwise (1) would imply that f(xv..., xn),... were zero polynomials. Strictly speaking, we should use a different symbol for an unknown, but it will save a multiplication of our symbols if we use the same letters for unknowns and indeterminates. There will be no confusion provided that we observe the following conventions: if f(xl9...,xn) is written alone, it is a polynomial in the indeterminates xl9..., xn; if we write /(^...,^) = 0 the x ^,..., xn are unknowns; and if we wish to express the fact that the polynomial f(xl9...,xn) is the zero polynomial, that is, that f(xl9...,xn) is identically zero, we shall write /(*!>• •., *J=0. The main purpose of this chapter is to find criteria for the existence of a solution of a given set of equations. While we shall often discover solutions in a field JST* which is a transcendental extension
140 IV. ALGEBRAIC EQUATIONS of K9 our reasoning will show that if a set of equations with coefficients in a field K has a solution, the set necessarily has a solution in an algebraic extension of K. We shall be much concerned with homogeneous equations. The equations (1) are homogeneous equations if all the polynomials f{xl9...9xn)9 g(xl9 ..., xn)9 ... appearing in the equations are homogeneous. The polynomial f(xl9...,xn) is homogeneous of degree r if, for an indeterminate A, f(Axl9...,AzJ = fcf(xl9...,xn). All terms in f(xl9...,xn) then have the same total degree r. A homogeneous polynomial is usually called a, form. If (£1?..., £n) is any solution of a set of homogeneous equations, so is (A£1?..., A£n), where A lies in any extension of K. Thus (0,..., 0) is always a solution, but we shall regard this as the trivial solution. We shall find it convenient to consider two solutions (^,...,^), (Vi> • • • i Vn) which are non-trivial as equivalent solutions if there exists a a such that ,. /., x U = Mi (*=l,...,n). Any non-trivial solution of a set of homogeneous equations will be identified with all equivalent solutions. So far we have said nothing about the number of equations in a given set, and have not even assumed the set to be enumerable. We shall be concerned with the number of equations in our next section. We remark here that there are two ideals associated with a given set of equations. First, in the non-homogeneous case, let h(xl9..., xn)9 k(xl9..., xn) be two polynomials which vanish for every solution (glf..., £J of (1). Then ^,...,^) = ^,...,^) = 0. But if a(xv ..., xn) is any polynomial in K[xv ..., xn], the equations h(xlf ...,xn)-k(xl9 ...,xn) = 0 and a(xl9...9xn)h(xl9...9xn) = 0 are satisfied by (£l9..., £n), since *(gi,...,gj-*(gi,-,fn) = <>i and ^,...,^)^,...,^) = 0. Therefore the polynomials h(xl9...,xn), k(xv ...,xJ,... form an ideal [I, § 2]. Now consider the case in which the equations (1) are homogeneous. Let us write
1. INTRODUCTION 141 where hi(xv ...,xn) is homogeneous of degree i in xl9...,xn. Since the set (1) is a set of homogeneous equations, if (£lf...,£n) satisfies Hxi> -..»»tt) = °> so does (*£i> •••» Agtt). That is, MSi.---.^) + AAi(gi,...,^) + ...+A^(g1,...f^)-0. Since A may be taken as an indeterminate over K(£lt...,£n) it follows that *o(£i> '-U = *i(fi,..,£J = -. = M£i» ...>L) = o. Therefore every polynomial of the ideal is a sum of homogeneous polynomials, each of which belongs to the ideal. Again, suppose that we may write the equations (1) in the form A(xv->xn) = °> Mxl9...,xn) = 09 • • • Consider the polynomials 2 a{\xi> • • • j xn)fi\xly ...»xn)> i where the a^x^ ...,¾) are in K[xv ...,a;J, and the various summations are all over a finite number of values of i. It is clear that these polynomials form an ideal. Also, every polynomial of this ideal vanishes for every solution of the equations (1), and therefore belongs to the first ideal. In the case in which (1) consists of homogeneous equations we may write ai(xv...}xn) = aoi(xv...9xn) + ...+ar%i(xv...9xn)} where aji(xv ..., #n) is homogeneous of degree j, and hence S ai(xv ' • • > Xn)fi(xV ..-, Xn) =5 2 S aji\xV .-., Xn)fi\xV • • •, Xn)> 3 i where each term in the summation is homogeneous and belongs to the ideal. The two ideals we have associated with the set of equations (1) do not necessarily coincide. To take a very simple example, let the set (1) be simply 2 n X-± = u. Then the polynomial xx lies in the first ideal, but not in the second.
142 IV. ALGEBRAIC EQUATIONS 2. Hilbert's basis theorem. We now show that in considering sets of equations, homogeneous or not, it is suflficient to confine ourselves to sets containing a finite number of equations. We first consider non-homogeneous equations, and then deduce the result for homogeneous equations. Theorem I. In any non-vacuous ideal i in K[xl9...,xn] there exists a finite number of polynomials g${xl9..., xn) (j = 1,..., r) such that any polynomial F(xl9 ..., xn) in i can be written in the form r F(xl9...9xn) = ^lai(xv...9xn)gi(xv...9xn)9 where the ai(xv ..., xn) are in K[xv ..., xn]. Such a set of polynomials gi(xv ..., xn) is said to form a basis for the ideal i. It is evident, conversely, that if the gi{xl9 ..., xn) are in i, and the ai(xl9 ..., xn) in K[xv ..., xn], then F(xl9..., xn) is in i. We prove the theoreip. by induction on n. If n = 0, the result is trivial. In this case K[xv ...,xn] = K. If i contains only the zero of K there is nothing to prove. If i contains a non-zero element a of K it contains any other element /? of K, since /3 = (/¾¾-1) a. Hence, to prove the theorem, we need only take g± = a. Now suppose that the theorem has been proved for ideals in K[xv ...9xn_1], We write the elements of i in K[xv ...,xn] as polynomials in xn, with coefficients in K[xv ...,xnrm{\. We consider the coefficients of the highest powers of xn in the various polynomials, that is, the leading terms of each polynomial. We write F(Xl9 ..., Xn) =f1(Xv ..., Xn_1) X% +/2(^1, • • •, #n-l) xnT + • • • > and prove that the polynomials f1(xv ...,xn_1) form an ideal in K[xv ...,#n_i]- To prove this, let 0(xv...9xj^g1(xv...9xn^1)x^ + g2(xv...9xn_1)x^1 + ... be another polynomial in i, and suppose that a ^ /#. Then F(xv...9xn)-**-fiG(xlf...9xn) s lfl\xV • • • 9 xn-l) — ^1(^1» • • • > xn-l)J xn + • • • is a polynomial in i with leading coefficient fl(xl> • • • > #n-l) "" ^lv^l* • • • > ^n-l)*
2. hilbert's basis theorem 143 Also, if a(xv ..., #n_i) is any element of K[xv ...,a?n.J, then a(xv ...,xn_1)F{x1, ...,xn) = a(xv ...,^-1)/1(¾ ...,vK + - is a polynomial in i whose leading coefficient is a(xv ...,xn_1)f1 (xv ...,xn_1). Hence, from the definition of an ideal, we deduce that the leading coefficients of polynomials in i, together with zero, form an ideal in K[xv ...,xn__1]. By hypothesis, there exists in this ideal J a finite set of polynomials such that any element of J can be written as the sum S«<(*1, ••• . *n-l)&(*l> — »*n-l)> i = l where the ty(&i, ...,&n_i) are elements of K[xv ...,xn_1]. Since ^(^, ..., &n_i) is in j, there is a polynomial gi(xv ...,xn) of i which, considered as a polynomial in xn, has ¢^(¾...,xn_±) for its leading coefficient. Select polynomials gi(xv ...,xn) corresponding to the r0 values of i considered, and let Nt be the degree of g^x^ ..., xn) in*-Let N-max{Nv...9N,J. Then if F(xv ...,xn) =/(¾ ...,xn_x)»« + ... is any element of i, of degree oc^N,we can write To i-i and JFfci, ...,^)-1^(¾ ^^n-i)xn~Ni9i(^v ...>*n) i-i is a polynomial in i of degree less than a. Repeating this argument at most a — N further times, we show that there exist polynomials 6^(¾ ...,xn) such that F(x1,...,xn)-^lbi(xv...,xn)gi{x1,...,xn)^h(xl9...fxn) i»i is a polynomial in i of degree less than N. We therefore consider the polynomials in i of degree less than N
144 IV. ALGEBRAIC EQUATIONS in xn. We show that the coefficients of x*-1 in these polynomials form an ideal jt in K[xl9 ...,av-il- Let and fc(»lf ...,^) = ^(^, ...,^1)^ + - Then h(xl9...,xn)-k(xv...,xn) = [h1(xl9 ---yXn_1) — k1(xl9 ---9xn__1)]xn" + ..., and a(*i,...,*tt-i)£(^1,...,¾) = a(xl9 ...,^-1)/^1(^, '",xn_1)xn"~ + ... are polynomials in i of degree less than N9 and their leading coefficients are ^1\XV ...,^71-1)^^1(^1, .--,^n-l), a(xv ...,^71-1)^1(^1, ••., #n-i), where a(#i, ...,#n_i) is an element of K[xl9 ...9xn_1\. The leading coefficients therefore form an ideal jv which, by hypothesis, has a oasis j 1 v jl t \ Yr0+l\XV ---,^71-1), ••-, Yn\XV ---J^n-l)- If the polynomials of degree N - 1 whose leading coefficients are these <f>ro+i(xl9..., #n_i) are flVo+i^l' • • •, ^n) — Pro+i^l' '"> Xn-l) Xn~ + • • •, we can write any element of i of degree iV -1 in the form ^1,---,^) = ^1(^1,---,^-1)^^+--- ri~r0 = 2 ^0+^(^1^---^71-1)^0+1(^1^---^71-1)¾- +•••, i=»l and therefore n-r0 "(^1» • • •, xn) ~~ S ^ro+A^l' • • • * Xn-l) 9r0+i(XV • • •, Xn) i = l is an element of i of degree N — 2 at most. Repeating this process JV — 2 times at most, we obtain a finite number of polynomials gi(xl9 ...,¾) (i = 1, ...,5) in i such that for suitable a{(xl9 ...,¾) in JT[#i,..., xn] we find that ^(*li ...,*n)-2M*1> —,*n)0<(*i, --,¾) i«l is a polynomial of degree zero, and a further application of the hypothesis of induction shows that r F(xlf...fxn)sJ^ai(xl9...9xn)gi{xv...9xn). i«i This proves Hilbert's basis theorem.
2. hilbert's basis theorem 145 Now suppose that we are given an enumerable sequence of equations ^ ^ _ 0 y_ ^ } (1) We have seen that the polynomials Saj{xl9...,xn)fj{xl9...,xn)9 where the summation is over a finite number of values of j, form an ideal i. This has a finite basis consisting, say, of polynomials 9k(xv~,*n) (k = !> •••>*•). Since each of these polynomials lies in i, 9k\XV -)^)=¾¾ ---yxn)fj\xV •••j^n)? where the summation is over a finite number of values of j. There exists, therefore, a finite number s such that 8 9k\xV • • • > xn) = 2 akj\xV • • • 9 xn)fj\XV • • ■ ? xn) {* = 1, • • •, ?*)• i=l Now, any polynomial of the sequence (1), say ft(xv ..., a?n), is expressible in terms of the gk(xv ..., xn); that is, r ft(xl9 ...9xn) = ^btk(xv ...9xn)gk(xv ...}xn) k=l 8 == £%(#1> • • • 9 xn)fj(xl> "'9 xn)> (2) i-1 r where c(j{xl9 ...,xn) = £ btk(xv ...,xn)akj(xv ...,a;J. From (2) we see that in the sequence of equations (1) we may consider only the first s of them, the remaining polynomials fj(xl9..., xn) belonging to the ideal defined by the first s9 and vanishing at all common zeros of these s polynomials. The argument given above holds whether the set (1) is homogeneous or non-homogeneous. Another application of the basis theorem is to the ideal defined by polynomials which vanish for every solution of a given set of equations. Let 9j(xl9...9xn) (i=l,...,r) , be a basis for this ideal. Then (7^,...,^) = 0 (j =1,...,/-) (3) is satisfied by all the solutions of the given set of equations. If /(*l,".,*n) = <> HP I 10
146 IV. ALGEBRAIC EQUATIONS is any equation of the set, f(xv ...,xn) is in the ideal we are considering, and therefore r f(xV • • • > xn) ~ S aj(xi> • • • , xn) 9j(xv • • •, xn)> Hence every solution of the set (3) is a solution of the set of equations considered. The two sets of solutions therefore coincide, and we may replace the set of equations by the finite set (3). If the given set of equations is homogeneous we can replace it by a finite set of homogeneous equations. For we saw in § 1 that if we write where g^(xv ..., xn) is homogeneous and of degree i, the homogeneous polynomials (orforms) g^x^ ...,xn) belong to the ideal. Hence, the set of homogeneous equations can be replaced by the finite set of homogeneous equations 9{j(xi>~>>Xn) = ° (* = 1,...,ft,;.? = 1,..., r). 3. The resultant of two binary forms. The results of the previous section enable us to confine our study of systems of algebraic equations to finite systems. We proceed to investigate the conditions that a system of equations may have a solution. While we are ultimately concerned with equations whose coefficients he in the given ground field K, it is essential to consider, as part of our method, equations with coefficients which may belong to an extension K' (possibly transcendental) of K. If the coefficients of the various equations in the system are a,/?, ...,8, the elements of K' obtained by applying the operations of addition and multiplication to a,/#, ...,8 and to the elements of K lie in an integral domain which is a subring of K'. We call this subring of Kf the ring of the coefficients, and its quotient field is the field of the coefficients. An extreme case is that in which the coefficients are all independent indeterminates over K. The ring of the coefficients is then K[a,fi,..., 8], while the field of the coefficients is K(a,fi, ...,8), the field of rational functions of these indeterminates. The system of equations will be called a system with indeterminate coefficients in this case. The idea of the specialisation of polynomials has been described in I, §5. Let ^ ^ _ Q {i = lf_s) (1) be a set of equations with indeterminate coefficients. We obtain a
3. THE RESULTANT OF TWO BINARY FORMS 147 set of polynomials in the coefficients (that is, elements in the ring of coefficients) which has the following property: a necessary and sufficient condition that the set of 8 equations ft{*i,-,*n) = 0 (t=l,...,«) (2) obtained from (1) by the specialisations fi(xV •••fXn) -*fi (XV " • > xn) should have a solution is that the corresponding specialisations of the set of polynomials all vanish. As a first step in our reasoning, and as an illustration of the method, we consider the case of a pair of homogeneous equations in two unknowns, /(*o>*i) = °» 0(*o»*i) = °- In the case of a single non-homogeneous equation f(x) = 0 it has already been shown [III, §4] that there is an algebraic extension of the field of coefficients in which f(x) is completely reducible, so that f(x) = a{x-a1)...{x-am). UK* is any other extension of the field of coefficients in which f(x) is completely reducible, and if /(aOsa(*-£i)...(*--£m)> we know that K(£lf...,£m) and K(av ...,am) are equivalent extensions of K, in which ^ and at correspond if the factors are suitably arranged. For our present purpose we need not distinguish between ^ and 0¾. Hence we can say that the only solutions off(x) = 0 are al> "'fam' Iff(x0, x±) is a form, that is, a homogeneous polynomial of degree m, we can write f{x09x±) = a{/31x1 -a^o) ...{fimx1-amx0) in some algebraic extension of the field of coefficients, and the only solutions, in the homogeneous sense, oif(xQ9 xx) = 0 are (/?!,<*!), ..., (#n,aj. IO-2
148 IV. ALGEBRAIC EQUATIONS Now consider a pair of forms in x0, x^ (usually called binary forms) in which all the coefficients are indeterminates over K: Then f{x0, xj)=a0x?+axx? -i«0 + ... + amx%,\ 9(x0>xi) = Mi + hxi~lxo +...+bnx%. J (3) a0x^+n~1+... +amxl-^xf a0af+n~2x0 +...+ am^-2<+1 = Xi f{x0, Xj), =a?i~ x0f{x0, Xy), a0xfx^ +...+ am<+«-1 = x$-if(x0, xj, b^m+n-i + _ + 6na^-1«J s zT-igiXo, xx), b0x?+n~*x0 +... + bnxf-*xZ+i = x?-2x0g(x0, xx), b0xfx?-1+...+bnxF+«-1 Ejfi^*,). If J? is the determinant of the square matrix a„ 0 a0 «! a., n rows a„ 60 6X 0 bn &i h (4) }* m rows the multiplication of these m + n equations in order by the corresponding cofactors of the successive elements in a column of the matrix (4) gives us the identity itejzg = An(x0, xx)f(x^ xx) + Brs(x0, xx) g(x0, xx), (5) where Ar8(x0, xx) and Br8(x0, x±) are forms of degrees n — 1 and m — 1 respectively whose coefficients are in the ring K[a0,..., am9 60,..., 6n]. There is one such equation for each pair of values of r, s such that r + 8 = m + n— 1. In particular, jfcm+n-i = A(x09 xx)f{xQ9 xx) + B(x0, xx) g(x0, xx)9 (6) when r = 0, * = m + n— 1.
3. THE RESULTANT OF TWQ BINARY FORMS 149 Now consider any specialisation in K'[x0, zj of the forms (3), where K' is any extension of K. If, for this specialisation, R is not zero, the specialised equations f(xo>*i) = °> 9(xo>xi) = ° can have no solution. Suppose, indeed, that they have the solution (co> <a), in which c0, say, is not zero. From (6) we have the equation = 0, which is a contradiction. Suppose, on the other hand, that for some specialisation of the forms (3) the determinant R is zero. The rows of the matrix (4), considered as vectors, are then linearly dependent. Hence we can find elements ,, , a0f aj,..., an_j, c0, cl9..., cm_i in K'y not all zero, such that d0 u0 + d1u1+... +dn^1un_1-c0v0-c1v1-...-cm_1vm_1 = 0, where the vectors u0,..., un_v v0,..., vm_1 are the row-vectors of the matrix (4). We therefore have the set of equations dQa0-c0bQ = 0/| doai + diao-co6i_-ci6o = o I pj dn-lam-cm~lh = °J If d(x0, xx) = d0*?-i +...+ dn^x^\ and c(a;0, a^) = c0xf -1 + ...+ c^ag1"1, the equations (7) express the fact that d(xo> xi)f(xo> xi) = c(Xo> xi) 9(xo> *i)> (8) being merely the equations which express the equality on both sides of (8) of the coefficients of a;f+"-1,..., ag^-i. Now resolve both sides of (8) into irreducible factors in K'[x0, &J. By the unique factorisation theorem [I, § 7], any irreducible factor on one side must also appear on the other. Since d(x0, x±) is of degree n — 1, at most, one of the factors of gix^Xj) must be a factor of f(xQ9 Xy). Hence if R = 0,f(x0, x±) and g(x0, xx) have a common factor
160 IV. ALGEBRAIC EQUATIONS of degree greater than zero. Any solution of the equation obtained by equating this common factor to zero is a solution of the equations /(^0^1) = 0, 00^1) = 0. The common factor can be found in K'[x0,x^\ by the division algorithm. The common solution of the two equations is in an algebraic extension of K'. Calling R the resultant of the two binary forms (or equations), we have proved Theorem I. A necessary and sufficient condition that two homogeneous equations with coefficients in afield K' have a common solution is that their resultant vanishes. The common solution, when it exists, lies in an algebraic extension of K\ We may use this result to obtain a theorem for two polynomials in K'[x]. Let £l . L J f(x) = a0xm+...+am9 g(x) = b0x"+... + bn. Make these polynomials homogeneous by introducing a new indeterminate y\ we obtain the forms /(*, V) = a0x™+... + amy™f g(x,y) = b0xn+... + bnyn. Let R be the resultant of these two forms. If J?#= 0, the equations /(*) = 0, 0(*)-O (9) cannot have a solution. For if they had a solution x = c, the homogeneous equations f(x,y) = 0, g(x,y) = 0 would have the solution (c, 1). If R = 0, the homogeneous equations have a solution. Let this be (c, d). If d # 0, this solution is equivalent to (cd*1,1), and x = cd~x is a solution of (9). But this reasoning fails if d = 0. In this case substitution in f(x, y) and g(x, y) shows that aQ = 0, bQ = 0. The determinantal expression for R makes it evident that if a0 = b0 = 0, then R = 0. Hence we have Theorem II. A necessary condition for the existence of a solution of two non-homogeneous equations is R = 0. It is not a sufficient condition unless we impose the additional condition that the leading coefficients of the equations are not both zero.
3. THE RESULTANT OF TWO BINARY FORMS 151 We deduce from (6) that the resultant of two polynomials f(x) and g(x) can always be expressed in the form B = A{x)f(x) + B(x)g(x)9 (10) where both A (x) and B(x) are polynomials in x with coefficients in the ring of coefficients of/(#) and g(x). We conclude this section by indicating how an alternative approach to the foregoing argument can be extended to give a set of conditions to be satisfied when the binary forms f(x0, x±) and 9(xo> xi) i*1 (3) have a common factor which is at least of degree t. If this is the case, f(x0, x±) =Mx09 xx) d(x0, xx)9 and g(x09 xx) == gx{x^ xt) d(x0, xj, where ft(x09 xx) == d0x? ~l +...+ d^aft-*, gi\XQ9x1)^cQx1 +...+ cn_iX0 . Since gx(x0f xx)f(x09 x±) =fx{x09 x±) g(x0, x±), we have the equations c0aQ-d0b0 = 0,^ c0a1 + c1a0 — d0b1 — d1b0 = 0,| (ii) cn-tam~~ dm_tbn — O.J Hence the rows of the (m + n — 2t + 2) x (ra + n —1+1) matrix (12) m — t+1 rows are Unearly dependent. Its rank is therefore not greater than m + 7i — 2t+ 1. Conversely, if the rank of this matrix is less than m + n — 2t + 29 the rows are Unearly dependent, and we can find elements d09 ...9dm__49 c0,..., cn_j, not all zero, in the field of the
152 IV. ALGEBRAIC EQUATIONS coefl&cients such that (11) is true. Defining fx{xQ9 x±) and g^x^x^ as above, it follows that 9i(Xo> xi)f(xo> xi) =/i(*o> xi) 9(xo> xi)> and from the unique factorisation theorem we deduce that /(#<>> #1) and g\x0, x±) have a common factor which is of degree t at least. Hence Theorem III. A necessary and sufficient condition that f(x0, x±) and g(x0, xx) have in common a factor of degree t at least is that all the determinants ofm + n — 2t+2 rows and columns in (12) should be zero. 4. Some properties of the resultant. Let ocl9..., ocm, fiv ..., fin he m + n indeterminates over the ground field K, and let cl9...,cm, dlf..., dn be the elementary symmetric functions of these two sets. That is, let c1 = a1+... + aw, <*i = A+ — +Ap c2 = aia2+ ... +am^am, d2 = AA+-+LX <>m = axa2... am, dn = fi±fi% ...fin. Let a0, bQ be two further indeterminates, and write «i = (-1)^0¾ (•= *>—>™)» &<=(-l)'M< (*=l,...,n). If f(x) s a0#m + 0^-1 + ... + am, and gr(a;) = b0xn + 6^-1 + ... + bn9 then f(x) = a0(xm-c1xm-1 + c2xm-2+... + (- l)mcm) ^olK^-a,.), n and 0(aOs&oII(3-A)- Let i? be the resultant of/(#) and gr(a:). Then i? is an element of K[a0,...,am, 60,...,6J, and we express this by writing R = R(a0,...,am, 60,...,&n). The determinantal expression for the resultant shows that it is homogeneous, of degree n, in a0,..., am, and homogeneous, of degree
4. SOME PROPERTIES OF THE RESULTANT 153 ra, in 60,..., bn. For every term in R contains one and only one element from every row, and one and only one element from every column of the determinant. The principal diagonal of the determinant gives the term aj b% as one term in B. We may write, by the above remark, R(a09...9am9b09...9bn) = anb$R(l9-cl9...9±cm9 l,-dv ...,±dn). Now, R(l9 — cl9..., ±cm, 1, —dlf ...,±dn) is an element of K[ocv. ..,¾^ jJJ, and may therefore be regarded as a polynomial in ai9 with coefficients in K[av ...,0^, oci+l9..., ocm9 fil9...9 fin]. This polynomial vanishes for each of the specialisations since /(a;) and gr(a;) then have a common factor. Hence, since the fiv •••> fin are aU distinct, a factor of the polynomial is (*i-p1)...{*i-fin). This is true for each value of i9 and therefore B(l,-cl9...,±cm, l9-dl9...9±dn)9 regarded as a polynomial in K[al9..., am, /?x,...,/?n], contains as a factor; or JK(a as a factor. But and therefore that is, Similarly m n i «* 1 j -«1 'o> • • • i am> fy» • • • >6n) contains m m n n »(«<) = 6? n n(««-A), (1) (2) (3) since /(») = a0 n (*-a») = (- l)ma0 II («*-*)•
154 IV. ALGEBRAIC EQUATIONS Now (2) can be expanded as a homogeneous polynomial of degree m in the coefl&cients 60,..., bn ofg(x). The coefficient of each term is Oq x (a symmetric function of al9...,am). Hence [III, §4, Th. V] S is a form of degree m in bQ9...,bn, with coefl&cients which are polynomials in a09cl9 ...,cm. We deduce from (3) that S can also be expressed as a form of degree n in a0,..., am with coefficients which are polynomials in b09dl9...,dn. Comparing these two results, we see that S is an element of K[a0,...,am9 60,...,6J which is homogeneous of degree n in a0,...,am9 and homogeneous of degree m in 60,..., bn. It also follows from (2) that one term in S is a%b%. We have already noted that R(<*o>-a0cl9...9±a0cm9b09-b0dv...9±b0dn)9 regarded as a polynomial in a0, b09 al9...,am, fil9 ...,/?n, contains S as a factor. That is, B(a09-a0cl9...9±a0cm9b09-b0dl9...9±b0dn) = ST9 where T is a polynomial in a0, 60, alf..., am, fil9...90n which must be symmetric in al9..., am and in /?1?..., /?n. Hence T can be expressed as a polynomial in a0, 60, cl9..., cm, ^,..., dn. By comparing degrees on each side and remembering [III, §4, Th. VIII] that a0i bQ9 cl9..., cm, dl9...9dn are algebraically independent over K9 we see that T is in K. Hence, expressing S as a polynomial in aQ9...,am, &o> --^^, we have a relation B(a09al9 ...9am9b09 ...9bn) = ST between the at and bi which must be an identity. By considering the coefficient of a % b% on each side we find that T = 1. Hence m n m n i-l i=l i —1 i—1 Now consider any speciahsation of the coefl&cients off(x) and gr(#), and let the roots of the specialised polynomials be £x,..., £m, ^, ...9rjn. Then, since the speciahsations <*<-*£;> A->%» ^o-^Aq, 60-^/^0 imply the speciahsations <*>i->hi9 bj'+fip we obtain the equation m n -R(A0, ...,^, /*0,...,ftn) = a?/*? n n (&-**)•
4. SOME PROPERTIES OF THE RESULTANT 155 We show next that R is isobaric in the a, and bj9 of weight mn. In other words, if cafyal1 ...alfibl9... bft is any term of m n then 2 nr + S SJ8 = mn- 0 0 To prove this, consider the expressions for R given in (4). If in these we substitute ol\ = Aa,, fij = A/?^, where A is an indeterminate, we obtain a = ajfty n n k-#) = a-x&s* n n (««-#), that is, R = Amni?, where 5 is the resultant of the polynomials f(x) = a0 n (*-oJ) = a0 II («- Aa<), i i ^) = 60n(*-^) = 60n(*-w. i i But these polynomials may also be written in the form f(x) = a0xm + Aa^™'1 + \2a2xm-2 +...+ Amam, g(x) = b0xn +\b1xn~x +A2b2xn~2 +... + An6n, R = R(a0, \al9..., Amam, 60, Aft^ ..., Xnbn) = A^i?(a0,...,am,60,...,6J. This proves the theorem. In later applications we shall need to consider the case in which the coefficients a^bj are forms in an integral domain K[£l9 ..., £8], obtained by adjoining the indeterminates £l9..., £8 to K. We suppose that the degree of at ish + i, where h is constant, and that the degree of bj isk+j, where k is constant. Since R is homogeneous, of degree 7i, in the ai, and homogeneous, of degree m, in the 6^, the degree of a term u ,, u v% m n in 22 is Ti(h + t)it+^(k + t)jt m n m n ^h^it + kZJt+Xtit+ZtJt = hn + km + mn> R being isobaric, of weight mn. Hence R is a form in K[£l9...,§J of degree nh + mk + mn.
156 IV. ALGEBRAIC EQUATIONS 5. The resultant of a system of binary equations. We consider the system of r homogeneous equations /i(Vi) = 0 (i-=l,...,r), (1) and suppose, in the first place, that the coefficients are independent indeterminates over K. Let /^(¾ #i) be of degree mi, and let m = max [mv ...,mr]. Then if '} (i=l,...,r), , (i= 1,...,r), <l>r+i (*0. xi) = 6< xi-mifi(x0> xi)J where the ai9 b{ are new indeterminates, we consider the auxiliary system of equations &G*o>*i) = 0 (* = 1 2r). (2) Clearly if, for any specialisation of the coefficients in an extension K' of K, the system (1) has a solution, so has the system (2). Conversely, if the system (2) has the solution (c,d), so that Mm"m\AM) == 0 j then, since aiy bt are not zero, and c, d! are not both zero, it follows that fi(c,d) = 0 (i=l,...,r), and therefore the system (1) has a solution. Thus, in order to find conditions for a solution of (1) when the coefficients are specialised in K', we merely need to find the conditions for a solution of (2) under the same specialisations. We do this by combining the system (2) into two equations. Let ul9..., u2r, vv ..., v2r be 4r further indeterminates, and let 2r %«i)sS«#o.*i)) 1 2r i If R(u, v) is the resultant of this pair of binary forms <P(x0, xx) and ^(#0,2^), we first note that R(u, v) is a polynomial in ul9...,u2r, vv ..., v2r. The coefficient of is evidently of the form D8 = daa\^ ... 4'+j' 6j'+»+*'+i... 6J»h-^, where d, is a polynomial in the coefficients of the original forms
5. RESULTANT OF A SYSTEM OF BINARY EQUATIONS 157 fi(xo>xi)- These polynomials dl9..., dN constitute a resultant system of the set of homogeneous equations (1). If, for any specialisation of the coefficients, the system (1) has a solution, so has the system (2), and hence also the pair of equations ¢(^0,^) = 0, !P(»0,»i) = 0. Then B(u, v) = 0. This implies that every coefficient D8 vanishes, and therefore, since the ai9 bi are indeterminates, . dx = 0, ..., dN = 0. Conversely, if each d8 vanishes, the forms 0(xo,x1)f W(x09xx) have a common factor. This can be determined by the division algorithm [I, § 6], and it is therefore an element of K'ia^bpU^ix^x^. We may assume that the common factor is in K'[aiy bp u8, vt] [xQ, xx] and is not divisible by an element of K'[ai9 bp u8, vt]. Now, 0(xo, xx) does not contain vl9...9v2r, so that this common factor does not contain any Vp and W{x^ xx) does not contain ul9...9 u2r, and so the common factor is independent of the ut. Hence every common factor of 0(xo, xx) and ^(Xq, xx) is independent of ui9 Vp and must therefore be a common factor of the forms $<(*•> *i) (*' = 1,..., 2r). The system (2), and hence the system (1), therefore has a solution. From § 3, (5), we may write B(u, v)xr0x{ = ar8(xQ, xx)0(xO9 xx) + br8(x0, x±) W{xQ9 xx), (3) where r + s = 2m—1, and ar8(x0,xx), br8(xQ,xx) are forms in x0, xx whose coefficients are polynomials in the coefficients of 0(xo, xx) and tPfa^Xi). Remembering that 0(xO9x1) and W(xQ,xx) are hnear expressions in the ^(x^x^ and how these are obtained from the original forms fi(x0,xx), we obtain, on equating like powers of ui> vp as> fy on both sides of (3), r diXf0x{= 2 bf8tij(xQ9 xdfs{xi>9 xx). (4) As in the case of two forms, we can deduce from these results others applicable to a system of r equations in one unknown. As
158 IV. ALGEBRAIC EQUATIONS before, we make the equations homogeneous, and construct the resultant system. Then we show that for specialised coefficients the members of the resultant system vanish if the given equations have a common solution, but that if the resultant system vanishes the equations either have a common solution, or the leading coefficient in each of the equations vanishes. We have proved the following result: Given r polynomials fx(x),...,fr(x) in one indeterminate with indeterminate coefficients there exists a set dv ..., dN of polynomials in the coefficients with the property that for specialisations of the coefficients the conditions dx = 0,..., dN = 0 are necessary and sufficient to ensure that (i) the equations fx(x) = 0,...,fr(x) = 0 are soluble in some extension field; or (ii) the leading coefficients infx{x),..., fr(x) all vanish. Moreover, if the equations have a solution, they have a solution in an algebraic extension of the field of coefficients. From (4) above we derive important expressions for dv ...,dN. Putting s = 0, we find that i<=SMW (6) Finally, we extend the result proved at the end of § 4 for two forms /(#<),#!), g(x0i xx) whose coefficients are themselves forms in K[£i> • ••>£«]• We consider r forms /<(*o>*i) (*= 1.-»*0» and suppose that the coefficient of x\xjQ in fk{x^xx) is a form in ^[£o> •••>£«] °f degree hk+j. If h = max[^,..., Ar], then in the preparatory stage of finding a resultant system for the fi(x0, xx) we choose ai9 bi to be forms in £v ..., £8, with indeterminate coefficients, of degrees h — hi + m — mi and h — hi respectively. At this stage our orms are ^^ ^ _ a^-^f^, xx), The coefficient of x[+m"m^xjQ in <f>k(x0, xx) is a form of degree h +j in £1?..., £8. Hence we may apply the result of § 4 to i 2r and W(x09 xx) = £ vk$k(x09 xx). l
5. RESULTANT OF A SYSTEM OP BINARY EQUATIONS 159 The coefficient of u{1... vig v{1... vfc; in B(u, v), which is Dt = dtO%+h ... 0,*'+*' 6jr+t+^+l ... 6*"+J" is therefore a form of degree 2mh + m2 in £lf..., £8. Since the a^, fy are forms in §lf..., £8, it follows that the dt are forms in ^,..., £,. 6. The resultant system for a set of homogeneous equations in several unknowns. We now consider r homogeneous equations in n+1 unknowns », \ a /• i \ /i\ /^0.-.^) = 0 (* = 1,...,r), (1) again beginning with the case in which all the coefficients are in- determinates. Let £0, §x be two further indeterminates, and write ( (» = 0,...,71-1). xn =3 bl We then obtain from (1) a set of equations, homogeneous in £0, glf /<(£o*o. -.£0<-i.£i) = ° (» = !> -.r). (2) The coefficient of £\ £{ in fM0xf09..., g0«n-i> £i)is a form in #o> • • •.xn-i of degree j. If we construct the resultant system of the equations (2), using the result of the preceding section, we obtain a set of forms in x'Qy ...,x'n_l9 di(x'09...X-i) (% = i,...,jy), whose coefficients are in the ring of coefficients of (1), with the following properties: (i) <W = 5X,(£o> §i)/j(£o*o. •••> £o^4-1. £i) 3 for a suitable integer pi9 where the a#(£0> £i) are forms in £0, £x with coefficients in the ring which contains x'Q, ...yx'n_l9 K9 and the coefficients of the equations (1); (ii) if for the specialisations (x0y..., ^n-1) of (x'09..., x^), di(xQ,..., 2n_i) = 0 (* = 1, ...,#), the speciahsed equations (2) have a solution, and conversely. This result remains true for any specialisation of the coefficients of (1), the coefficients of the forms di(x'09...9xfn_ml) being speciahsed to correspond to the specialisation of (1). Now, if the equations (2) have the solutions (c, d) for the specialisations (x0y...,^i) of (x'0,...,<_!), evidently (cx0y ...ycxn__l9d) is a solution of the set (1). Conversely, suppose that the set (1) has
160 IV. ALGEBRAIC EQUATIONS the solution (x0>...,xn). If xQ,...,xn_x are not all zero, the set (2) has the solution (l,Sn) for the speciahsation (xQ9 ..., xn_^) of (x'0,...,x'n_x). If, on the other hand, x0 = ... = xn_x = 0, #n4=0, the equations (2) have the solution (0,xn) for any speciahsation of #o> -y^n-v Hence, in this case, ^(^,...,0:^) = 0 (i =1,...,#). We see then that a necessary and sufficient condition that the equations (1) have a solution is that the equations ^K,...,<_i) = 0 (i=l,...9N) have a solution. Since these are homogeneous equations, and xi = £0x'i9 this is simply the condition that the equations di(x0f...9xn^) = 0 («= 1,..., N) (3) have a solution. We see that if the equations (3) have a solution which is algebraic over the field of coefficients, so has the system (1). As a final preparatory step towards our first theorem, consider the reverse transformation connecting xt with x\y xi = xd«\\ (i = 0 n-l), £l = xn I and its effect on the identity ditft = S %(&» £i)M£ox'o> • > £o<-i> §i)- This becomes »*(#<) bo J--->^n-lb0 )bSl—S r^/ fj\x0> --">Xn)> J bO where biJ(x09...9xn) is a form in .r0,..., #n; the coefficients are polynomials in £0. On multiplying by a suitable power of £0 we obtain the equation di\xQ,..., #n_i) £01 = 2boiybij(xQ9...,xn)fj(x09 ...,#„), and on equating like powers of £0 we have, finally, di\X0, ..., #n_i) = 2 Gij\XQ, ..., %n)fj\%o> • • •, #n)' W We are now in a position to prove Theorem I. Let /,.(^,...,^) = 0 (t = l,...,r) (1) be a set of homogeneous equations with indeterminate coefficients, and let f<{x0,...,xa) = 0 (t = l,...,r) (5)
6. EQUATIONS IN SEVERAL UNKNOWNS 161 be the equations obtained from (I) by a given specialisation of the coefficients in (1). Then there exists a finite set of polynomials in the coefficients of (1), dv ...,^, with the properties r (i) ^^ = 2^(¾ -,^)/^(¾ — >*n) for some value of m, the coefficients of a#(#0> • • • > xn) being in the ring of coefficients of (I); and (ii) a necessary and sufficient condition that the equations (5) liave a solution in an algebraic extension of the ring of coefficients is that the specialisations of the polynomials di arising from the given specialisations of the coefficients should be zero. We have proved this theorem when n = 1 [§ 5] and we now proceed by induction. Beginning with the set of equations (1), we construct the set of forms d{(xQ9..., aV-i) which satisfy (4) and have the property that the equations (3) have a solution if and only if the equations (1) have a solution. This property remains true for any specialisation of the coefficients. Now consider a set of N homogeneous equations %-,Vi) = 0 (i=l,...,iV) (6) with indeterminate coefficients, Dt(x09 ...,xn_x) being of the same degree as d^x^ ..., #n_i). Applying the hypothesis of induction to (6), we construct the set Et(i = 1, ...,&), of polynomials in the coefficients which are such that E^^^A^Xq, ...,^^)2)^0,...,^^) (i = 1,...,¾ and which also have the property that a necessary and sufficient condition that any specialisation of (6) should have a solution is that the corresponding specialisations of the Ei should vanish. We can specialise the coefficients of (6) in two stages. First we specialise the coefficients of (6) so that Di(xQ9...9xn_1) becomes di(xQ9..., xn__t)9 and then we specialise the coefficients of d^(#0,... #n_i) by the specialisation which leads from (1) to (5). The first specialisation induces a specialisation Ei-^di (i =l,...,fc) of Ei9 and we have diX0 = 2j*%(#o> • • • > xn-l) dj\%Q, • • •, %n-i) = SS^^/ftKr'^J (7) by (4), where 2a^fyfc is a form in x09...,xn9 with coefficients in the ring of coefficients of (1). HP I II
162 IV. ALGEBRAIC EQUATIONS When we make the further specialisation which leads from (1) to (5) we induce a further specialisation of the d€. The vanishing of these is a necessary and sufficient condition for the existence of a solution of the specialised forms of (3) in an algebraic extension of the field of coefficients; that is, in an algebraic extension of the field of coefficients of (5). This, in turn, is a necessary and sufficient condition for the existence of a solution of (5) in an algebraic extension of the field of coefficients. Our result is therefore proved. Moreover, if the set (5) has any solution, it follows from (7) that the di are all zero, and hence that the set (5) has a solution in an algebraic extension of the field of coefficients of (5). 7. Non-homogeneous equations in several unknowns. We now consider how we may apply a process similar to that of § 6 to test for the existence of a solution of a set of non-homogeneous equation* ^,...,^),0 « _ 1 ,), (1) having coefficients in the ground field K. It will be recalled that the condition given in § 5 for the case n = 1 was necessary, but was only sufficient if we imposed the restriction that the actual degrees of the equations were equal to the assigned degrees. We make a similar restriction in the present case, but even when this is made we may not be able to proceed in the obvious manner. For, let the polynomial gi(xv ..., xn) be of degree mi9 and let it be of degree m^ when regarded as a polynomial in xn with coefficients in K[xl9..., #n_i]. Write 9i(Xl>»->Xn)=*i(xl>"->*n-l)^ (2) at(xl9...,xn_^) is a polynomial in K[xl9...,&n_i] of degree mi — mfi. Our restriction implies that it is not zero. When we apply the results of § 5 to (1), regarded as equations in xn> we obtain a set of polynomials , x ) (i-1 s) having the properties r (i) hi(xu...,xn_1) = ^a{j(x1,...,xn)gj(a1,...,xn) (i = 1,...,a); (ii) necessary conditions that the equations ftfa.--,^-1.¾) = 0 (i = 1,..., r) (3) (where xt is a specialisation of xt) have a solution are
7. NON-HOMOGENEOUS EQUATI0N8 163 (iii) if K0V ...,Sn-i) = ° (*' = !> •••>«), then, either the equations (3) have a solution, or else a*(£i>...>£n-i) = 0 (»= l,...,r). In order to remove this last possibihty we proceed as follows. Let ul9...9 un_x be n — 1 independent indeterminates, and write yi^Xi-UiXn (i= 1,...,71-1), Vn = **• Then we have gt(xl9..., xn) e= (2,(^,..., yn), where Oi(yl9 ..., yn) is a polynomial with coefficients in ^[t^,..., un^\ of degree mi in yx,..., ?/n. In this polynomial y™» appears with coefficient ¢£(^,...,un_l91), where (/*(#!,...,xn) is the set of terms in gi(xv ..., #n) which are of degree m^ We apply the results of § 5 to the equations %.^j = 0 (i-I,...,,), (4) regarded as equations in yn. We then obtain polynomials #i(2/l>..->2/n-l) with coefficients in K[ul9 ..., Mn_J, such ^18^ r (i) Hi{yl9...9y^ = j:Aii{yl9...9yjai{yu...9yn) (i = 1,...,a); i-i (ii) necessary and sufficient conditions for the existence of a solution of the equations flWyi.—>yn-i.yJ = o (»= i,...,r), are Ht{yl9 ...9yn^) = 0 (i = 1,..., a). In this case the possibility (iii) does not arise since which is a non-zero element of K[ul9...,un^\. There exist [III, §8, Th. I] specialisations ul9...9un^1 of fl#(Si,...,un_l9 1) + 0 (i = 1,...,r), and the results just stated hold when ul9..., wn_i are specialised to %> •••>un-v In what follows we assume that these specialisations have been made. II-2
164 IV. ALGEBRAIC EQUATIONS Any solution (x'v ..., aQ °f *^e equations (1) leads to a solution (x'x — uxx'n9..., x'n) of (4) and hence to a solution (x1 — uxxn,..., xn_1 — un_1 xn) of fii(2/i>...,2/n-i) = 0 (i = 1,...,^). (5) Conversely, any solution {y'l9...,y^-i) of (5) implies the existence of a solution y'n of the equations (4) obtained by specialising Vi> -~>yn-ito Vv --->y'n-i> and the solution (y'v ...,^) of (4) implies the solution (y[ + u1y/n9 ...,^n-i + *V-i2/n>2/n) of (1). Thus we have shown how to find a set of equations (4) in n — 1 unknowns, the existence of a solution of which is a necessary and sufficient condition for the existence of a solution of the equation (1) in 7i unknowns. Repeating this process n times we can determine whether a given set of equations (1) has a solution. Theorem I. If the equations 9i(xi>—>Xn) = ° (i =l,...,r) (1) have no common solution, there exist polynomials At(xl9...,xn) such that r H^i(xV -->xn)9i(xl9 '->xn)== !• We have already [§ 5, (5)] proved this result in the case n = 1, so we proceed by induction. To this end, we assume the truth of the result for equations in n — 1 unknowns, in particular for the equations (5) above. Since equations (1) have no solution, neither have (5) and therefore there exist polynomials B^y^ ...,yn^) such that 8 i = IiBi(yi,---,yn-i)Hi(yi>-~>yn-i)- Using the property found above of the polynomials Hi(yl9...,yn_i) we have 8 r l^TlBi{yl9...9yn__1)"ElAij{yl9...9yn)Oj(yl9...9yn) i~i i=i r = 1,0^,...9yn) 0^...^, j-1 8 where C$(yl9..., yn) = S ^(2/i> • • • > 2/n-i) ^;(</i> • • •. ifo)- i«i Replacing yi by ^-¾¾¾ (i = 1, ...,n- 1) and yw by xn we have r l - s ^(^, ...,«„) ^(xl ..., xn), i-l where 4^,...,xn) = (¾^-uxxn,...,xn).
7. NON-HOMOGENEOUS EQUATIONS 165 This important result will be used in the next section. Apart from this application we shall not have occasion to consider other properties of non-homogeneous equations, and we do not pursue the topic further. 8. Hilbert's zero-theorem. Let /<fa, •••,**) = 0 (»'=!,-,') (1) be any set of non-homogeneous equations with coefficients in K, and let g(xv ..., xn) be a non-zero polynomial which is such that (7(^,...,^) = 0 for every solution £l9..., £n of (1) which is algebraic over the ground field. We prove Hilbert's zero-theorem: Theorem I. There exist polynomials ai(xv ...,xn) and an integer s such that T Let z be any new indeterminate. Consider the equations (1) together with the additional equation 2g(*i,...,*n)-l=0 (2) as equations in K[xl9 ...9xn9z], This new set of equations has no solution in any algebraic extension of K9 since for any such solution the left-hand side of (2) is — 1. There can be no solution in any extension of K9 since this would involve the vanishing of the resultant system, and hence lead to a solution in an algebraic extension of K. Hence [§7, Th. I] there exist polynomials ^■i\xV • • • > xw> z)> "\xv • • • > xn> z) in K[xv ..., xn9 z] such that r 1 = 2^t(#l> '••9Xn>z)fi\xl> "'fxn) itml +B(xl9...,xn,z)[zg{xv...,xn)-1]. This identity remains true when we replace z by [g(xl9 ...,¾)]-1. Hence r l = 2-4^(#i, ...9xn9g~~ )fi(xv ...9xn)9 i=i and multiplying both sides by a suitable power of g(xv ...,xn) we have the required result: r [g(xv ...,«n)]8 = Sai(»i» •■.,*«)/<(*!, • ••»*»). (3) i = l the 0,(¾....,xn) lying in K[zv ...,xn].
166 IV. ALGEBRAIC EQUATIONS From this we deduce immediately that if (1) is a system of homogeneous equations, and g(xl9 ...9xn) is a form of degree m which is satisfied by all non-trivial solutions of (1) in an algebraic extension of K, then a relation (3) holds, the a^x^ ...,xn) being homogeneous and of degree ms — mi9 where mi is the degree offt(xly..., xn). 9. The resultant ideal. The procedure we followed in obtaining the resultant system of a set of homogeneous equations was not symmetric in the indeterminates x0,...,xn. We eliminated these in a definite order, and a different order of elimination may well lead to a different resultant system. Also, the number of forms which appear in the resultant system has not been indicated. We show here that a single resultant suffices for n +1 forms in n +1 variables, whereas for any number of forms less than n + 1 no condition for solubility in a suitable extension of the ground field is necessary. The method is based on the consideration of resultant-forms. Let the set of homogeneous equations considered be fi(x09...9xn) = 0 (* = 1,..., r), where ft(x09 ...9xn) sa^+ ...+(0^ (% = 1,..., r), all coefficients being indeterminates. In §6 we have already encountered polynomials dl9..., dN in the coefficients of the forms f{ with the properties expressed by the identities r diXl. = ^Aj(xQ, ...,#n)/j(#o, •••9%n) j =1 for a suitable r and a suitable fc, the Aj being polynomials in xQ9..., xn- with coefficients in the ring of coefficients. We use this property to define resultant-forms. Any polynomial T(av ..., ar9..., oj19 ..., cor) in K[al9..., cor] for which there is a relation r TxI^^A^Xq, ...,^n)/j(^o» •••J^n) (1) for a suitable r and a suitable k9 the Aj being polynomials in x09...,#n with coefficients in K[al9...,a)r]9 is called a resultant-form. The polynomials dl9...9dN of §6 are all resultant-forms. We may give another definition of resultant-forms. Write fi = aixfi+... +<*)iX%i=ft +w,a#<.
9. THE RESULTANT IDEAL 167 Then if we substitute ^ = _/?/<< (2) in (1), theft all vanish, and therefore xlT{ax,...,ar,...,-^,..., -^)-0. (3) Conversely, if for any polynomial T(av ...,ar, ...9a)v ...,a)r) the relation (3) holds, we have T(av...,ar,...,a)v...,(or) (f* f* /»* /* \ al9...,ar,..., - ±±- +0), + ^-,..., - *£- +a)r+ i{-\ on expanding the polynomial T(al9..., ar,..., o)v ..., o)r) in ascending powers of wi + (/*/#™*)• Multiplying this last equation by a suitable power of xn, and using (3), we find r x°T(av ...,<tfr)s2 jfyfro, ...,^)/^0, ...,¾). l Hence T is a resultant-form. From (1) we can deduce (3), and from (3) arrive back at (1), but with xn instead of xk. Since the suffix n plays no special part in this process, we see that from (3) we may deduce (1), k being any one of the numbers 0,1,..., n. The resultant-forms can be used as a resultant system for the set of homogeneous equations fi(x0,...,xn) = 0 (i = 1,..., r). (4) For if, after specialisation of the coefficients, there is a non-trivial solution, and we substitute this in (1), the right-hand side vanishes, and since not all xi are zero, T == 0, T being any resultant-form. On the other hand, if all resultant-forms vanish for a given specialisation of the coefficients of (4), those resultant-forms dv...,dN which belong to the resultant system of (4) also vanish, and the set of equations has a non-trivial solution. As we shall see immediately, the resultant- forms constitute an ideal in the ring of polynomials K[av ...,o)r]. If we can find a basis for this ideal [cf. § 2], this same basis will serve as a resultant system for the equations (4).
168 IV. ALGEBRAIC EQUATIONS If S(av...,G)r) and T(av ..., o)r) are two resultant-forms, and JR(av ...,G)r) is any element of K[av ...,a)r], then since 4 -£M- -§H 4- -£)4--fH it follows that 8{al9 ...9a)r)-T(alf...,o)r) and JB(alf ...,a)r) S(av ..., <*)r) are resultant-forms. Hence resultant-forms constitute an ideal in K[al9...,cor]. We call this the resultant ideal. If the product of two elements of K[av ...,o)r] lies in this ideal, one of the elements must lie in the ideal. For if R(av ...yo)r)9 U(av ..., cor) are the elements, we have 4 -£)4---fH- One of the factors on the left must vanish, and therefore one of the polynomials must be a resultant-form. An ideal with this property is called a prime ideal. We now prove Theorem I. If the number r of forms fi is less than n+1, there are no resultant-forms different from zero. If T is a non-zero resultant-form, 4,.^---,-!>•••>-f>o, (5) but T{alt...,0,,.,...,0)^...,0),)^0. (6) We show that a relation of this type continues to hold even when the coefficients av ...,ar,... are no longer indeterminates, but specialisations av ...,ar,... in K. Consider, first of all, the specialisation ¢4^%. It may happen that T{av ...,ar,...,o)v ...,(or) = 0. If this is so, T(av..., (1,,...,^,..., o)r) = {a1-a1)9 T*(au...,o)r) for some integer s, where T*(av ...,ar, ...,o)lt ...,0),)^0.
9. THE RESULTANT IDEAL 169 Since it follows that (0,-0,)-^(0,,...,0,,...,-^ -|)S», and so T*(av ...,ar,o)v...,o)r) is a non-zero resultant-form. Specialising each of the indeterminates present in turn, we find at each stage a non-zero polynomial which vanishes when the substitution (2) is made; that is, a resultant-form. Hence, finally, we obtain a non-zero resultant-form when a1->av ..., ar->ar,..., provided that we began with a non-zero resultant-form before any indeterminates were specialised. Now, since r <n + 1, we may specialise av ...,ar... so thatft->fi9 where fi^xTii + UiX™* (i = l,...,r). The existence of a non-zero resultant-form before the specialisation involves the existence of a non-zero polynomial T{(DV ...,0),) such that But x0,..., #r_i, xn are independent indeterminates, and we are therefore led to a contradiction. This proves the theorem. It follows that the resultant system of r(r < n+ 1) homogeneous equations in n 4-1 unknowns vanishes identically, and therefore that the equations always have a solution in some algebraic extension of the ground field. Theorem II. There is a non-zero resultant-form for n 4-1 equations inn+1 unknowns. There must be one, at least. For if not, n 4-1 homogeneous equations always have a solution, for any specialisation of the coefficients. But there is an evident specialisation for which the equations become xfi = 0 (* = 0,...,n), and these equations have no non-trivial solution. Our assumption is therefore false, and there is at least one non-zero resultant-form!
170 IV. ALGEBRAIC EQUATIONS The argument of the previous theorem shows us that when r = n + 1 there can be no non-zero resultant-form which is independent of o)n+1. Assumption of the contrary hypothesis leads to an algebraic equation connecting xQ,...,xn. Hence if we choose, from the ideal of all resultant-forms in K[av ...,a)n+1], a polynomial whose degree in o)n+1 is the least possible, this degree is not zero. If this polynomial is reducible, one of its irreducible factors must belong to the ideal, which is prime, and this factor must also be of the minimum possible degree in o>n+1. We call this resultant-form R, and prove Theorem III. Every resultant-form is divisible by R. R = Aa)*+1+...9 when arranged in descending powers of (c>n+1, and let be any member of the ideal, similarly arranged. Since /i > A, AS-Bo)f^l\R=T is another member of the ideal, with a lower degree in o)n+1 than 8. Repeating this process, using T instead of S for the next stage, we finally arrive at a polynomial T* = AkS-QR in the ideal which is of lower degree than R. This must be a zero polynomial, and therefore AkS = QR, showing that AkS is divisible by R. Since R is irreducible and cannot divide A, which is independent of con+v R divides S. This proves the theorem. This resultant-form R is therefore a basis for the ideal of resultant- forms in K[al9 ...,o>n+1]. A necessary and sufficient condition for the 7i+l homogeneous equations fi(x0,...yxn) = 0 (i= 1,...,71+1) to have a non-trivial solution is given by the vanishing of R. We call R the resultant of the system.
10. THE ^-RESULTANT OF A SYSTEM OF EQUATIONS 171 10. The n-resultant of a system of equations. Suppose that we know that a system of homogeneous equations over K fi(z09...,xn) = 0 (»'=l,...,r) (1) has a finite number of solutions. Let these be in some extension of K. To the equations (1) we add the equation with indeterminate coefficients „ fr+1 = U0X0+...+UnXn = 0. (2) Since the ui are indeterminates, StfigP + 0 (j =1,...,^), t = 0 and therefore the resultant-forms of equations (1), (2), of which (1) are specialised, do not all vanish. These resultant-forms are polynomials in u0,..., un. If one such is T(u), r+l T(u)xrn = ^Ai(x0y...,xn,u) fi(xQy...9xn). (3) Now the fi} if they contain u0,...,un, are homogeneous in these indeterminates. If we equate sums of terms of equal degree in the u'a on each side of (3), we obtain equations similar to (3), but the T(u) are now homogeneous in u0,...,un. These T(u) are still resultant-forms, satisfying equations of type (3), and we therefore need only consider resultant-forms which are homogeneous in ^0> •••> unm Let 6^),...,6^) (4) be a basis for the resultant-forms of equations (1), (2). We may assume that each b^u) is homogeneous in the w's, and is also of degree greater than zero, since all resultant-forms of the set (1) vanish. The set of forms (4) serves as a resultant system for the r+l equations given by (1) and (2). This resultant system vanishes for the specialisations ^-+¾ (i = 0,...,n), if and only if a solution of the equations (1) satisfies (2); that is, if and only if f^u0^+... + ujf = 0, for some value of j. In other words, the common zeros of the
172 IV. ALGEBRAIC EQUATIONS forms (4), considered as forms in the set uQy...,un> are precisely the zeros of the product Q(u)=lJf<fi=m»o6P+-.-+Un8& i-1 1 in a suitable extension ring of K[u]. It follows by Hilbert's zero-theorem that on the one hand [W^i^SW (i =1,...,0, (5) and on the other that t [GMY^OiMbiiu). (6) Now, the fa are linear forms in uQ,..., un, and therefore irreducible. From (5) the b^u), and therefore their highest common factor D(u), can be divided by the linear factors/^). But (6) can also be written [G(u)Y = B{u)D(u)9 from which it follows that D(u) contains no other linear factors but these/(i). We deduce therefore that D(u) = n(u08P+...+ungfri (cr,>l). Since D(u) is the highest common factor of the forms (4) which are a basis for the resultant-forms of equations (1), (2), D(u) divides all resultant-forms, and, in particular, D(u) divides every member of any resultant system of the r+1 equations (1) and (2). We have proved Theorem I. The linear forms fa which determine the solutions of equations (1) are found by the resolution into factors, in a suitable extension field, of the form D(u). This form D(u) is the highest common factor of the resultant-forms offv... ,/r, /r+1, and is called the u-resultant of A,-.,frit the set of equations (1) has an infinite number of solutions, we may extend the notion of the ^-resultant. We consider the equations (1) together with the set of linear equations with indeterminate coefficients n /r+i=£ Vj = 0 (* = 1,...,5). (7) i-o If s is large enough, the sets of equations (1) and (7) have no common solutions. For if we take s = n+ 1, the set (7) taken by itself has no solution.
10. THE tt-RESULTANT OP A SYSTEM OF EQUATIONS 173 Choose s to be the smallest integer such that (1) and (7) have no common solution. Consider the resultant-forms of these equations. Generalising our previous argument, we see that the only resultant- forms we need consider are those homogeneous in each set of indeterminates ui = (ui0>-~yuin) (* = 1, •••>*)• If for the specialisations usj-+u8j U = 0,...,71) the equations (1) and (7) have a solution (£0,..., £n), all resultant- forms T(uv ..., u8) vanish for this specialisation. That is, r(t*i Vi»5i)s0« Reasoning as above, we see that n is a factor of T(uv ...,u8). If D(ul9...,us) is the highest common factor of all resultant-forms, the factor (7) divides D(uv ..., us). But since this highest common factor, regarded as a form in w«o> --->u8n has only a finite number of linear factors in any extension field, the equations /, = 0 (i = l,...,r + *-l) (8) have only a finite number of solutions. The ^-resultant of (8) is D(uv ...,us). It may also be called the ^-resultant of equations (1). In a case which will be considered in vol. H, in which the equations (1) define an irreducible variety, this generalised ^-resultant is of considerable importance. *
* BOOK II PROJECTIVE SPACE CHAPTER V PROJECTIVE SPACE: ALGEBRAIC DEFINITION 1. Introduction. It has been customary to follow two distinct and alternative lines of approach in defining a projective space of n dimensions. One is purely algebraic, and the other is based on the so-called 'propositions of incidence'. We propose to consider both methods of approach, devoting this chapter to the algebraic approach, and the next to the synthetic method. One of our main objects will be to compare the two methods in considerable detail. For this reason we shall begin by defining a more general space than that with which we shall ultimately be concerned. We shall impose new conditions only when we have examined their significance in some detail. Our first action in defining a projective space algebraically is to select the ground field K. Initially we shall impose no restrictions on K. It may be non-commutative, and of finite characteristic. When K is non-commutative we shall be led to consider another field K, which bears to K a relation similar to isomorphism, but differing in one respect. To define this relation, which we call skew- isomorphism, in general terms, let 8 denote a set of elements in one-to-one correspondence with the elements of K, a and a denoting corresponding elements of if and 8 respectively. We define addition and multiplication in the set by the rules a + b = a + b, ab = (ba). It is easily proved that the elements of S, with these laws of composition, form a field, which we denote by K. This field K is isomorphic with K, the reversal of the order of the factors in the product being, simply, a change of notation for the product of two elements a and b in K. But we shall find it convenient to distinguish between
176 v. projective space: algebraic definition an isomorphism between two fields K and K in which ab corresponds to a&, and one in which ab corresponds to ba. We shall refer to the first case simply as isomorphism, and to the second as skew-isomorphism.* In the case of a commutative field, of course, there is no difference between isomorphism and skew-isomorphism. 2. Projective number space. A set (a0,..., an) of n + 1 elements of the ground field K will be called an (n+ l)-tuple if not all the elements a0,..., an are zero. Two (n+ 1 )-tuples (a0,..., an), (/?0,..., /?n) are said to be right-hand equivalent if there exists an element A of K such that 0 ^ ,. - x <Zi = fliA (i = 0,...,n), and left-hand equivalent if there exists an element /i of K such that ai = M (i = 0,...,n). Clearly, when the two (n+ l)-tuples are right-hand equivalent, A cannot be zero, since, otherwise, each ai would be zero. Similarly, if they are left-hand equivalent, /i cannot be zero. Right-hand (left-hand) equivalence is clearly reflexive, symmetric and transitive. A set of right-hand equivalent (n+ l)-tuples is called a point of right-hand projective number space of dimension n over K. The aggregate of such points is called a projective number space of dimension n over K. This is denoted by PNrn(K). If left-hand equivalence is used in place of right-hand equivalence we obtain a left-hand projective number space of dimension n over K, denoted by PNln(K). If K is commutative, pN^K) _ P^;(Z)# If K is not commutative, and K is skew-isomorphic to K, two (w+l)-tuples (a0, ...,an), (/?0,...,/?n) over K are right-hand (left- hand) equivalent if and only if the corresponding (n+ 1)-tuples (a0,..., an), (/?0,..., /?n) are left-hand (right-hand) equivalent. Hence, in an obvious sense, PNln(K) and PNrn(K) are similar spaces, and PNln(K) and PNrn(K) are also similar. For this reason it will be sufficient to develop the properties of PNrn(K), those of PNln(K) being read off from those of PNrn(K). Let A = (a{j) be any regular (n+1) x (n+1) matrix whose elements are in K, and consider the equations n yi = Ti*ijXj (i = 0,...,71), (l) j-0 * van der Waerden uses the term inverse-isomorphism.
2. PROJECTIVE NUMBER.SPACE 177 and also the equivalent equations Xi^itPiiVi (i = 0,...,71), (2) where B = (/¾) is the inverse of A. If (#0 A,:.., xn A) is a set of right- hand equivalent (w+ l)-tuples, the equations (1) transform these (n + 1 )-tuples into the set (*/0^> • • • > 2/n^) °f equivalent (n + 1 )-tuples; that is, the equations (1) transform a point of right-hand projective number space into a point of right-hand projective number space. By virtue of (2) this correspondence is one-to-one. A set of equations (1), which we shall usually write in matrix form as y = Ax, (3) determines a projective transformation of PNrn(K) into itself. To each regular (n+ 1) x (n+ 1) matrix A over K there corresponds a proj ective transformation (1). Let A x and A 2 be two regular matrices, defining the transformations V = Axx, (lx) V = A2x. (12) We operate on a point x with (1^, and follow with (12). The point x becomes y = Atx, and y is transformed into z = A2y = ^^a;; and so the relation between x and z is z = A2Axx. Since [II, §3] A2AX is regular, the product of the two projective transformations of PNrn(K) given by (lx) and (12) is a projective transformation of PNrn(K), and hence we have Theorem I. The projective transformations of PNrn(K) into itself form a group. Projective transformations of PNln(K) can be defined in a similar way. It should be noted that the equation of the transformation corresponding to the matrix A is now y = xA9 (4) where x and y are now 1 x {n+ 1) matrices. Let K* be an extension of K. The points of PNrn(K) are a subset of the points of PNrn(K*). The projective transformations of PNrn(K*) correspond to the regular (n+ 1) x (n-h 1) matrices over HPI 12
178 V. projective space: algebraic definition K*, and a subset of these are matrices over K. The projective transformations corresponding to the matrices of this subset form a subgroup of the group of projective transformations of PNrn{K*), and under the transformations of this subgroup the points of PNrn(K) are transformed into points of PNrn(K). We say that PNrn{K*) is the extension of PNrn(K) corresponding to the extension K* of K. Similarly, the group of projective transformations of PNrn(K*) is an extension of the group of projective transformations oEPN*n(K). 3. Projective space of n dimensions. Let S be a set of elements which can be put in one-to-one correspondence with the points of PNrn(K), and let T denote a correspondence such that if A is any element of 8, and B any point of PNrn(K), there corresponds to A a unique point T(A) of PNrn(K), and to B a unique element T-X{B) of S. Then to each element A of S there corresponds a set of equivalent (n+ l)-tuples (a0A,...,anA), where T(A) is (a0, ...,an). Any (n + 1)-tuple of this set is called a set of coordinates of A. Now consider a projective transformation in PNrn(K). If this takes the point T(A) of PNrn(K), which corresponds to the element A of S, into T'(A), the correspondence between the elements A of #and the points T'{A) of PNrn(K) is one-to-one, and thus a new coordinate system is defined in S. Indeed, a coordinate system is defined for each projective transformation of PNrn(K). The various coordinate systems obtained in this way will be called allowable coordinate systems in S. From the group properties of projective transformations, discussed in § 2, it is clear that the same set of allowable coordinate systems in S is defined if the initial correspondence T between S and PNrn(K) is replaced by any other correspondence T' which gives an allowable coordinate system. The set S, with the set of allowable coordinate systems, is called a right-hand projective space ofn dimensions over K, and the elements of S are called the points of this space. The space may be denoted by P^K), but various other notations will be used. In the case in which K is commutative, and there is no doubt about the ground field, the symbols [n]9 Sni S'n will be used to denote a projective space of n dimensions, the notation Sn being particularly useful when a number of projective spaces have to be considered simultaneously. If (x0i...,xn) is a set of coordinates of a point P in P^iK) in a given allowable coordinate system, the coordinates (y09...9yn) of
3. PROJECTIVE SPACE OP U DIMENSIONS 179 this point in another allowable coordinate system are given by the equations n Vi = Ti^iM (i = 0,...,^), where A = (a#) is a regular (n+ 1) x (n+ 1) matrix over K. The equations for transformation of coordinates in Prn(K) can therefore be written, in matrix notation, as y = Axy and there is one such transformation for each regular (n + 1) x (n + 1) matrix over K. We consider how P^iK) can be extended when the field K is replaced by an extension K*. In § 2 we saw how to extend PNrn(K). We now embed the set S in a set #* which can be put in one-to-one correspondence with the points of PNrn(K*) in such a way that the elements of $* which lie in S correspond to points of PNrn(K*) which are in PNrn(K). This is clearly possible; a set such as S* can be formed by the elements of S and the points of PNrn(K*) which are not in PNrn(K). Using the given correspondence between 8* and PNrn(K*), and the projective transformations of PNrn(K*)> we construct a projective space of n dimensions P^K*) whose points are the elements of $* in which P^iK) is embedded. Amongst the allowable coordinate systems of P^(iC*) there is a subset which induces the allowable coordinate systems of Prn(K). In later chapters we often have to replace a field K by an extension K*. When this is done we shall assume that P^K) has been extended to P^(Z*), as above. The effect of this is to add new points to P^(iT), and to exteijd the group of allowable transformations. Let us suppose that we have a coordinate system in P^(Jf*) in which the points of Prn{K) can be given coordinates in K. Consider the coordinates (aj,..., a*) of a point R in P^(iC*). (i) If, for some A* in J5T*, a* A* (i = 0,...,71) is in K, the point R is in PTn{K), and is sometimes called a rational point of Prn(K). (ii) If, for some value of A*, each ocf A* is algebraic over K, R is called an algebraic point of P^iK). (iii) If, however, for all A* in K*f at least one a? A* is transcendental over K, R is said to be a transcendental point of P^K). 12-2
180 v. projective space: algebraic definition This completes our description of a right-hand projective space P^(JSl). The method of construction of a left-hand projective space Pln{K) is now evident, and the reader can immediately read off the relations between Pln(K) and P^(J?), where K is skew- isomorphic to K. 4. Linear dependence in P&K). Consider k +1 points A\...,Ak in«(*),andlet {< ^ (. _ „ k) be their coordinates in an allowable coordinate system. We say that the points A0,..., Ak are linearly dependent in P^K) if there exist elements A0,..., Xk of K, not all zero, such that S^ = o (i = 0,...,^). (l) In order that linear dependence may be considered as a property of the points A0,..., Aky it must be shown to be independent of the coordinates selected for A0,..., Ak in the chosen coordinate system, and also to be independent of the allowable coordinate system. To prove the first result, let (/?j, ...,/¾) be other coordinates for A1 in the given allowable coordinate system. Then where /i€ is some non-zero element of K. From (1) we deduce that 2/^(/^) = 0 (i = 0,...,^). Hence, if a relation of type (1) holds for one choice of the coordinates in a given allowable coordinate system, a relation of this type also holds for any other choice in the given coordinate system. Now consider any allowable transformation of coordinates Siven ^ y = Cx, (2) where C = {ytj) is a regular (n+ 1) x (n+ 1) matrix over K. In the new coordinate system the coordinates of A1 can be taken as (/&...>An), where # = 27*,«,- Then 2$A< = S Sr^-0, i«0 1-0 />«0 from (1). Hence, again, if a condition such as (1) holds in one
4. LINEAR DEPENDENCE IN P^(k) 181 allowable coordinate system, a similar condition holds in any other allowable coordinate system. Thus the linear dependence of points in F^K) is a property of the points, and does not depend on the allowable coordinate system. Recalling the properties of the linear dependence of vectors derived in Chapter II, we have the following alternative statements of the conditions for the linear dependence of the points ^4°,..., Ak. (i) If (al,...,ai) (i = 0,..., k) are the coordinates of A0,..., Ak in a given allowable coordinate system, a necessary and sufficient condition that A0,..., Ak be linearly dependent is that thek+1 right-hand vectors (aj, ...,<) (i = 0, ...,&) be linearly dependent. (ii) A necessary and sufficient condition that A0,..., Ah be linearly independent is that the matrix 'oc°o a? K aj *\ • *l m • • . a? <x\ «5. be of rank k+1. An immediate corollary is Theorem I. Any mpoints of Prn(K), where m>n+l, are linearly dependent. We can introduce a useful symbolic notation if we assume that the symbol A1 for a point represents a particular (n 4-1)-tuple in the class of equivalent (n+ l)-tuples which give the coordinates of Ai in a given allowable system. The transformation (2) then defines a particular set of coordinates for each A1 in the new coordinate system, and this set will also be denoted by A1. When we use the symbol A1 for a point in this way, we extend our notation by multiplying A*, on the right, by a non-zero element of K, so that if Ai = (a0,...,an), then AlX = (a0A,...,anA), and the symbols Ai and u4*A represent the same point. The equation (1) can then be written as k £4^ = 0, <-o and this relation does not depend on the allowable coordinate system used.
182 V. PROJECTIVE space: algebraic definition Any point A such that k A/i^^A^ for some non-zero /i in K, is said to be linearly dependent onA°9...9Ak. If lc = n9 and in a given allowable system of coordinates we take A* = {*{qA» •••>*«, -..,*<tt) (*' = °> •••»*) = (0, 0,...,1,...,0), we obtain n+1 linearly independent points. Since any point A = (a0, ...,an) can be written as n A = ZA%9 every point in Prn(K) is linearly dependent on these n+\ points. They are said to form the simplex of reference for the given coordinate system. The properties of the linear dependence of points can now be read off from the properties of the linear dependence of vectors [II, § 1]. It will not be necessary to give proofs in detail, since these can be deduced at once from the corresponding proofs for vectors. The main results are as follows. In the first, A1 has the symbolic meaning just defined. Theorem II. If A°9 ..., Ak are linearly independent, and if k A\^J^A% is linearly dependent on them, the (Jc+ \)4uple (A0,..., A^) is uniquely determined to within the class of equivalent (&+ \)-tuples. Theorem III. If B°, ...9Bh are linearly dependent on A°9..., Ak9 and P is linearly dependent on B°,..., Bh9 then P is linearly dependent onA°9...9Ak. Theorem IV. In any finite set of points A0, ...9Ak we can find a linearly independent subset A^,.:., A** such that any point linearly dependent on A0,..., Ak is linearly dependent on A{o,..., A**. Theorem V. If B°9 ..., Bh are h+\ linearly independent points, each of which is linearly dependent on A0,..., Ak, then h^k, and we can find k — h points A**+i,..., A1* in the latter set such that every point linearly dependent on A0,..., Ak is linearly dependent on B°,..., Bh, 4^+1,...,21^. (Exchange Theorem.)
4. LINEAR DEPENDENCE IN ^(k) 183 If u4°,..., Ak are k +1 Unearly independent points of P^iK), the aggregate of points of Prn(K) which are linearly dependent on them constitute a linear space of k dimensions. We shall call a linear space of 1 dimension a line, and a linear space of 2 dimensions a plane. From the first part of the Exchange Theorem we deduce that any h+l points of a linear space of k dimensions, where h>k, are necessarily linearly dependent. On the other hand, there exists at least one set of k+1 independent points, A0,..., Ak. If B°,...,Bk is any other set of k+ 1 Unearly independent points of the space, the second part of the Exchange Theorem tells us that the linear space determined by B°,..., Bk is the same as that determined by A»,...,Ak. Any set of k + 1 Unearly independent points of a linear space of k dimensions is called a basis for that space. We use the above results to show that a linear space of k dimensions is a Prk(K). Let A°,..., Ak be any basis for the Unear space. Then any point Q of the space is given by an equation where the (&+l)-tuple (A0,..., A^.) is uniquely determined by the point Q, to within right-hand equivalence. This estabUshes a one-to- one correspondence between the points of the linear space and those of PNrk(K). Now consider what happens if B°,..., Bk is any other basis for the linear space. We have k k Bi^^iA^ji, and 4* = 2#a«. k k Hence Ai: = 2 2 ^Vw^. But this must be an identity, since ^4°,..., Ak are Unearly independent. The matrices n , n . A , x J? = (/y, A = (0¾) are therefore inverse matrices. It follows that A and B are both regular. k Now, if Q = 2 4*Af, we also have o 0 0 0 ft where ^ = 2½^ (3) o
184 v. projective space: algebraic definition and therefore the change of basis from A0,...,Ak to B°,...,Bk results in the projective transformation (3) in PNrk(K). Conversely, any projective transformation in PNrk(K) corresponds to a change of basis in the linear space. Thus a linear space of k dimensions is a PkiK), each allowable coordinate system corresponding to a choice of basis in the linear space. From Th. Ill we deduce at once that if Sh is a linear space of h dimensions determined by a basis B°,..., Bh, and if each Bl lies in a linear space Sk, then every point of Sh lies in Sk. That is, Sh lies in Sk. We write this 0 0 It is evident that if Sa^8b9 and Sb^Sc, then Sa^Sc. Also, if Sa^Sb and Sbs Sa, then Sa = Sb. From Th. IV we deduce that any k + 1 points lie in an Sh (h < k) and that h < k if and only if the points are linearly dependent. Again, let Sp and Sq be two linear spaces in P^iK), and suppose that P°,..., P* is a basis for Sp, and Q°,..., 0* a basis for Sq. If the p + q + 2 points P°,..., Pp, Q°,..., Q1, are linearly dependent, so that SP^ + SQ^O, o o we deduce that Sp and Sq have a point in common. For neither P Q 2 P*A^ = 0, nor 2 QlH = 0, since P°,..., Pv are linearly indepen- oo p dent, and so are Q°,..., Qq. Hence 2 J0*^ ig a point in /S^ which, Q o since it can be written as — 2 Q1/^, is linearly dependent on Q°,..., QQ, and therefore lies in Sq. ° Finally, we note the trivial fact that if P and Q are distinct (that is, linearly independent) points of P^iK) the point which is represented symbolically by P + Q is necessarily distinct from both P and Q; hence, since any St contains at least two points, any S± contains at least three points, and we may sum up the results established in this section as follows: In P'niK) there are linear spaces of dimensions 0,1,2,...,n, a space of 0 dimensions being a point. Between any two such spaces 8h and Sk there may exist a relation of one ' lying in' the other, say Sh £ Sk. This relation satisfies the conditions I. If 8h s Sk, and Sk c Sh then Sk = 8h. II. IfiSf^s^and^c^then^s^
4. LINEAR DEPENDENCE IN Prn{K) 185 A set of p + 1 points is linearly dependent if there exists a linear space Sq containing them, where q <p. III. Every S± contains at least three distinct points. IV. Given p + 1 linearly independent points there exists at least one Sp containing them. V. In any Sp there exists at least one set of p+1 linearly independent points. VI. If P°,..., Pp are p + 1 linearly independent points which lie in an Sq, any Sp containing them lies in Sq. VII. If Sp and Sq are any two linear spaces of dimension p and q, and P°,...,Pp are p+1 independent points of Sp and Q°, ...,0^ are q + 1 independent points of Sq, then if the p + q + 2 points P°,..., Pp, Q°,..., Qq are linearly dependent there exists at least one point R which lies in 8p and in 8q. VIII. There exists at least one set of n+ 1 linearly independent points, but any set of m points, where m > n + 1, is linearly dependent. We have, indeed, proved a great deal more, but it is desirable at this stage to emphasise that a right-hand projective space satisfies these propositions, which we call the Propositions of Incidence. In the next chapter we shall begin by taking an aggregate of undefined elements, which we shall call linear spaces of dimensions 0,1,2,.... We shall postulate the existence of a relation <= between certain of these spaces, and assume that this relation satisfies I and II above. The statement which follows II is then taken as a definition of linear dependence. We shall then impose the further conditions III-VIII, noting that it is only in VIII that the number n appears. It will be our main object in Chapter VI to show that these linear spaces of dimensions 0,1,2,... can necessarily be represented as the linear spaces of a P^(if), for a suitably chosen K. This will follow directly from the above postulates if w > 2. If w = 2 a further postulate will be necessary. The case n = 1 is without interest. For a left-hand space Pln(K) a theory of linear dependence, similar to that developed for P^if), may be stated. If 4* = («$,..., 4) (» = 0,...,¾) are & + 1 points, the points linearly dependent on them are the points k I k k \ 0 \ 0 0 / and the theorems proved earlier in this section can be proved for
186 V. projective space: algebraic definition Pln(K), with the necessary formal alterations. In particular, we arrive at the same propositions of incidence, and we shall therefore be able to prove that, with the necessary reservations if n ^ 2, these propositions serve to define a left-hand projective space. This implies that a left-hand projective space may be regarded as a right-hand projective space, and conversely. The explanation of this fact will be found in our method of constructing a projective space from the propositions of incidence. We shall have to construct a field, and we shall see that there are two ways of doing this, leading to two fields which are skew- isomorphic. Now, if K is skew-isomorphic to K, there is a one- to-one correspondence between the points of P^K) and those of Pln(K), and a one-to-one correspondence between their allowable coordinate systems. Choosing corresponding allowable coordinate systems in Pln(K) and P^iK), the points (aj,..., a*n) and (3j,...,a£) correspond. The point (27^,...,27¼¾) of Pln(K) corresponds to the point (27^,...,27¾¼) of P'niK). Hence, the points of Pln(K) which are linearly dependent on a given set of m points of Pln(K) correspond to the points of P^K) linearly dependent on the m corresponding points. Indeed, corresponding to any property of spaces 8a, #6,... in Pln{K) which remains unchanged under any allowable projective transformation in Pln(K) there is a property of spaces 8ai Sbi... in Prn(K) which remains unchanged under any allowable projective transformation in P^(JSl ). We say that the two spaces are projectively identical. 5. Equations of linear spaces. Duality. We consider a right- hand projective space P^(K), and a left-hand projective space Pln(K). In each of these we fix, quite arbitrarily, an allowable coordinate system, but, having done this, we regard the two as associated. We now show how we can proceed to associate with each allowable coordinate system in P^(K) a unique allowable coordinate system in Pln{K).
5. EQUATIONS OP LINEAR SPACES. DUALITY 187 If the transformation from the selected coordinate system in P^K) to the new coordinate system is y = Axy the transformation from the selected coordinate system in Pln(K) to the coordinate system which we associate with the new system in F'niK) is taken to be , , A _x y — x j± (where xf and y' are, of course, 1 x (n + 1) matrices). It can be verified at once that this association of allowable coordinate systems in the two spaces is one-to-one, and is determined by the above rule when the selected coordinate systems are any pair of associated systems. Now consider a linear space 8k in P^K). Let A0,..., Ak be a basis for this space, where A* = (aj, ...,<) are the coordinates of A1 in a given allowable system. Consider the equations n S^j = o (; = o,...,&). (1) Since the (w+ 1) x (fc+ 1) matrix (of) is of rank k+1 [§4, (ii)], the solutions of the homogeneous equations (1) form a left-hand vector manifold of dimension n — k [II, §6]. Let the non-zero vectors of this manifold be taken as the coordinates of the points of a linear space in the associated coordinate system in Pln(K). Its dimension will be n — k - 1. We denote the space by En_k_1. If «...,<) (l = 1,...,71-¾) is a basis for Zn_k_v the (n — k) x (n + 1) matrix (uj) is of rank n — k, and from (1), StijaJ = 0 (i = l,...,?i-&; A = 0, ...,&). (2) But, since (wj) is of rank n — k, the equations 5>fo = 0 (i = l,...,n-&) (3) are satisfied by vectors of a right-hand vector manifold of dimension k+ 1. The non-zero vectors of this manifold are the coordinates of points of a linear space of dimension k in P^iK). By (2) above, this space contains A0,..., Aky and therefore coincides with 8k. Thus Sk determines -Tn_fc-i> and conversely 2?n_fc_i determines Sk.
188 v. projective space: algebraic definition We now show that this association does not depend on the particular associated pair of coordinate systems in P^iK) and Pln(K) of which we have made use. To see this, we first note that the relation between Sk and -£n_fc_i can be expressed as follows: the equation n i = 0 is satisfied for every point (u0,...,un) of Zn^k-V for all choices of (xo>'"9xn) *n &k> and conversely. Now, if we perform the transformation of coordinates A y = Ax in jP^(jfiT), and the associated transformation v = uA"1 in Pln{K), then (noting that u, v are both lx(n+l) matrices) n n 2 viVi = VV = uA"xAx = UX = 2 1^0¾ = 0. 0 0 Hence the property characterising the relation between Sk and 2?n_fc_i remains unchanged in the new associated pair of coordinate systems. The spaces Sk and Zn-k-i are said to be dual spaces. If we take any n — k points of £n-k-i which form a basis for this space, and which have the coordinates «...,<) (» = 1,...,^-¾) in an allowable coordinate system in Pln(K), then the points of Sk in P'niK) satisfy the equations Sttfo = 0 (* = l,...,n-fc) (3) j = 0 in the associated coordinate system. We may call (3) the equations ofSk. Given any set of equations 2^==0 (i = 1,2,...), (4) o we may consider the space in Pln(K) linearly dependent on «...,<) (< = 1,2,...). If this is a Zn_k_ly the solutions of (4) are the points of the dual Sk, and the equations therefore determine this Sk. In algebraic
5. EQUATIONS OF LINEAR SPACES. DUALITY 189 language, if the matrix of coefficients of a set of linear homogeneous equations (4) is of rank n — lc, these equations determine a unique Sk. Now consider two linear spaces 8p and Sq in P^iK) with respective equations n 3>ja;,= 0 (i = l,...,n-2>), o n and y£lvjxj = 0 (i = 1, ...,n —g). o Suppose that the (2n—p -q) x (n+ 1) matrix /u\ (u = (tty), v = (t^)) (5) is of rank n — h. Then the two sets of equations, taken together, determine an Sk. Each point of Sk lies in Sp and in Sq, and each point common to Sp and Sq must lie in Sk. We call Sk the intersection of #p and #¢, and have proved Theorem I. The intersection of two linear spaces is a linear space. If the matrix (5) is of rank n - fc, the 2n — p — q points «...,<) (* = l,..-,n-p),J « •••><) (*'= l,...,n-g) j in Pln(K) are Unearly dependent on n — fc of them, and therefore the duals En_p_1 and En_q_1 of #p and Sq lie in a En_k_1, which is the dual of Sk. Moreover, there is no smaller space containing both Zn-p-x and En_q_1. For any space containing both of these spaces must contain the 2n—p — q points (6), of which n — k are linearly independent. Calling the smallest linear space which contains two linear spaces their join, we have Theorem II. The dual of the intersection of two linear spaces is the join of their duals (and conversely). As a corollary we deduce that If Sp 2 Sq, then E^_x <= Zn_q_v We now find a relation between the dimensions of the intersection 8k and the join Sr of two linear spaces Sp and Sq. Let A0,..., Ak be a basis for the intersection. By the Exchange Theorem we can find bases P°,..., P» for Sp, and Q°,..., Q<* for 8q such that pi = A< = Qt (i = 0,...,¾). Then clearly, any space containing P°,..., Pv, Qk+1, ...,01 contains Sp and Sq, and therefore Sr, their join. Therefore r^p + q-k.
190 v. projective space: algebraic definition A similar argument applies to the duals £n„p_v En-q-i and their join, which, by Th. II is Sn_k^1. Hence n — k— 1 ^(n-p — l) + (n — q—l) — (n — r—l), that is, r^p+q— k. It follows that k + r = p + q. Hence we have Theorem III. If Sp and SQ meet in Sk, their join is a linear space of dimension p + q — k. This theorem remains true when Sp and Sq have no intersection, provided that we then put k = — 1. We have established a correspondence between P^K) and Pln{K) in which an Sk of the former corresponds to a Sn_kr_1 of the latter, and intersections and joins correspond to joins and intersections respectively. If we establish any theorem for P^K) which is based only on properties of linear spaces, their intersections and joins, this correspondence enables us to state forthwith a theorem for Pln{K)> without further proof. For historical reasons this relation between theorems for P^iK) and for Pln(K) is called the Principle of Duality for projective space. If if is a field skew-isomorphic to K, there is a correspondence between Pln(K) and Prn(K) in which a linear space of k dimensions corresponds to a linear space of k dimensions, and intersections and joins correspond to intersections and joins respectively. Hence, by the Principle of Duality, if we have any theorem in P^iK) which concerns linear spaces Sa,Sb,..., their intersections and joins, we obtain a corresponding theorem in P^iK) concerning linear spaces Sn_a_l9Sn^b_1,..., their joins and intersections. If K is a commutative field we can take K = K, and the dual theorem is therefore true in P^K). Again, if the original theorem is true for all ground fields K, it is true in particular for Prn(K), and therefore the dual theorem is true in P^iK). This last case is the most usual one. The theorems we shall prove will be either (i) true for all ground fields Ky or (ii) true for commutative ground fields K, or (iii) true for all ground fields with a given characteristic,
5. EQUATIONS OF LINEAR SPACES. DUALITY 191 It will be noted that if K belongs to one of these classes, the skew- isomorphic field K belongs to the same class. Hence, every theorem which we shall prove in P^iK) will have a dual in P^K). 6. Desargues' Theorem. In this section we consider the celebrated Theorem of Desargues which deals with two triangles in perspective in P^K). In order to avoid needless repetition in the next chapter, we shall prove the theorem and its converse as consequences of the propositions of incidence [§4]. We notice here some immediate deductions from the propositions of incidence, which we shall use in this section. A more complete treatment will be found at the beginning of the next chapter. Every Sx (or line) contains at least three distinct points. Two distinct points determine one line. Any S2 (or plane) contains three linearly independent points. If a line I is determined by the points P°, P1, and a line m is determined by the points Q°, Q1, and the points P°, P1, Q°, Q1 are linearly dependent, that is, the lines are coplanar, there is at least one point R on both lines. If P is any point in a plane n, at least three distinct lines can be drawn through P; for if P, Q, R is a set of linearly independent points in n, the line QR contains at least one point different from Q and R. Let this be S. Then PQ, PR, PS are three distinct lines through P. To complete the deductions from the propositions of incidence for a plane we show that if A, B, U are three distinct collinear points in n, we can draw lines a, b, u through A, B, U respectively which lie in n, are distinct from the line A U, and do not all pass through one point. For there is at least one point X in n such that X, A, U are linearly independent. The line XU contains a point Y distinct from X and U. The lines a = AX, b = BY,u = TJX satisfy the conditions. The theorems we shall use in #3 are: a line meets a plane in one point, or it lies in the plane; two planes in Sz meet in a line; at least three distinct planes pass through a given line. We now return to the consideration of Desargues* Theorem. In order that we may avoid the consideration of a large number of special cases it will be useful to use the terms collinear and concurrent in the following sense: three points are collinear if there is at least one line containing them (if the points coincide there may be more than one line); three lines are concurrent if there is at least one point lying on all three of them. We shall use the word triangle, or the term proper triangle, to denote a triad of points which are not collinear.
192 v. projective space: algebraic definition Two triangles ABC, A'B'C are said to be in perspective from 0 if there exists a point 0 such that OAA', OBB' and OCC are each collinear triads. We note that 0 need not be unique as, for example, when A = A'y and B = B'. With these definitions Desargues* Theorem is Theorem I. If two triangles ABC, A'B'C are in perspective there exists a line I passing through a point common to BC and B'C, a point common to CA and CfA', and a point common toAB and A'B'. The converse theorem is Theorem II. If ABC, A'B'C' are two triangles such that there is a line I passing through a point common to BC and B'C, a point common to CA and CA', and a point common to AB and A'B', then the triangles are in perspective. We remark that in both theorems the whole configuration lies in a linear space of either two or three dimensions. (Since A, B, C, are not collinear the space cannot be of less than two dimensions.) For in Theorem I there must be three lines a, b, c through O containing, respectively, AA1, BB', CC. On these lines take three points P, Q, R different from O. Then the set O, P, Q, R determines an Sr (r<3) which contains O, P, Q, R and therefore 0, A, A', B, B', C,C, and hence the whole configuration. The line I of the theorem lies in this Sr. We also note that if r = 3 the two planes determined by A, B, C and A', B', C are necessarily distinct, except in the trivial case A = A', B = B', C = C. In Theorem II the triangles A BC, A 'B'C each determine a plane. Let these be n and n'. The fine I of the theorem meets BC, CA and AB, and, since A, B, C are not collinear, at least two of these intersections are distinct. Hence I has at least two points in n, and therefore lies in n. Similarly I lies in n'. The planes n, n' having at least a line in common, their join is an Sr (r ^ 3). The point O must lie in this Sr, except in the trivial case A = A', B = B', C = C. We may therefore confine ourselves to the two cases (i) r = 3 and (ii) r = 2. In case (i) the planes n and n' determined by A, B, C and A', B', C are distinct, in case (ii) they are not. Theorem I, Case (i). The lines a, b, c containing OAA', OBB', OCC are distinct, otherwise they would all he in a plane. Hence any two of them determine a plane. The points B, C, B', C all lie in the plane determined by b and c. Therefore BC and B'C have at least one point P in common. Similarly there is a point Q common to CA
6. DESARGUES' THEOREM 193 and CA'f and a point R common to AB and A'B'. These three points P, Q, R all lie in n and n'. Now, n and n', being planes in S9i meet in a line, and this is a line I satisfying the theorem. This line I is unique. Indeed, if there existed a line V different from I satisfying the conditions of the theorem, V could not lie both in n and in n'. Suppose it did not lie in n. Then it would meet n in a point common to BC, CA, AB, which is impossible, since A, B,C are not collinear. Theorem II, Case (i). The line I must be the intersection of n and n'. Since BG and B'C' have a point in common, B, B', C, C are linearly dependent, and hence BBf and CC have at least one point in common. If they have more than one point in common the four points are collinear, and must therefore lie on the intersection I of 77 and n'. In this latter case A and A' cannot lie on I. Since AB and A'B' meet, we must have B = B'. Similarly G = C, and therefore the triangles have two pairs of corresponding vertices coinciding. When two pairs of corresponding vertices of the triangles ABC, A'B'C coincide, the triangles are clearly in perspective from any point on the join of the other two vertices. The case in which just one pair of corresponding vertices, say A and A', coincide is also easily disposed of. Then BB' and CC meet, but do not coincide, and their intersection O is a point from which the two triangles are in perspective. Finally, we have the case where no two corresponding vertices coincide, and therefore the lines AA', BB', CC are each uniquely defined, and each has a single point in common with each of the other two. Let 0 be the point common to BB' and CC. If O does not lie on A A'', the points 0, A, A' determine a plane containing A A', BB' and CC, and therefore n = n', contrary to hypothesis. Therefore A A' passes through O, and the triangles ABC, A'B'C are in perspective from O. We now turn to Case (ii), when the two triangles lie in a plane. It is desirable to deal, first of all, with certain special cases which may arise. Theorem I. If all three pairs, A and A', B and B', C and C coincide, the theorem is trivial. If two coincide, say B = B' and C = C, the line BC satisfies the theorem. If only one pair coincide, say A = A', the line joining this to any point common to BC and B'C (there is at least one) satisfies the theorem. Again, if the lines BC, B'C coincide without any pairs of corresponding vertices coinciding, BC and B'C are the only pair of corresponding sides HP I 13
194 V. PROJECTIVE space: algebraic definition which coincide. Hence AB intersects AfBf in just one point R, CA intersects C'A' in just one point Q, and the line QR satisfies the theorem. Next, suppose that the corresponding vertices A, A'; B, B'; C, C are distinct, but that O coincides with at least one of the vertices, say A. Then B'C satisfies all the requirements of the theorem. This is also true if a non-corresponding vertex, B' say, coincides with A and with O. Finally, if two vertices, say B and C, coincide, but not in 0, the lines BG, B'C must coincide, and we are back at a case already examined. There only remains the case in which the seven points A, B, C, A', B', C, O are all distinct, and corresponding sides, such as BC, B'C' are also distinct. We call this the general case. We now consider the various special cases which may arise in proving Theorem II. Theorem II. We consider the possibility of the pairs A, A'; B, B'\ C, G' coinciding. If all three pairs coincide, the theorem is trivial. If two pairs coincide, say B = B', C = C, A^A', the triangles are in perspective from any point on AA'. If only one pair coincide, say A = A', 2? =j= 2?', C =j= C", the triangles are in perspective from the point (any point) common to BB' and (7(7'. Again, if two corresponding sides coincide, say BG = B'C, the triangles are in perspective from a point on this line collinear with A and A'. Next, suppose that I coincides with BG. Then C'A' must pass through (7, and B'A' must pass through B. The triangles are therefore in perspective from A'. Finally, if two non-corresponding sides coincide, say BG = C'A', but not in I, we must have G = C, and we return to a case considered above. The only remaining case to be considered in proving this theorem is therefore the general case in which the seven lines BC, CA, AB, B'C', C'A', A'B', I are all distinct, and the pairs of points A, A'; B, B'; C, C' are also distinct. We note that all the special cases considered above illustrate the two theorems, the proofs following directly from the propositions of incidence. We now prove the two theorems in the general case on the assumption that n>2, that is, that there exists a point not in the plane n = n'. Theorem I {general case). Since there exists a point not in n, there exists a line A through 0 which is not in n. Take two distinct points V and V on A, distinct from 0. The lines VV, A A' meet in 0, and
6. DESARGUES' THEOREM 195 therefore lie in a plane. They do not coincide, because VV = A does not lie in n. Hence VA, V A' have a unique point A * in common. Since A ^A' (for we are in the general case) and VA is not in tt, A * is not in n. Nor is it on A, since this would imply A = 0 or A' = O. Similarly VB, VB' have a unique point B* in common, and VC, VC' have a unique point (7* in common. The points A*, B*, (7* do not lie in n, and they are distinct from V and V. If A*9 jB*, C* were linearly dependent, the points V, A*, U*, C* would lie in a plane which would be distinct from n. But this plane would contain A, B, C, and these would therefore be collinear points, contrary to hypothesis. Hence ABO, A*B*C* are two triangles in perspective from V. Therefore BG passes through the point P common to U*<7* and n, GA passes through the point Q common to C*A* and n, and AB passes through the point R common to A*B* and n. Similarly A'B'C and A*B*C* are in perspective from V, and we deduce in the same way that B'C, C'A' and A'B' pass through P, Q, B respectively. But P, Q, R are on the intersection I of n with the plane of A*B*C*. This line I therefore satisfies the requirements of the theorem. Theorem II (general case). Through I draw a plane n' distinct from n. P, Q, R are the intersections of BG, GA, AB respectively with I. If Q = R, then we must have A = Af, which is not the case (since we are dealing with the general case). Hence P, Q, R are all different, and we can draw through them three lines in nf to form a triangle A*B*(7*. By the converse of the three-dimensional Desargues' Theorem, ABG and ^4*iJ*C* are in perspective from some point V, and similarly A'B'C' and A*B*C* are in perspective from some point V. If V = V we should also have V = V = A* = 5*= C*9 whereas the points A*, J5*, 0* are constructed so as to be distinct. Again, neither V nor V lies in n, for if either did, we should have n' = n. The line VV is therefore uniquely determined, and meets it in one point 0. Since VA and VA' meet in A*, VV and AA' meet in 0. Similarly BB' and CG' pass through 0. Thus we have proved Desargues' Theorem and its converse, using only the propositions of incidence, except in the general case in two dimensions. We then had to make the additional assumption that the plane is contained in a space of higher dimension. This additional assumption can easily be justified for a space defined over a field K by the method of § 2, and we shall give a simple algebraic proof of 13-2
196 v. projective space: algebraic definition Desargues' Theorem in this case. But we first remark that assuming Desargues' Theorem in the plane, the converse follows without any additional assumption. We need only consider the general case. Let BC, B'C meet in P, CA, CA' in Q, and AB, A'B' in R. In the general case neither BBfR nor CCQ can be collinear triads. We consider the two triangles BB'R, CC'Q, and we see that they are in perspective from P. Now, B'R and C'Q meet only in A', and RB and QC meet only in A. Hence, by Desargues' Theorem in the plane, BBf and CC meet on A A'. That is, the triangles ABC, A'B'C are in perspective. We conclude this section with a simple algebraic proof of Theorem I which is valid as long as 0 is distinct from the vertices of at least one of the triangles. Let this triangle be ABC. We use the symbolic notation introduced in § 4. Since A' is on OA, and is assumed to be distinct from A, we may choose the multipliers of A', B', C so that A' = 0 + Aa, and similarly B' = 0 + Bfi, and C = 0 + Cy, where a, ft, y are in K. Then P = B'-C = B/3-Cy is a point on B'C and on BC. Similarly, Q = C'-A'= Cy-Ax is on CA1 and on CA, and R = A'-B' = Aol-B/3 is on A'Bf and on AB. But P + Q+R = 0. Therefore P, Q, R are collinear. This proves the theorem. 7. Some fundamental constructions. We now discuss some constructions which are of fundamental importance in the next chapter. They are valid in any Prn(K) for which n ^ 2. Let 0 and U be any two distinct points in P^(K). We suppose that the multipliers of these points are fixed, and we define a point E> which is distinct from 0 and U, by the equation E = O+U.
7. SOME FUNDAMENTAL CONSTRUCTIONS 197 Let A, B be two points of OU which are both distinct from U. We may suppose that their multipliers are chosen so that A = O+Ua, 5= O+U/3. We propose to show how to obtain the points 0+U(ot + J3), O+Uccfi by geometrical construction. In each case the construction will be given in italics, and the justification for the construction will follow. I. A construction for 0+ U(a + fi). In any plane through the line OU draw three lines a, b, u through A, B, U, respectively. These lines are distinct from OU and do not all pass through one point. Let M be the intersection of u and a, B' the intersection of u and b, A' the intersection of OB' and a, and L the intersection of UA' and b. The intersection P of LM and OU is the point 0+ £7(a + /?). [See Plate I.] By Proposition of Incidence V, if n is any plane through the line OU there exists a point X in n which is independent of 0 and U. The point Y = U + X is different from X, and does not he on OU or on AX. Hence the lines AX, BY, UX are three lines through A, B, U in n which are each distinct from OU, and do not all pass through one point. Hence the first part of the construction, the selection of the lines a, b, u, is possible.* The points M, B' are uniquely defined. So also is the point A'. For if OB' intersected a in more than one point, we should have OB' = AM = a. Then A,M,0,B' would be collinear, that is, MB' would contain 0, and we should have u = OU, contrary to hypothesis. Similarly, the point L is uniquely defined. The points L, M can only coincide if both are at the intersection of a and b. But M is on u, and u does not pass through the intersection of a and b. Hence the fine LM is uniquely defined. This line cannot coincide with OU, for M would then coincide with A, and we should have u = OU. The intersection P of LM and OU is therefore a unique point. We now show that P = 0+U{a + fi). This is evident when one or both of the points A, B coincides with O, and we may therefore confine our attention to the case represented by the first two figures. In these cases A' is easily * This was proved in § 6, using only the propositions of incidence. Here we merely select the point Y on UX.
PLATE 1 0=A=A' B=L=P 0=B A=P 0-A=B=A=L~P *
7. SOME FUNDAMENTAL CONSTRUCTIONS 199 shown to be distinct from A and M. Hence, in writing the equation M\ + A'/i+Av = 0 which expresses the collinearity of M, A', A, the multipliers A and v are different from zero, and we may choose the multipliers of A and M so that M = A+A'/i= O+Ua + A'/i. Hence M — Ucl = O + A'/i is on MU and on OAf, and is therefore the point B'. We define the symbol for Bf by the equation jB' = O + A'/i. Now B'-B= 0 + A'/i-0-Uj3 = A'fi-Up, and therefore the point L, which is on A 'U and B'B has the symbol A'/i-Up=B'-B ETftnce M-L = A+A'p-A'p+Up = A + Ufi = 0+U(a + P) is the symbol of a point on LM and on OJ7. This point is P, and we have therefore proved that P = 0+ U(oc + P). II. ^L construction for O+Uap. In any plane through the line OEU draw three lines a, 6, u through A, B, U, respectively. These lines are distinct from OU and do not all pass through one point. Let P be the intersection of u and a, Y the intersection of u and b, X the intersection of EY and a, Q the intersection of OX and b. Then PQ meets OU in the point R = 0+ Ua,p. By definition E = 0+U, and it is therefore different from 0 and U. As above, the selection of the lines a, 6, u is possible. The points P, Y are distinct, and neither lies on OU. The intersection X of EY and a is unique. For otherwise EY = a, and Y would coincide with the intersection of a and 6. But Y is on u, and a, 6, u are not concurrent. Again, X is distinct from 0, since X is on EY, which does not coincide with OU, Y not lying on OU. Hence OX is defined. The point Q, the intersection of OX and 6, is unique. For otherwise OX = 6 and we should have, in the first place, B = 0. If B =j= 0, we have a contradiction. If 5 = 0, and OX = 6, we can still obtain a contradiction. For X and Y would lie on 6, and since XY cuts OU in E, and 6 is distinct from OU, we should have E = B = 0, contrary to hypothesis. Hence Q is unique. If P coincides with Q, it can only be at the intersection of a and 6. Since P is on u, this
200 v. projective space: algebraic definition would contradict the fact that a, 6, u are not concurrent. Hence the line PQ is defined, and since P is not on OU the line PQ meets OU in a unique point R. We now prove that R = O + Uafi. [Plate II shows the possible special cases.] The proof is again trivial in the cases in which A or B coincides with 0 or E, and it will simplify matters if we confine ourselves to the cases in which A and B are different from O or E. When A is not at E the points P, A and X are all different. In the first place P is on u and not at U, hence it is not on OU, and therefore cannot coincide with A. If P = X, then E, P, Y are collinear. But PY is u which does not contain E, and we have a contradiction. Finally, since X is on a and on EY, it can only coincide with A when A = E, which is contrary to our hypothesis. Hence in the relation P\ + A/i + Xv = 0 which expresses the colhnearity of P, A, X, the multipliers A, /i, v are all different from zero. Now if B is not at O, ft is not zero, and it follows that we can choose the multipliers of P and X so that P = Afi + X. Then P+ U{l-ot)j3 = (0+ Ua)/3+ U(l-oc)]3 + X = (0+U)/2+X = EJ3+X; it follows that this is the symbol of a point on PU and on EX. Hence we may write Y=(0+U)]3 + X.
PLATE II O-A-R U <J> u o E-A-X B-Q-R U 0 E-B A=R U O-B-Q-R O E A-B R V t
202 v. projective s£ace: algebraic definition Then Y-B = (0+U)/$+X-0-U/3 = X+0(/?-l); we have a point on YB and on XO which is therefore Q. Finally, P-<2 = Afi+X-X-0(fi-l) that is, the intersection of PQ and OU, namely B, is represented by the symbol 0 + Uafi. 8. The condition for a commutative field; Pappus* Theorem. The construction given in the last section for the point B = 0 + Uafi is not symmetric in A and B, and if the roles of these two points are interchanged, the point B' = 0 + Ufa is obtained. This is distinct from JS, unless fa = a/?. In all the various cases considered in the last section, except in the first or 'general9 case, the condition aft = /?a is satisfied; for either a = ft, or one or both of a, /3 is zero or unity. We therefore confine ourselves to the first case. Fig. 2. With the same figure as for Fig. 1, let 07 meet a in P', and let XU meet b in Q'. Then P'Q' meets OU in B' = 0 + Ufa. Hence the condition that a/? = fa is the condition that the two collinear triads P'PX and QQ'Y are such that the intersection 0 of P'Y, QX and the intersection U of XQ', YP should be collinear with the intersection of PQ and P'Q'. We can express this more briefly by saying that the intersections of the cross-joins of the two triads PP'X and p p* x V Q Y Q'Q Y are collinear.
8. THE CONDITION FOR A COMMUTATIVE FIELD 203 We consider this theorem by itself, altering the notation. Let Z, V be two coplanar lines, and let ABC and A'B'C be two triads of points on I, V respectively. Let P, Q, R denote the intersections of their cross-joins BC, B'C; CA', C'A; AB', A'B. It can at once be verified that if the three points of either triad are not distinct the points P, Q, R, as long as they are determinate, are collinear. On the other hand, if one or more of the points P, Q, R is indeterminate, there is at least one line containing a point P common to BC and B'C, a point Q common to CA' and C'A, and a point R common to AB' and A'B. The theorem can also be proved at once when any one of the six points A,B,G,A',B',C coincides with 0, the intersection of Z and V. We notice further that if P, Q, R are not distinct, A9 B, C, A\ B', C cannot all be distinct and distinct from 0. If, for instance, Q = R, and A ^ A', we must have B = C, B' = C, in which case P is indeterminate, or A = B= C, withP = Q = R = A. If ABC, A'B'C are two triads of points on two coplanar lines Z and V and the intersections of the cross-joins are collinear, we say that Pappus9 Theorem is satisfied. We have seen that Pappus' Theorem is always satisfied in the case when either triad contains a pair of coincident points, or a point coinciding with the intersection O of I and V, and we need only consider the theorem when none of these special cases arises. We can prove Pappus' Theorem for the case in which the lines AA', BB', CC have a point in common. We apply Desargues' Theorem to the triangles A'BC, A B'C. We deduce that the points 0, Q, R are collinear. Similarly, by considering the triangles AB'C and A'BC we deduce that P, O, R are collinear. By hypothesis these are triads of distinct points. It follows that P, Q, R, 0 are collinear. On the other hand, if the ground field K is not commutative Pappus' Theorem cannot be true for all pairs of triads. For there must exist a pair of elements a, ft in K such that a/?=#/?a. We perform the construction (Fig. 2) given above for R = 0+ Uafi and R' = 0+ Ufia. As we saw above, if Pappus' Theorem is true for the triads PP'X, Q'QY, then R = R' and therefore a/? = /?a, contrary to hypothesis. We conclude that if Pappus' Theorem is universally true in Prn(K), then K is a commutative field. We now show, conversely, that if the ground field K is commutative, Pappus' Theorem is universally true in Prn(K). This
204 V. projective space: algebraic definition can be done by showing that any two triads can be taken as PP'X, Q'QY for a suitable Fig. 2, but the following proof is simpler. As explained above, we need only consider the case in which A, B, C, A'y B'y C are all distinct, and distinct from the intersection O of I and V. We denote the intersections of the cross-joins by P, Q, R. Fig. 3. Then, since the special cases are excluded, P, Q, R are all distinct. If QR passes through 0, the converse of Desargues' Theorem shows that the triads A'BC, AB'C are in perspective. Hence, as above, Pappus' Theorem holds for the triads ABC, A'B'C. We assume then that QR does not pass through 0, and cuts I in X, V in Y'. It can be immediately verified that since A, B, C are distinct, and since also A\ B', C" are distinct, the point X is distinct from points of the first triad, and Y' is distinct from points of the second triad. We may therefore choose the multipliers of 0, X, Y' so that A = 0 + X9 £' = 0+r. Since R is on AB' and on IF, R = A-B' = X-Y'. Now, if B = 0 + X/$, A'^O+Y'ol', A'B meets XYf in Xfi— Y'a'. Since this is the point R, we must have a' = /?.
8. THE CONDITION FOR A COMMUTATIVE FIELD 205 Again, if C=0 + Xy, C' = 0+Y'y', CA' meets XY' in Q = Xy- 7'a'. But CA meets X 7' inX- Y'y\ Since this is the point Q, we must have y'y = a'. Now consider the intersection P of 5(7' and B'G. The line J5C meets X Yf in the point Xa' - 7'y', and U'C meets X Y' in the point Xy - 7'. The point P is on X 7', that is, Pappus' Theorem is true, if and only if X«'-Yy = (X7-Y')y', that is, if and only if , , , J yy' = a = y y. But this is always the case, since K is commutative. Hence we have proved Theorem I. A necessary and sufficient condition that Pappus' Theorem be universally true in P^(K) is that K be commutative. There are, of course, other geometrical theorems equivalent to the condition that K be commutative. But since Pappus' Theorem is one of the simplest of these, and since it is traditionally taken as the geometric equivalent of the assumption of commutativity, we take it as our geometric criterion. 9. Some finite geometries. The constructions given in §7 were possible because we were able to draw a sufl&cient number of distinct lines through a given point to make the various intersections determinate. If the ground field K contains an infinite number of elements, we can draw as many distinct fines through a given point O as we please. For in any plane through O there exist two points, X, 7 such that 0, X, 7 are linearly independent. The lines joining 0 to X + 7a, where a ranges over K, are infinite in number, and we can always select from these a finite number which satisfy our requirements. If K is a finite field containing k elements (including 0 and 1) the fine XY contains the k points X+Yot (a in K) and also the point 7. Through 0 we can therefore draw exactly k+1 lines (k > 2); that is we can draw at least three distinct fines through O in the given plane. As we saw in § 7 this is a sufficient number to make the constructions given there determinate. The plane through 0 contains at least seven points, namely 0, X, 7, 0 + X, 0+7, X+7, 0 + X+Y. If K is the field of integers reduced modulo 2, in which case K
206 v. projective space: algebraic definition contains only the elements 0, 1, these seven points are the only points in the plane. Then, since o+x+x+y+o+y = 2(o+x+y) = o, the three points 0 + X, X + Y, 0 + Y are collinear. We shall return to this finite geometry of seven points in the next chapter. x x+y y Fig. 4. The number of points in P^K), where K is the field containing only the elements 0,1, is 2n+1 — 1. For in a given allowable coordinate system these are the points (a0,..., an), where each at takes the values 0, 1 independently, only the case in which every a{ is zero being excluded. The number of points in Pln(K) is the same, since Prn(K) = P'n(K). In the same way we can show that if K is the field of integers reduced modulo p, so that K contains only the elements 0, 1, ..., p-l9 the number of points in P^iK) = Pln(K) is p"*1 -1 10. r-way spaces. Many generalisations of P^iK) are possible. Only one will be used in this work. The ground field K is here taken to be commutative. We consider r projective number spaces PNJK) (i=l,...,r), and suppose that (4,...,4.) (»=l,...,r)
10. r-WAY SPACES 207 are the coordinates in allowable coordinate systems of points in these r spaces. Now consider an ordered association of these points These are called the points of the r-way projective number space An allowable transformation of PNni nr{K) is defined to correspond to a projective transformation of PNni(K), a projective transformation of PNni(K),..., a projective transformation of PNnr(K). We consider a set # of elements which can be put in one-to-one correspondence with the points of PNni # ^(K). This correspondence establishes a coordinate system in 8. The allowable transformations of PNni nr(K) lead to new correspondences, and hence to new coordinate systems in S. The elements of S, taken together with these allowable coordinate systems, constitute the points of r-way projective space Pni _ nr{K). In fact Pni _ nr(K) is simply an aggregate in one-to-one correspondence with all the sets of points (Av ..., Ar)f where Al9...,Ar are arbitrary points of P7ll(K),...,P^K), the coordinate systems being obtained by associating allowable coordinate systems in each of these spaces. *
* CHAPTER VI* PROJECTIVE SPACE: SYNTHETIC DEFINITION 1. Propositions of incidence. In Chapter V we defined a projective-space of n dimensions over a field K, and derived [V, § 4] certain theorems which we called the Propositions of Incidence. The object of this chapter is to consider to what extent these propositions characterise a projective space. We begin by postulating the existence of certain objects which we call linear spaces of 0,1,2,... dimensions (or points, lines, planes, ...) and a relation which may exist between a linear space Sp of p dimensions and a linear space Sq of q dimensions. If this relation connects Sp and Sq, we say that Sp lies in Sq, and write Sp c Sq. This relation is also written Sq^Spy and is then read 'Sq contains Sp\ If P0,...,Pp are any p+ 1 points we say that they are linearly dependent if there is a Unear space Sq of q dimensions such that PiS8q (i = 0,...,^), where q <p. Otherwise the points are said to be linearly independent. We now postulate the following properties for the incidence relation c [cf. V, §4]. I. If Sh s Sk and Sk c Sh then Sk = Sh. II. If Sp S Sq and Sq c Sr, then Sp s Sr. III. Every line contains at least three distinct points. IV. Given any p + 1 linearly independent points, there is at least one linear space of p dimensions which contains them. V. Any linear space of p dimensions contains at least one set of p + 1 linearly independent points. VI. If P0,...,Pp are p+1 linearly independent points which lie in an 8q, any Sp containing them is contained in Sq. * This chapter is almost completely independent of the rest of the book, and may be omitted at a first reading.
1. PROPOSITIONS OP INCIDENCE 209 VII. If P0,..., Pp are p + 1 hnearly independent points of an Sp, and Q0,..., Qq are q +1 hnearly independent points of an Sq, and if the p + q + 2 points P0,..., Pp, Q0> •.•,#$ are -inearly dependent, there exists at least one point R which lies both in Sp and mSr VIII. There exists a finite integer n > 1 such that there is at least one set of n + 1 hnearly independent points, but any set of m points, where m > n+ 1, is linearly dependent. That these are a set of consistent postulates is proved by the example of a projective space, defined in the previous chapter. We now draw certain conclusions about the relation ^, about linear dependence, and, later on, about the intersections of linear spaces. (1) Each linear space Sh lies in itself. For by V, Sh contains ^+1 hnearly independent points. By VI, any S'h containing these points lies in Sh. Choose S'h = Sh, and the result follows. (2) If Sp £ Sq, then p ^ q. For by V there is a set of p + 1 hnearly independent points lying in Sp. By II these he in Sq. If q<p, these points would be linearly dependent. Hence p ^ q. (3) If Sp^S'p, then Sp = Sp. For if P0,...,Pp are p+1 linearly independent points of Sp, by II they are p + 1 hnearly independent points of Sp. Hence, by VI, Sp £ Sp, and therefore, by I, 8p = Sp. Corollary. The Sp containing p + 1 linearly independent points is unique, and any set of p + 1 independent points of an Sp serves to determine it. We shall call any set of p + 1 linearly independent points in an Sp a basis for the Sp. (4) Let P0,..., Pp be p+ 1 independent points, and suppose that P0,...,Ps is a linearly dependent subset. Then P0,...,P8 lie in some 8i9 where t<s. Let Q0,..., Q( be a basis for S(, and consider the q+1 =p-s + t+l points Q0,...,Qt, P8+v,Pp- If these are hnearly independent they determine an Sq. In any case, they he in an Sr, where r^q<p. Now, 8r contains Q0>..., Qt and hence, by VI, St. Therefore, by II, Sr contains P0, ...9P8. Also, Sr contains Ps+i>->Pp- Hence P0, ...9Pp are contained in Sr (r<p) and are therefore hnearly dependent. From this contradiction we deduce that any subset of a linearly independent set of points is necessarily a linearly independent set. Corollary. The points of a hnearly independent set are all distinct. HP I *4
210 VI. projective space: synthetic definition (5) Let P0, ...fPp be a basis for Sp; Q0, ...>Qq a basis for Sq. If Sp and Sq have no common point it follows from VII that P0,...,Pp, Q0,..., Qq are linearly independent. Suppose that Sp and Sq have at least one point in common. Considering the set of common points there exists an integer r (0 < r < n) such that there is at least one set A0,..., Ar of r + 1 linearly independent points common to Sp and Sq, but any set of m points (m>r+ 1) common to Sp and Sq must be linearly dependent. Now, A0,..., Ar determine an Sr. Since Ai(^Sp (i = 0,...,r), it follows from VI that 8r^Sp. Similarly Sr^Sq. Hence r^min(p,q). Also, if the point R lies in Sr, R^Sr^ Sp, and therefore, by II, R c ^. Similarly P c ^. On the other hand, if the point B lies in Sp and in Sq, the points -40,..., Ar, B are linearly dependent, and therefore lie in an St (t^r). Since Ai^Si (i = 0,...,r), Sr^St and therefore r^t. Hence r = t and Sr = St. Therefore the points common to Sp and Sq are all the points of a linear space Sr. If now St is any space lying both in 8p and in Sq, and C0,..., Ct is a basis for St, then the points of this basis lie in 8p and in Sq and therefore in Sr. Hence St lies in Sr. (6) With the notation of (5) we prove the converse of VII, that if 8p and Sq have at least one point (and therefore all the points of an #r) in common, the p + q + 2 points P0,...,Pp, Q0,...,Qq are linearly dependent. If p > r there exists a point in Sp which is not in 8r. Otherwise Sp ^ Sr, and hencep^r. Let Br+1 be a point which is in Sp but not in 8r. Then A0,..., Ar, Br+1 must be linearly independent. Otherwise we could show, as above, that Pr+1 is in Sr. Hence A0,..., Ar, Br+1 determine an Sr+1 in Sp. If p>r+ 1 we can, similarly, find another point in Sp, Br+2, say, such that A0,..., Ar, Br+1, Br+2 are linearly independent. If we continue in this way we obtain a basis A0,..., Ar, Br+1,..., Bp for Sp whose first r-h 1 points are a basis for Sr. In the same way we can construct a basis A0,..., Ar, Cr+i> •••>£* for Sr If the Points Ao> —>Ar, Br+1,...,Bp, Cr+1, ...iCq are linearly independent, they determine a linear space of p + q — r dimensions. In any case they determine an St (t^p + q — r). Since St contains A0,...,Ar, Br+x,...,Bp, it contains Sp, and therefore, by II, P0,...,Pp. Similarly it contains Q0,..., Qq. Hence the set P0,..., PP,Q0, ...,QQ lies in St(t^p + q-r <p + ¢+1) and is therefore linearly dependent. (7) We now prove that in fact the points A0,..., Ar, Br+1,..., Bp, Cr+1,..., Gq are linearly independent. For, let us suppose that they are linearly dependent. By construction, AQ,..., Ar, Br+1,..., Bp
1. PROPOSITIONS OF INCIDENCE 211 are linearly independent and determine 8p. Also, Cr+1, ...,Ca is a subset of a linearly independent set. By (4) above this subset is a linearly independent set, and it therefore determines an Sq_+_1. By VII, Sp and 8q^r_1 have a point, R, say, in common. Since Ct ^Sq (i = r + 1,..., q), Sq_r_1^Sq, and therefore jR lies in Sq and in Sp. Hence, by (5) above, it is in Sr. Since Sr and Sq^r^1 have the point R in common, the points A0,..., Ar, Cr+1,...,Gq must be linearly dependent, by (6) above. This contradiction proves that A0,..., Ar, Br+1,..., Bp, Cr+1,...,Cq are linearly independent. These points therefore determine an Sp+q_r, the result still holding, as at the beginning of (5) above, if 8p and Sq have no point in common, provided that we put r = - 1. Clearly Sp c Sp+q_r, and Sq £ Sp+q_r. On the other hand, if Sk is any linear space such that Sp £ Sk, and also Sq^Sk, then .40,..., Ar, Br+1,..., Bp, Gr+1,...,Gq are all in /¾. Therefore Sp+q__r^Sk. Finally then, /Sp+a_r is contained in every linear space which contains both Sp and Sq. If we call this space the join of Sp and SQ, the space Sr being called the intersection of #p and $g, we have proved Theorem I. The dimension r of the intersection of two linear spaces Sp and SQ is connected with the dimension t of their join by the relation . r + t = p + q. We know, by VIII, that there exists a set of n+ 1 linearly independent points. Let these be A0,..., An. They define an Sn. If R is any other point, then, again by VIII, A0,..., An, R are linearly dependent, and therefore lie in an Sm (m^n). Since Ai^Sm (i = 0,...,n), it follows that Sn^Sm and therefore, by (1) above, n^m. Hence Sm = Sn, and all points (and similarly all lines, planes, ...) lie in Sn. The complete configuration of points, lines, planes,... lying in Sn is said to form an incidence space of n dimensions. A set of n +1 linearly independent points in Sn is called a simplex. Let A0,..., An be a simplex. The points ^0,...,^4^, Ai+1,..., An are linearly independent, and determine an Sn_1 (or prime) Zi, which we call a face of the simplex. For the sake of later applications [§ 6] we now prove that there is a point E which does not lie in any face of the simplex A0,..., An. We prove this by induction on n. If n = 1, the result is simply Postulate III. We therefore suppose that we have proved the theorem for a space of n — 1 dimensions, in particular for 27°, the 14-2
212 VI. PROJECTIVE space: synthetic definition face of the simplex which does not contain AQ. Let E12 n be a point of Z° which does not he in any face of the simplex Al9..., An of that space. Since any point common to Z° and Zi (i > 0) lies in a face of this simplex, E12 n does not he in any Zl (i > 0). Since A0 is not in 27°, the points A0 and E12 n must be linearly independent. They determine, in consequence, a line l0. This line cannot lie in 27°, since A0 is not in Z°; nor in Zl (i = 1,...,n), since E12 n does not Ue in Zl. The line meets Zl (i = 0,..., n) in unique points, namely, in the points E12 n, A0,..., A0. Now, Z0 contains at least one point E distinct from A0 and E12 n, and this point satisfies our requirements. Now let i0,...>in be any derangement of the integers 0,..., n, and consider the points Aik , ---,Ain (k^O). These are linearly independent. If, therefore, the set Aik ,..., Ain, E was linearly dependent, E would lie in the space determined by Aik+i,..., Ain, and therefore in ZV This is contrary to hypothesis. The n — k + 1 points considered therefore determine an Sn_k. Now, Aio, ..., Aik determine an Sk, and since the join of Sk and Sn_k is the whole space Sn, the spaces Sk and Sn_k have just one point in common. We denote this by Eio_ik. The points Eu_ih, Aik+l, ...,Ain are linearly independent. If Eio ik lay in any of the faces of the simplex Aio,..., Aik of Sk, say in that face determined by Ait,..., Aik, Sn_k would lie in the space determined by Ah, ...,Aik, Aijc+i,...,Ain, that is, in ZV Hence Z{<> would contain E, and this contradiction proves that EiQtik does not he in any of the (k— l)-spaces determined by the points Aio,..., A^, Aij+1,..., Aik. We obtain in this way 2n+1— 1 points Eio t ik, including Ei = At and EQ n = E. It should be noted that any incidence space of n dimensions contains such a set of points, at least, and that they are all distinct. Let us now consider the space Sk determined by Aio,..., Aik. In this space let us use the point Eio^ik, which does not lie in any of the faces of the simplex Aio...,Aik, to construct points E'JoJ8, where j&...>j8 is a subset of i0,...,ik. Arrange i0,...,ik so that (Jo> • • • >Js) = (h> • • • >h)' Then 2¾ . js is the intersection of the space determined by Aio,..., Aia with that determined by Eio _ik, Ais^,..., Ai]c. But this last space is the intersection of the space determined by Aio,...,Aik, or Sk, and the space determined by E, Ais+l, ...,Ain. It follows that E'JQ u = EJo u, so that no new points are obtained in t.hifl Txrfl.v
1. PROPOSITIONS OF INCIDENCE 213 We shall always assume that n > 1 in subsequent sections, the case n = 1 being of little interest. Indeed all that we can deduce from the propositions of incidence when n = 1 is that there are at least three points, and that two points are linearly independent if and only if they are distinct. 2. Desargues* Theorem, This theorem is fundamental for the development of this chapter. Since the proof of both Theorem and converse given in V, § 6 was based only on the propositions of incidence, except in the case n = 2, we may state the theorems without further proof for the case n > 2. Theorem I. If ABC and A'B'C are two triangles in perspective there exists a line I passing through a point common to BG and B'C, a point common to CA and C'A' and a point common to AB andA'B'. Theorem II. If ABC and A'B'C' are two triangles such that there is a line I passing through a point common to BC and B'C, a point common to CA and C'A' and a point common to AB and A'B\ the triangles are in perspective. We shall refer to these theorems as Desargues' Theorem and its converse. The theorems were only proved from the propositions of incidence in the case n = 2 for a number of special cases. The connection between the algebraic proof given in V, § 6 and the propositions of incidence is necessarily not very clear at present. But we have no a priori reasons for thinking that the case n = 2 is not a consequence of the propositions of incidence. However, we now show, by means of a counter-example, that when n = 2 the theorems cannot be deduced from the propositions of incidence. In fact, we construct an aggregate of points and lines which satisfy the propositions of incidence in the case n = 2, but for which Desargues' Theorem is not true. This construction is due to F. R. Moulton. As a preliminary, we consider a Euclidean plane referred to rectangular Cartesian coordinates (x, y), and in it a certain set of loci, which we call L-loci. An L-locus is defined by the equation y = mf(y,m)(x-a)f where m, a are real numbers, and/(i/, m) is defined as follows: (i) ifm^O, /(2/, m) = 1; (ii) if m > 0, f(y, m) = 1, when y < 0, f(y>™) = h when«/>0.
214 VI. projective space: synthetic definition We also include the case m = oo, the line x = a being an -L-locus, and m = 0, a = oo, the line y = b being an L-locus. It is clear that an -L-locus is completely determined by that part of it which lies in the region y ^ 0, or by that part which lies in y ^ 0. Moreover, if an Zrlocus has m > 0 (and is therefore not a line), the line joining any two points of it has positive slope. Bearing these facts in mind, it is easily seen that there is an L-locus joining any two points A, B, of the plane, and that it is uniquely determined by these points. In fact, if A and B both lie in y > 0 the upper half of the L-locus is the upper half-line through A and B\ and this determines the Z-locus uniquely. A similar result holds if A and B both he in y < 0. Suppose now that A lies in the region y > 0, and B in y<0. If the line AB has negative slope, the unique L-locus containing A and B is the line AB. If, finally, AB has positive slope, let A be (xl9y^j9 and let A' be the point (#1,2^). Let A'B meet y = 0 in M. Then the L-locus containing A and B is necessarily that determined by the half-line MA. We show now that any two distinct -L-loci y = rnf(y,m)(x-a)y y = nf(y,n)(x-b) (m + n) have a unique point in common. This is evident if both m, n are negative, or, by the above remarks, if one of m, n is negative. If both m, n are positive, and we solve the equations y = m(x — a), y = n(x — 6), we find that y = mn(b — a)/(n — m); and solving the equations y' = \m(x' - a), y' = \n(x' - 6), we find that y' = |mn(6 — a)/{n — w). Thus «/ gives the intersection of the L-loci if (b — a)/(n — m) is negative, and y' if this ratio is positive. This proves the result, and we deduce again that one and only one L-locus joins any two points. If m = n the -L-loci do not meet, unless they coincide. They are then said to be parallel. The process by which we pass from a Euclidean plane to a projective plane by the introduction of elements at infinity is well known. In this projective plane Desargues' Theorem is, of course, true. We can now follow a similar process to arrive at a set of
2. debauches' theorem 215 elements for which our postulates (with n = 2) are satisfied, but Desargues' Theorem is not true. To distinguish the points and lines of this space from those of the original Euclidean plane, we shall call them Points and Lines. The points of the Euclidean plane are Points. Corresponding to each set of parallel Z-loci we postulate a further Point (at infinity) which is regarded as a Point of each locus of the set. The Lines are the L-loci of the Euclidean plane together with one other element, the Line at infinity, which contains all the Points at infinity. It is now a simple exercise to verify that the Plane which contains these Points and Lines, and the Points and Lines themselves satisfy the Propositions of Incidence of p. 208. In this Plane we now construct a configuration which violates Desargues' Theorem. The construction involves certain points and lines in the Euclidean plane which are treated as Points and Lines. It is important to observe the difference between points and lines on the one hand, and Points and Lines on the other. The justification of the assertions concerning the elementary geometry involved may be left to the reader. Fig. 5. We construct two triads ABC, A'B'C for which the lines AA', BB\ CC are all parallel to 0 Y, G and C lying on 0 7, for simplicity, A on OX, and the intersections of corresponding sides lying on a
216N VI. projective space: synthetic definition line WW parallel to OX in the upper half of the plane. We also suppose that the slopes of BC, B'C, CA', CA are negative, while those of AB, A'B' are positive. Then the Lines AA', BB', CC are parallel, and therefore concurrent, the Lines BC, B'C1 intersect in U, the Lines CA, CA' in V. The Line AB is the same as the half-line AB (since A is on OX), and passes through W, whereas if A' is on the opposite side of OX to Bf, as in the diagram, the line A'B' passes through W, and therefore the Line A'B' cannot pass through W. Hence ABC, A'B'C are two triads in perspective whose corresponding sides do not intersect in the points of a Line, and Desargues' Theorem does not follow from the Propositions of Incidence when n = 2. An even simpler example than the above shows (as we should expect) that the converse of Desargues' Theorem is also not necessarily true in an incidence S2. Geometries in which Desargues' Theorem is not true have been the subject of numerous researches, but since they cannot be identified with the geometry of Chapter V we are not concerned with them here. We add to our postulates Villa. If n = 2, Desargues' Theorem is true. We only need to postulate Theorem I of the present section, and not the converse, Theorem II, since, as we saw in V, § 6, the latter can be deduced from the former. 3. Related ranges. A set of points P, Q, R, ... on a line I is called a range of points on I. A set of coplanar lines p,q,r,... through a point L is called a pencil of lines through L. From the Propositions of Incidence it follows that if 0 is any point not on I, and I' is any line in the plane joining 0 and I which does not pass through 0, V will intersect any line of the pencil 0(P,Q,R,...) in a point, giving a range of points P', Q', R', ... on V. The ranges on I and V are said to be in perspective from 0, and we write (P,Q,R,...)°(P',Q',R',...). Such a perspectivity establishes a one-to-one correspondence between the points on I and those on V. If I contains only k points, V contains only k points, and every line in the plane contains exactly k points. Through 0 there pass exactly k lines, and through every point in the plane there pass exactly k lines lying in the plane. By Postulate III, k ^ 3. We shall see that
3. BELATED RANGES 217 there are geometries in which a plane contains only a finite number of points, and our methods must be devised to take account of these finite geometries, especially when k = 3. The Propositions of Incidence assure us of the existence of 7 points in a plane in this case, and there are finite geometries in which a plane contains exactly 7 points [V, §9]. Returning to our perspectivities, suppose that another per- spectivity has been established between the points of V and those of a fine Z", the vertex being 0'. The fine V must cut Z', but need not he in the plane joining 0 to Z, which we write 0(1). We describe the perspectivities thus: (P,Q,*,...) °(P',Q',R',...) V(P»,Q»,£",...). A one-to-one correspondence, which is not necessarily a perspec- tivity, is set up between the ranges on Z and Z", and we say that these ranges are protectively, or homographically, related, or simply that they are related ranges. We write (PyQ,R,...)K(P",Q"9R\...) in this case. The process may be continued, and we say that two ranges are related if they can be put into a one-to-one correspondence by means of a finite chain of perspectivities. It is clear that in a perspectivity distinct points are projected into distinct points. Hence, if two ranges are related, distinct points in one range correspond to distinct points in the other range. An important problem in the theory of related ranges is to determine when two given ranges, which are in a one-to-one correspondence, are related by a series of perspectivities. The following theorem is a first step towards an answer. We shall use the term projectivity for series of perspectivities. Theorem I. Given two triplets of distinct points Pl9 Qv Rx and P& Q& ^3 tying9 respectively, on two distinct lines l± and Z3, there is a projectivity which assigns Pv Qv Rx to P3, Q3, j?3 respectively. Since Pl9 Qv Rt are distinct, and lx is distinct from Z3, at most one of these points lies on Z3. Similarly, at most one of the points P3, Qz, Rs lies on lv Hence we can always choose one pair of points which are not required to correspond in the projectivity, say Px and Qz, neither of which lies on both lx and Z3. Let PXQ3 be the fine Z2.
218 VI. PROJECTIVE SPACE: SYNTHETIC DEFINITION The join Qi#s is distinct from lx and Z2. On it take a point Ot which is not on either of these lines, and let 01R1 meet Z2 in R2 (Ox and Rx lie in the plane Ox(Z2)). Since Z3 and Z2 meet, PXPZ and jR2J?3 meet. Indeed, their intersection is unique, since P8#=J?3, and PXP3 has a single point in common with Z3. Let this intersection be 02. (It is not necessarily distinct from Ov) Then (Pl9Ql9B1)(^(Pl9 Q^R^iP^RJ. Pt R* it Qs Fig. 6. An immediate corollary is that a projectivity which is the result of three perspectivities, at most, can be set up between two ranges on the same line, so that three distinct, arbitrarily chosen points correspond to three distinct, arbitrarily chosen points. It may be thought that the projectivity established between Zx and Z3 is a special one, because it is the result of only two perspectivities. This is not the case, for we now prove Theorem II. If a range of points on a line lx is related by a chain of m perspectivities to a range of points on a line Zm+1 #= lv this projectivity is equivalent to at most two perspectivities. Two preliminary theorems are necessary. Consider a sequence of two perspectivities between the lines lx and Z2, and l2 and Z8. We shall call l2 the intermediary line for the projectivity between lx and lz. lt and l2 must intersect: let them meet in the point L12, and let Z2, Z8 intersect in the point L2Z; the points L12 and L2Z may or may not coincide. If they do, we have Theorem III. If lv Z2, Z8 are concurrent and lx 4= Z3 the sequence of two perspectivities between the points of lx and Z3 is equivalent to a single perspectivity.
3. RELATED RANGES 219 Let (^01,¾...)^¾¾.¾...). and (P2><?2^2,...)^23(A><?3>^3>...). There is nothing to prove if 012 = 023, so we may suppose that 0124=6)23. We consider the triads P1P2PZ> QXQ2Q& and we see that these are in perspective from L12 = L23. If both triads are proper triangles we are necessarily in the general case of Desargues' Theorem, and the sides P1P3, Qx Qz intersect on the j oin 012 02Z of the two centres of perspective at a point 0, say. Thus QXQZ, and similarly RXRZ, passes through 0, the intersection of 01202S with PXPZ. Thus If the points Pv P2, P3 are collinear, which is the case when Pt lies on 01202a> the point 0 may be taken at the intersection of QXQZ with 0l2 02Z = PXP2PZ. If both sets of points Pv P2> P3 and Qv Q2, Qz are collinear then we must have 012 = 02Z. If, on the other hand, L12 =(= jL23, we have Theorem IV. In the sequence of two perspectivities between the points of lx and Z3, the intermediary line may be replaced by any other
220 VI. PROJECTIVE SPACED SYNTHETIC DEFINITION line ZJ which joins a pair of points on lx and l3 which are neither corresponding nor coincident. Fig. 8. We apply the previous theorem twice. Let 1$ be the line joining Xx of lx to Yz of Z3. By hypothesis, Xx and Yz are distinct, and do not correspond in the projectivity between lx and ls. Let us suppose, in the first place, that Xx does not correspond to L2Zi the intersection of l2 and l3. Since L2Z is self-corresponding in the perspectivity, centre 023, between l2 and l3, we deduce that Xx L2S does not pass through 012. We take the line X1L2S as l2. It satisfies the requirements of our theorem, since Xx and L2S are not corresponding points. We can project the range (P2,...) on l2 from 012 into the range (P2,...) on V2. For l2 Ues in the plane ^23(^1) = 012(lx). By a sequence of two perspectivities, vertices 012 and 023, we have set up a projectivity between the ranges (P2,...) on V2 and (P3,...) on Z3, and the lines, V2y l2, l3 are concurrent. By the previous theorem there is a point 023 on 012 023 such that 0 Of (Pv...) _ (P2,...) _ (-Fa,...)* Similarly, since X± and F3 are not corresponding points, and Xx is self-corresponding in the perspectivity, centre 012, between lx and V2y the line XXY3 = Z£ does not pass through 023. If we now project
3. BELATED RANGES 221 the range (P2,...) on l2 into the range (P*,...) on If from the point 023, then again, by the previous theorem, since 1$, l2i lx are concurrent, there exists a point 0'12 on 012 023 and therefore on 012 023, such that The line I* therefore satisfies the requirements of our theorem, and may be used instead of l2 as an intermediary line. If Xx corresponds to L2S, the argument fails. If L12 does not correspond to F8 we can prove the theorem, as above, by taking L12YZ in place of V2 as the first intermediary line used in our argument. There remains the case in which X± corresponds to L23, and L12 corresponds to F3. We consider this case. On lx there exists at least one other point besides L12 and Xv Let Tt be such a point. If ls meets lv and Tx is the point of intersection, the point Tz of ls which corresponds to Tx is different from Tv unless the join 01202Z of the centres of perspective passes through Tv Suppose that 0120^ passes through Tl9 and that Tx = Tz. There are two possibilities: (i) There is a point Z3 on lz distinct from T3, L2Z, F3. We can pass from the intermediary line l2 to the intermediary line If by taking as intermediary lines, in order -^12^23) L12ZZy Z3Xly X± I3, none of which joins corresponding points, or passes through the intersection of lx and l3. (ii) Every line contains only three points. Then the ranges on lx and lz are in perspective from the intersection of X1X23 and YZL12, and this point is 012 = 028. The case in which 012 = 023 is, of course, trivial. Finally, if l± and lz meet, and Tz and Tx are diflFerent, or if lx and Z3 do not meet, and Tz corresponds to Tv we replace Z3 by T3 in the proof of (i), above. This concludes the proof of Theorem IV. We are now in a position to complete the proof of Theorem II. In order to do this we first consider the case of four lines lv l2, Z3, Z4, with ranges (Pl9...), (P2,...), (P3, ...)> (P4,...) on them, and suppose that (pv...)°£(p2,...), (p2,...)^8(p3,...), ^...)^(¾...).
222 VI. projective space: synthetic definition By the definition of perspective, lx and l2 are distinct intersecting lines, and similarly for l2 and Z3, and for l3 and Z4; but other coincidences of the lines lv l2, Z3, Z4 consistent with these restrictions are permissible, and must be taken into account. Fig. 9. We first prove that if lx and Z4 are distinct lines we can pass from the range (Pl9...) to the range (P4,...) by two perspectivities. We first observe that in the case in which each line of the configuration contains only three points the result follows by the argument used to prove Theorem I. Indeed, the ranges on lx and Z4 each consist of three points and are in one-to-one correspondence. Using the argument adopted to prove Theorem I it follows at once that we can pass from the range on lx to the range on Z4 by two perspectivities. In proving the result generally we may now confine ourselves to the case in which each line of the configuration contains at least four distinct points. We first consider the cases which arise when the lines ll9l2, Z3, \ have certain special mutual relationships. (i) lv l2, l3 distinct but concurrent. By Theorem III, we can pass from (Pv ...) to (P3,...) by one perspectivity, and the result follows at once. (ii) l2, lz, Z4 distinct but concurrent. This case follows in exactly the same way as (i). (iii) lx = Z3, Z4 not through L12 = L2Z. Then there is a line 1$ through L12, different from Z3, which meets Z4 in a point different from a possible intersection of l2 and Z4 and from the point of Z4 which corresponds to L12 on l2. By Theorem IV we may replace lz by 1% and then apply (ii). (iv) l2 = Z4, lx not through L2S = L34. This case follows as in (iii). (v) lx = Z3, l2 = H4. Let m be any line through L12 = L2Z = L34 different from lx = Z3 and from l2 = Z4. From a point Vx in the plane
3. BELATED RANGES 223 of lx and m we project the range (Pl9...) on lt into the range (P,...) on m. Then (p,...)^(pv...)°^(p^...)0»(pu...)°^(p1f...). Using Theorem III, we obtain, in turn, (/>,...)£* (P„...) (^ = 1,2,3,4) for suitably chosen points V2, Vz, Vt. Hence we have {Pv...)%(P,...)%(P4f...), and indeed, by another application of Theorem III, we see that the ranges on lx and Z4 are in perspective in this case. The cases considered above deal with all the cases in which lv Z2, Z3 are concurrent, or Z2, Z3, Z4 are concurrent. The remaining cases are first, the general case illustrated in Fig. 9, and the cases in which lv Z2, Z4 are concurrent (but all different) or lv Z3, Z4 are concurrent (but all different). These last possibilities cannot occur simultaneously, otherwise we should be in a case already considered. We need only consider the case in which lv Z2, Z4 are not concurrent (whether lv Z3, Z4 are concurrent or not), since the case in which lv Z2, Z4 are concurrent and ll9 Z3, Z4 are not concurrent is obtained by interchanging the roles of Z2 and Z3. Our argument, of course, includes the general case. Since lv Z2, Z4 are not concurrent we may replace Z3 by a line ZJ joining L12 to a point of Z4 not on Z2, nor corresponding, in the range on Z4, to L12 regarded as a point of Z2. We then have lv Z2, Z£ concurrent, and may apply (i). This completes our proof of Theorem II in the case m = 3. Now consider m perspectivities, that is, the general case of Theorem II. The case in which each line has only three points can be dealt with as when m = 3, and may be omitted. If, in the sequence of lines 777 7 / o\ we can find a line Zi which is different from Zi+3 we can apply the case m = 3 to reduce the number of perspectivities. We proceed in this way until either we reach the case of two perspectivities, or else k = k+s (* = l,...,m-2).
224 VI. PROJECTIVE SPACE: 8YNTHETIC DEFINITION In this case m > 3 since we have assumed lx + lm+v Then lv l29 ls are distinct lines. We can, by Theorem IV, replace l2 by a different line ZJ without altering any other l{. Then, since 1$ #= l5 we can apply the result for m = 3 to shorten the sequence of perspectivities. Thus in all cases we can reduce the number of perspectivities to two and our theorem is proved. Corollary. A one-to-one projective correspondence between two ranges on the same straight line can be obtained as the result of at most three perspectivities. 4. Harmonic conjugates. On a line I take two distinct points A and B. If C is any point of I we construct a new point D on I as follows. n. r* Fig. 10. In a plane n through I take three lines a, 6, c passing respectively through A, B, C, each distinct from I and not all meeting in one
4. HARMONIC CONJUGATES 225 point. Since at least three distinct lines pass through every point in n, it is possible to choose a, b> c to satisfy these conditions. Let b and c meet in P, c and a in Q, and let a meet b in the point P. Then A cannot coincide with P, nor P with ¢. Since A, B, P9Q are coplanar points, and not collinear, AP and BQ meet in a point. Let this be 8. Then 8 must be distinct from P, since J? is distinct from A. The line RS has a unique intersection D with Z. This is the point we seek. It is immediately verified that if C = B then D = P, and if C = ^4, then D = A. In these cases the construction of D does not depend on the particular set of lines a, 6, c chosen. We prove now that if C is distinct from A and B the point D is independent of the coplanar set of lines a, 6, c chosen: in other words, D is determined only by the ordered triad A, B, C. We first examine the case in which there are only three points A, P, C on I. Then D must coincide with one of these points. If D = A, then P, 8, Q, A must be collinear, and hence Q = S or Q = A. If Q = $, then R = P, and a, 6, c are concurrent, contrary to hypothesis; if Q = A, then C = A9 which is again contrary to hypothesis. Similarly we cannot have D = P, and so the only remaining possibility is D = C. Hence in this case too, the position of D does not depend on the particular choice of A, P, C. We may now confine ourselves to the case in which I (and hence every line) contains at least four points, and through any point there pass at least four lines which lie in any given plane through the point. We consider any plane n' (not necessarily distinct from n) through I, and in it we draw three lines a', 6', c' through A, P, C respectively, these lines being distinct from I and not concurrent. We repeat the construction given above, denoting the points corresponding to P, Q, P, 8 by P', Q'9 P', S'. Let R'S' meet ImD'. Consider, first of all, the case in which the lines a, a'; 6, 6'; c, c' are all distinct. Corresponding sides of the triangles PQR, P'Q'R' intersect in ^4, P, C, which are collinear. Hence, by the converse of Desargues' Theorem, these triangles are in perspective. If P = P', then b = 6', and c = c', contrary to hypothesis; for similar reasons Q*Q\ P4=P'. Thus the lines PP', QQ\ RR' are determinate, and they have a point 0 in common. If two of these lines coincide, say QQ' = RR\ then QR = Q'R', that is, a = a', which is contrary to hypothesis. Therefore O is determined by any two of the three lines. HPI 15
226 VI. projective space: synthetic definition Again, corresponding sides of the triangles PQS, P'Q'S' intersect in 5, A, C, and therefore these triangles are in perspective. Hence $, S' and 0, the intersection of PP' and QQ'y are collinear. The triangles QRS, Q'R'S' are therefore in perspective from 0. Let us assume, first of all, that SQ=^S'Q'. Since SQ meets S'Q' in B, and QR.meets Q'R' in A, Desargues' Theorem shows that RS and R'S' have a point in common on BA = I. Since RS meets I only in D, and R'S' meets it only in D' it follows that D = D'. If SQ = #V, but SP^S'P' we may use the triangles PRSy P'R'S' in place ofQRS,Q'R'S' and prove that Z> = 2)'. IfSQ =£'0' and SP = S'P' then 5=5' and QQ'y PP' meet in S. Since PQ£, P'Q'R' are in perspective, .RjR' goes through S = #'. Hence jR$ = R'S' meets I in the same point Z>. This proof may break down if the lines a, a'; 6, 6'; and c, c' are not all distinct. For instance, if c = c', the triangles PQRy P'Q'R' are in perspective from the intersection of RR' with c, and PQS, P'tyS' are in perspective from the intersection of SS' with c, but since PP' = QQ' we cannot deduce from these facts that QRS, Q'R'S' are in perspective. We must therefore modify our proof. We note that if c = c' the planes n, n' coincide. We show that if the lines a, a'; 6, b'\ c, c' are not all distinct there exists at least one other set of lines a*, 6*, c* through A, B, C satisfying the requirements of the construction, such that a, a*; 6, 6*; c, c*, and a', a*; 6', 6*; c', c* are distinct sets of lines. The case in which c = c' is typical. Since through each point of I there pass at least four distinct lines of n we can find lines a* and 6* through A and B respectively distinct from I, a, a' and I, b, b' respectively. Let 6* meet a* in i2*. Since c = c', there is at least one line c* in n which passes through C and is distinct from c = c', I and Oi?*. The lines a*, 6*, c* satisfy our requirements. We use them to construct a point D*. By the proof given above D = D* and similarly D' = D*. It follows once more that D = D'. This unique point Z) is called the harmonic conjugate of C with respect to A and B, and is denoted by D = (A,B)/C. It is clear from the construction that D = (ByA)/C.
4. HARMONIC CONJUGATES 227 We now prove that C = (A,B)ID = (B,A)ID. To construct the harmonic conjugate of D with respect to A and B we draw three lines AQR, BSQ, DSR through A, B,D respectively to form a triangle QSR. Then BR and AS intersect in P, and QP meets I in the harmonic conjugate of D with respect to A and B (or B and A). This is the point Gy so that the result is proved. We saw in § 3, Th. I, that a projectivity can be established between the points A, J5, C, and the points A\ B\ C, these being two triplets of distinct collinear points. We now prove that such a projectivity makes (A, B)fC correspond to (A', J3')/C". If in fact C = A or C = B then we must have C = A' or C = 2?', and in these cases the result is evident. We therefore suppose that Ay By C (and therefore A', B', C") are distinct. Since a projectivity is the result of a finite number of perspectivities, we need only consider ranges A, J5, C, A'y B\ C on distinct lines I, I', which are in perspective. (i) A = A'. Let the ranges be in perspective from R. To construct D = (A, B)/C we draw the lines ARy BB\ CB' through Ay B, C, respectively, to form the triangle B'QR. Since AB'y BQ meet in S9 RS meets I in the required point D. To construct D' = (A\ jB'J/C, we draw through A', B\ C" the lines A'QR, B'QG and C'CR respectively to form the triangle CRQ. Since A'C and B'R meet in B, BQ meets V in D'. Therefore D' = S. Hence (A, J5, C,D) and (A',B',C'yD') are in perspective from R. 15-2
228 VI. projective space: synthetic definition (ii) B — B'. This case follows at once from (i) since (A,B)IC = (B,A)IC. (iii) C = C. If AB' and A'B meet in S, and RS meets I in D, V in D', we construct (A, B)jG by means of the lines AA', BB', CA'B', and (A', B')jC' by means of the lines A'A, B'B, CAB. It follows at once that D = {A, B)jC, and D' = (A', B')IC Hence (A,B,C,D)R(A',B',G',D'). (iv) Let I and V meet in a point distinct from A,B,C,A',B',G', and suppose that (A,B,C,D)^{A',B',G',D'), where D = {A, B)/G. To prove that D' = (4', #')/£', join GA' and let this line meet ODD' in Dx, and OBB' in -Br Since (,4,5,(7,2))^(^,^,0,2^), and D = (A,B)IC, it follows from (iii) that 2¾ = (A^BJ/C, and since (i'.^C.D^^'.F.CD'), it follows from (i) that D' = {A', B')\G'.
4. HARMONIC CONJUGATES 229 Fig. 13. These results are usually stated in the following form: Theorem I. Harmonic conjugacy is protectively invariant. An important corollary is Theorem II. If there exist three distinct points A> B> C such that G = (Ay B)jC, then, given any three points A\ B'y C> with A'' 4= J3', wehaveC = {A'yB')jC'. This is evident if C" = A' or C" = B'. If C is distinct from A' and B'y there is a projectivity which takes A\ B', C into A, J5, C. By Theorem I this takes (A'> B')IC into (A, B)jG. Since, by hypothesis, this last point coincides with (7, we must have G' = (,4',£')/C". When this happens we shall say that we are in the special case. For the purpose of completing the relationship between the pairs Ay B and Gy Dy we find it convenient to define (A, B)jG when A = B. The definition varies with the circumstances. If we are in the special case we allow any point of the line to be D = (AyA)/C. If we are not in the special case we define (AyA)/G to be A when G^Ay and to be any point of the line when C = A. We now prove that if D = (Ay B)jCy then A = (GyD)jB. It may be seen at once that this follows from the definition of harmonic conjugacy, with the above convention, if the points Ay By CyD are not all distinct. We need only consider the case, therefore, in which Ay By Gy D me four distinct points on I.
230 VI. PROJECTIVE space: synthetic definition With the notation of Fig. 10 let PD meet RA in Q\ and QD meet RB in P'. The lines RD, APy BQ pass through S, so that the sides of the triangles RABy DPQ must intersect in collinear points. Therefore P'Q' passes through C. Fig. 14. To construct (C,D)/B we draw the lines DP\ CP and BPP' through D, C and B, forming the triangle PP'Q. The lines DP and CP' meet in Q', and QQ' meets I in the required point, which is therefore A. To sum up, we have proved Theorem III. The harmonic relation is symmetric inA9Band in C, D and also in the pairs {A, B) and (C, D). We may now describe (A, B) and (C, D) as harmonically conjugate pairs. The harmonic property is not of much interest in the special case, but it is of fundamental importance in all other cases. If we are not in the special case we cannot extend § 3, Th. I and find a projectivity which relates four distinct points A, JS, C, D of one range to four distinct points A', B\ C", D' of another. For if D= {A,B)/Cy but D'*(A',B')IC'9 the existence of such a projectivity would contradict Theorem I. Finally, if (A, B), (C,D) are harmonically conjugate, and 0 is any point not on the line A BCD, consider the lines a = OAy b = OB, c = OCyd— OD. If V is any line in the plane of these lines which does not pass through 0, and cuts these lines in A\ B\ C", D' respectively, then n (A,B,CyD)°(A'yB'yC'9D'),
4. HARMONIC CONJUGATES 231 and so (A\ J5') are harmonically conjugate with respect to (C\D'). We express this property of the pencil of lines a, 6, c, d by saying that (a, 6) are harmonically conjugate with respect to (c9d). 5. Two projectively invariant constructions. We re-examine the constructions already given in V, §7, and prove that they are projectively invariant. Their importance lies in the fact that they can be used to construct a field by which we shall introduce coordinates into the incidence space [§§ 7,8]. On a line I take three distinct points 0, Ey U. Let A9 B be two points of ly each distinct from U. Construction I. In a plane n through I draw three lines a, 6, u through Ay J5, Uy respectively. These lines are to be distinct from I and must not all pass through one point. Let P be the intersection of u and ay Q the intersection of u and 6, R the intersection of OQ and a9 S the intersection of UR and b. Then PS intersects I in a definite point T. After justifying this construction we shall show that T depends on Oy Uy Ay B only, and not on the construction lines a, 6, u. The first part of the construction, the selection of the three lines a, 6, u9 is possible because, as we have already seen [§3], at least three distinct lines pass through every point in n. The point Q cannot coincide with 0, and so the line OQ is defined. The point R is unique, since even when A = 0, P 4= Q. Since R is on a, R 4= £7, and therefore UR is determined. The points P and S are distinct, and their join is distinct from Z, since u does not go through the intersection of a and by and does not coincide with I. Hence PS meets I in a definite point Ty and since PS is distinct from u> T^U. The 'general' case is illustrated by Plate III, (i). We first show, however, that T is independent of the construction lines a, by u in the following special cases, in which the points 0, Ay B are not all distinct (the numbering refers to the diagrams in Plate III): (ii) A = 2? + 0. We see that T = (Ay U)/09 performing the usual construction for the harmonic con j ugate, using only lines of the figure. (iii) 0 = A. Evidently T = B. (iv) 0 = B. Here T = A. (v) A = B = 0. We see that T = 0. If in a projectivity the points 0, Ay By U become the points 0', A\ B'y U'y the point T becomes the point Tf in the above special cases, since coincident points become coincident points, and
232 VI. PROJECTIVE space: synthetic definition harmonic conjugates become harmonic conjugates. In these cases, then, the constructions are projectively invariant. We now suppose that 0, A, B, U are distinct. We prove that P, Qy By S are distinct points, no three of which are collinear. By construction, P and Q are distinct points. Also OQ=^b9 and Q does not lie on a. Hence B is not on 6, and, in particular, B 4= Q. If B = P then we should have QB = QP; but QB meets Z in 0, and QP meets Z in U. Since 0^U> it follows that jR+P. Again, since ?7#=-4, f7P#= UBy and since ¢7 + 5 the points Qy S are distinct. Finally, B is not on 6, and therefore S^B. We conclude that P, Q, 5, S are distinct points, no three of which are collinear. If B were on Z, B would be A> and OQ would be Z; that is, Q would lie on Z, contrary to our construction. Similarly, if S were on Z, S would be By and B would lie on Z, which we have seen is not the case. Hence none of the points P, Qy By 8 lies on Z. These facts will help to clarify certain points in subsequent proofs. We must show that if n' is any plane through Z (which may coincide with n) and a', &', u' are lines of n' through A, By U respectively, all distinct from Z and not concurrent, then the construction given above, with a, 6, u replaced by a', 6', v! respectively, leads to a point Tf which coincides with T. Case (i). n^n'. By the converse of Desargues' Theorem the triangles PQB9 P'Q'B' are in perspective, corresponding sides intersecting at the collinear points 0, Ay U. Similarly, the triangles SQBy S'Q'B' are in perspective. Hence there is a point common to the lines PP\ QQ\ BB'> SS', so that the triangles SPQ> S'P'Q' are in perspective. Hence SP meets S'P' on the line UB; that is, T= T. Case (ii). n = n'. If the configuration lies in a space of dimension n > 2, then the uniqueness of the point T is at once established by a double application of (i); the two constructions in n = n' are compared with a construction in a plane through I distinct from n = n'. But if n = 2, we must proceed differently. We first establish the uniqueness of T for the cases, three in number, in which two and only two of the coincidences a = a', b = 6', u = u' occur. Having done this, we observe that there are at least three lines in n distinct from Z through each of the points Ay By Uy for Z has been assumed to contain at least four distinct points 0, Ay By U. The reader may now easily verify that we can pass from construction lines a, 6, u to construction lines a', 6', u' by a series of steps in each of which only
PLATE III 0 A 0) o~A=n B~S~ T (iii) 0=B A = T Ov) 0~A~B~R-S~T (T) *
234 VI. projective space: synthetic definition one of these lines is altered. Thus we have only to consider the three following special cases (iia), (ii&), (iic). (iia). a = a', b = &', u*u'. Let PS meet QR in Y. Clearly XY is the harmonic conjugate of XU with respect to the distinct lines XAy XB. Similarly, XY' is the harmonic conjugate of XU with respect to XA, XB. Hence Xy 7, Y' are collinear points. The triangles PQYy P'Q'Y' are therefore in perspective from X. Since PQ meets P'Q' only in U, and QY meets Q'Y' only in 0, it follows from Desargues' Theorem that PY meets P'Y' in a point of OU; that is, T = T. (iifc). a = a', 6 + 6', ti = i^'. Consider the triangles QRS, Q'R'S'. Corresponding sides meet in the points U, B9 0 respectively. Hence Fig. 16. the triangles are in perspective. But QQ' meets RR' in the one point P. Therefore P, S, S' are collinear, and hence T = T'.
5. TWO PROJBCTIVELY INVARIANT CONSTRUCTIONS 235 (iic). a*a', b = 6', u = u'. The triangles PMS, P'R'S' are in perspective from Q. Since PR meets P'R' only at -4, and RS meets j?'£' only at U (if S = S', R = #', since 0* 17, and then a = a'), it follows from Desargues' Theorem that PS meets P'S'onAU. That is, T = T'. The three special cases having been examined, the uniqueness of T is now completely established. We now proceed to establish the projective invariance of the point T\ that is, we shall prove that if there is a projective correspondence between I and V such that (A9B,U909...)7i(A'9B'iU\0'9...)9 and if T' is the point obtained by applying Construction I to A\ B\ U\ 0', then (A9B9U909T9...)7l{A\B\U\0'9T'9...). We need only prove the invariance of the construction in question under a perspectivity. If n ^ 3, and the ranges on I and V are in perspective from V9 there are planes n9 n' through Z, V respectively which do not contain V. We perform Construction I on I in n9 and project the configuration in n from V on to n'. The projected figure gives Construction I on V', and therefore the projective invariance of T is proved. If n = 2 we must use other arguments. As we have already noted, we need only consider the case in which A9 B9 09 U are all distinct.
236 VI. projective space: synthetic definition We suppose that AyByOyU and A', B', 0', U' are two ranges on lines I, V respectively, which are in perspective from 7. Case (i). 0 = 0'. To construct T we use the lines A Vy BVy UB' for a, 6, u. Then, in the notation of the figure, PQ meets I in T. To construct T we use the Unes AVy BV, U'B through A', B'y U' respectively. Then SR meets V in T'. Let PB'Uy SBU' meet in X. Evidently OX is the harmonic conjugate of OF with respect to I and V. Similarly, if A'U meets A U' in 7, OF is the harmonic con- Pig. 18. jugate of OF with respect to I and V. Hence, 0, Xy Y are collinear points, and XY meets BB' in (5, B')jV. Projecting from X, we see that -XT meets AAf in (P, #)/7, and, projecting from F, it meets BB1 in (Q, 12)/7. Hence, the intersection Z of PQ, £P is on IF. Now, the corresponding sides of the triangles T'U'R, TUQ meet at F, Z, 0 respectively, and therefore, by the converse of Desargues' Theorem TT' passes through 7, the intersection of UU\ QR. Case (ii). U = U'. To construct T we use the lines AVy BVy V for a, by u respectively. If OB1 meets AV in R and UR meets BV in Sy then -4'$ meets I in T. To construct T' we use the Unes A Vy BVy I. If O'B meets 47 in R'y and £772' meets 57 in S'y then 4£' meets V in T'. Let £'12 meet A'S in X, and BR' meet 4S' in X\ Then VX is the harmonic conjugate of VU with respect to VA, VB, and so is VX'. Hence 7, JC, X' are collinear. Similarly, if BR' meets P'J? at L, and AS' meets A'S at M, we see that E7X and J7Jlf are
5. TWO PROJECTIVELY INVARIANT CONSTRUCTIONS 237 both harmonic conjugates of UV with respect to UO> UO'. Hence (7, L, M are collinear. Now consider the triangles OXT, O'X'T. Corresponding sides intersect at the collinear points M, U, L. The triangles are therefore in perspective. But 00' and XX' meet only in V. Therefore T, T', V are collinear. Case (iii). The intersection of I and I' is distinct from 0 and U. Join O'U, and let this line cut VA, VB in 4*, B*. Using 0\ A*, £*, U, we construct T*. By Case (ii) above TT* passes through F, and by Case (i) above T*T' passes through V. Hence TT' passes through Vy and the proof of the projective invariance of Construction I is complete. We shall return to a study of this construction later. We note here that given 0 and U the point T is determined by A and B independently of the order in which they are taken; that is, T is symmetric
238 VI. projective space: synthetic definition in A and B. For, having already shown that T is invariant, we may use the line TJ8R as uy and although this interchanges the parts played by A and 5, we evidently obtain the same point T in the only case we need examine, namely A + 0> B^O. Construction II. In a plane n through I draw three lines a, 6, u through A, J5, U respectively. These lines are to be distinct from I and must not pass through one point. Let P be the intersection of u and a, Q the intersection of u and 6, R the intersection of EQ and a, S the intersection of OR and b. Then PS intersects I in a definite point T. Fig. 21. After justifying this construction we shall show that T depends on the points A, B, U, 0, E, but not on the construction lines a, 6, u. As in Construction I, we can select three lines a, 6, u to satisfy the given conditions. The points P, Q are distinct, and neither lies on I. The intersection R of EQ and a is unique. For otherwise EQ = a, and Q would coincide with the intersection of a and 6, so that a, 6, u would be concurrent, which is not the case. Again R is distinct from 0, since R is on EQ, which does not coincide with Z, Q not lying on I. Hence OR is defined. The intersection 8 of OR and b is unique. For if not, we should have OR = 6, so that B = 0, and since Q lies on 6, we should also have QR = 6, so that E = B. But we cannot have E = B = 0, for E is distinct from 0. Hence 8 is unique, and it does not coincide with P, because P does not he on 6. The line PS is therefore determinate, and since P does not lie on Z, PS has a unique intersection T with I. Finally, T=# U. For if T = U, S = Q, and QR = SRy and therefore 0 = E> which is contrary to hypothesis. To show that T is independent of the plane n and of the construction lines a, 6, u, we first consider a number of cases in which A, B,U, 0> E are not all distinct. These are illustrated in Plate IV.
PLATE IV *
240 VI. projective space: synthetic definition (i) If A = 0, or B — 0, the construction shows that T = 0. (ii) If A = Ey we see that T = B. (iii) If B = Ey we find that T = A. (iv) Suppose that A = J?, but that -4 and J5 are different from 0 and JS7. Let 0' = E, U' = A = B, A' = 0, J3' = C7. Taking 0', £/', -4', B' in place of the four points 0, U, A, B of Construction I, we draw 0i2£, UQP, BSQ through 4', 5', U' respectively. Using these lines as the lines a', b'y u' of Construction I it is clear that PS meets I in the invariant point T' of that construction. Hence T = T' does not depend on the choice of lines through A, B, U, and is projectively invariant with regard to the points 0, E, A = J5, 17. Hence, to prove the uniqueness of T and its projective in variance, we may confine ourselves to the case in which 0, Ey U, Ay B are all distinct, and therefore every line contains at least five points. Then T is distinct from 0, A, B and U. As in Construction I, we show that in this case P, Q, R, S are all distinct points, no three are collinear, and none lies on I. Now let tt' be any plane through Z, and draw Unes a', 6', u' through A, J5, U respectively which are not concurrent, and do not coincide with I. Complete the construction for T'. We show that T = T'. Case (i). tt^tt'. The triangles PQR, P'Q'R' are in perspective, since corresponding sides meet at the collinear points E9 A, U. Similarly, the triangles QRS, Q'R'S' are in perspective, corresponding sides meeting at the points 0, J5, E. Hence the triangles PQS, P'Q'S' are in perspective, and therefore PS meets P'S' on the line UB. Therefore T = T. Case (ii). n = n'. If the configuration lies in a space of n>2 dimensions, we can establish the uniqueness of T by comparing the constructions with a third construction in a plane through I distinct from n = n'y using the result of Case (i). If, however, n = 2, we must proceed differently. We consider three special cases. (iia). a 4= a', b = 6', u = u'. We use the notation of the previous figure. The triangles PRSy P'R'S' are in perspective from Q. Now PR and P'R' are distinct, and meet only in A. Also RS^R'S'. For if R = R', R = R' = A> and then E = A, contrary to hypothesis. Similarly, if 8 = 8\ OS = OS' = OR = 0£', and therefore RR' contains 0, so that 0 — E, which is also contrary to hypothesis. Hence if RS = JB'/S' we must have RR' = SS', which implies E = J5, which is also not the case. Hence RS meets R'S' in the one point 0. Applying Desargues' Theorem to the triangles PRSy
5. TWO PROJEOTIVELY INVARIANT CONSTRUCTIONS 241 P'R'S', we see then that PS meets P'S' on OA. That is, T = T'. (iift). a = a'y 6 + 6', u = u'. Since 6 #= 6', ##=#', and therefore iSJ+i?'. Also, RR' = a*n = ##'. Now, g# meets #'i?' at Ey RS LP-F Fig. 23. meets R'S' at 0, and QS meets #'#' only at B. Hence the triangles QRSy Q'R'S' are in perspective. But QQ'&nd RR' meet only in P. Therefore P, $, S' are collinear points, and so T = T7'. 16 HPI
242 VI. PROJECTIVE space: synthetic definition (iic). a = a', b = 6', u^u'. Let a, 6 meet in X, let PS, RQ meet in 7, and P'S', £'Q' meet in Y'. Let i?£ meet u in F, and let R'S' meet tt' in F'. The triangles RQV, R'Q'V are in perspective, since corresponding sides meet at U, 0, E respectively. Since RR' and Fig. 24. QQ' meet only in X> it follows that W passes through X. Now, XY is the harmonic conjugate of XV with respect to XA, XB, and X Y' is the harmonic conj ugate of X V with respect to the same pair of lines. Since XV = XV, X, Y, Y' are collinear points. Now consider the triangles PYQ, P'Y'Q'. By what we have just proved, these triangles are in perspective from X. Since PQ meets P'Q' only at U, and YQ meets Y'Q' only at Ey it follows that PY meets P'Y' on Z. Hence T = T\ Having proved that T = T' in these special cases, we can proceed, as in the case of Construction I, to prove that T = T' for any two constructions in the one plane which use the same distinct points 0, Ey U,A, B. Since at least four lines distinct from I pass through each of the points A, J5, U we can pass from the first construction to the second by a series of alterations of one of the lines a, 6, u. At each change T remains invariant, since we have one of the three special cases considered above, and finally we reach the lines a', 6', uf of the second construction. We now prove that if there is a projective correspondence such that (Q,E,UyA,By...) A {0',E',U',A',B'f...)
5. TWO PROJEOTIVBLY INVARIANT CONSTRUCTIONS 243 and if T' is the point obtained by applying Construction II to the points 0', E'y U'y A'y B'y then (0, Ey Uy A, By T) A (0'y E'y U'yA'y B'y T). We need only consider the case when the two ranges are in perspective from a vertex V. We may suppose that 0, E, UyAyB are all distinct, the cases in which Ay B coincide with 0 or Ey or A = B having already been considered. We suppose in the first instance that (i) 0 = 0'. To construct Tand T we use the lines VAA'y VBB' first as a, 6, then as a', 6'. Let EA' meet VB in P. We take UP as uy meeting a in Qy and QBf then meets I in T. Let 22'.4 meet 75 in P'. We take EPP' as < meeting a' in Q\ and then Q'J5 meets VmT. Let ^, meet J5'^ in Xy BT meet 5'T in F, and let PQ meet P'Q' in Z. Finally, denote the intersections of PQ and Vy and of P'Q' and ly by #' and H respectively. Consider the triangles A'PH'y AP'H. Since corresponding sides intersect at Z, 0, X respectively we can deduce that HH' passes through V if we prove that Zy 0, X are colhnear. But this follows at once from the fact that the triangles EPUy E'P'TJ' are in perspective from Vy since corresponding sides of these triangles intersect at Zy Oy X respectively. The triangles R'QB'y HQ'B are now seen to be in perspective, vertex V. Since corresponding sides intersect at F, 0, Z it follows that 0, X, F, Z are colhnear. It is clear that OX is the harmonic conjugate of OF with respect to I and V. If TT' meets BB' in K, the harmonic conjugate of OK with 16-2
244 VI. projective space: synthetic definition respect to I and V is 0 Y, which coincides with OX. Hence OK = 0V> and therefore TT' passes through V. (ii) Now let U = V. This case can be deduced from the one just proved by observing that if we rename the points 0, E9UyA, B by ^1118 (0,*,l7,Jf5)-(l7,tf,0,B,4), and construct T from the points 0, Ey U9 A, B, we have T = T. Thus the projective invariance of T when £7 = CT is covered by the previous case. Finally, let the lines I, V meet in a point distinct from 0 and U. If Z* is the line O'TJ we proceed in two stages, first considering the ranges on I and Z*, and then the ranges on I* and l\ and the proof of the projective invariance of T is completed as in the case of Construction I. The preceding proofs of the invariance of the points T obtained by the two constructions can be greatly simplified when n = 2 if we may assume that every line contains an unlimited number of points. It is then possible to ensure that all the required inter- gections in the uniqueness proofs are determinate. Alternatively, the assumption that we are not in the special case mentioned In §4 leads to simplifications. Our purpose has been to provide proofs which are valid when n = 2 without the use of these assumptions. We conclude this section with some remarks on the various constructions we have discussed. It will be observed that in the 'general' cases of Constructions I and II a set of four coplanar points P, Qy R9 S enters, with the property that no three of the points are collinear. These points form a quadrangle with three pairs of opposite sides: PQ, BS; PS, QB\ PB, Q8. The respective intersections Z9 X, Y of these pairs of opposite sides are the diagonal points of the quadrangle, and the joins of the diagonal points are the diagonals. If we apply the usual construction for (B> S)/Z we see that this point is the intersection of B8 and X Y. Hence X, F, Z are collinear if and only if we are in the special case. Let I be any line in the plane which is distinct from the sides of the quadrangle. It meets the sides in six points which form three paira, each pair corresponding to a pair of opposite sides. We can select a set of three points from the six, one belonging to each pair, in eight possible ways. These triads fall into pairs, two paired triads
5. TWO PROJECTIVELY INVARIANT CONSTRUCTIONS 245 making up the whole set of six points. The sides of the quadrangle to which a given triad belongs either form a triangle or are concurrent. In the first case we say the triad is a triangle triad; in the second case we say it is a point triad. Thus, in the figure, A> J5, G is a triangle triad, A', B'y C a point triad. If a given triad is a triangle triad (point triad) the triad with which it is paired is a point triad (triangle triad). If two triads have one point in common they are of like type; if they have two points in common they are of opposite type. n s z v Fig. 26. It may, of course, happen that the pairs are not all distinct. Suppose, for instance, that C = C" (= Z). Then A, B, C may be regarded as both a triangle and a point-triad, and so may A', B\ G\ but A, A\ C and JS, B\ C, which also have one point in common, may be regarded as each consisting of two triangle triads. Our Constructions I and II are, essentially, constructions for completing our quadrangular set of six points, given five of them. In I, one pair consists of the point U counted twice, and another is the pair A, B. The construction is then simply a method for determining the third pair of the quadrangular set, given one of its points, namely 0. The fact that triangle triads are point triads, and conversely, explains why the construction for the point T is symmetric in A and B. In II, the given points are paired as (0, U) and (A, B), and we are given one member E of the third pair. The problem is then to construct T so that E, JS, U is a point- triad.
246 VI. projective space: synthetic definition If we denote the quadrangular set of points which occurs in II by Q(OEA9 UTB), where OEA is a point triad, UTB a triangle triad, and 0 and U> E and T, A and B are respectively the intersections with the line of the set of the pairs of opposite sides of the quadrangle PQRS, we have proved that The property of being a quadrangular set Q(OEAy UTB) is invariant under any projectivity. 6. Reference systems. In an incidence space of n dimensions there exists at least one set of n + 1 Unearly independent points. We call such a set a simplex. We saw in § 1 that there is a point E which is not in any face of the simplex A0>..., An> and we found 2n+1— 1 points Eio ik which are the unique intersection of the Sn_k determined by the points Aik ,..., Ain, E with the Sk determined by the points Aio,..., Aik. The point Eio ik does not lie in any face of the simplex Aio,..., Aik ofSk. Also, 2¾ = Aiy and EQ n = E. We now construct a further set of \n(n+\) points Fy (iy j = 0,..., n\ i 4=j). Since Ai and Ai are distinct, and Ey is distinct from each of these points, the harmonic conjugate of Etj with respect to Ax and A^ is uniquely determined. It is always different from Ai
6. REFERENCE SYSTEMS 247 and Ajy and coincides with Etj if and only if we are in the special case mentioned in § 4. We denote this harmonic conjugate by Fy = Fjt. To construct F{j consider the plane AiAiAk through AtAp and in this draw AtEjky AjAky EijAk through Aiy Ajy Etj respectively. These lines form the triangle AkEmEjk. Since AtAk and AjEijk meet in Eiky it follows that EikEjk passes through (AtA^/Ey, that is, through Fy. This is true for all values of k. Suppose, first, that we are not in the special case. Then Eip Ejky Eki are not collinear. The triangles AiAjA^ EjkEkiEi:j are in perspective from Eijk. Hence the intersections of corresponding sides are collinear; that is, Fjk9 Fki9 Ftj are collinear. If we are in the special case, Eijy Ejk, Eki are collinear, and so Fy = Eijy Fik = Ejk and Fki = Eki are again collinear. We now prove that the %n(n+ 1) points F{j all lie in an Sn__v This result has just been proved when n = 2, and we prove the theorem by induction on n. We suppose that the theorem is true for the configuration in 27°, the face opposite the vertex A0. That is, we assume that the points Ftj (t, j = 1,..., n) lie in an #n_2- If a^ the points F^ he in this Sn_2 the result is evident. Suppose that F01 is not in this #n_2. Then F01 and Sn__2 determine a prime, £^ say. Consider the point Foi. Since the p6ints Foi, Fov Fu are collinear, and Fov Fu are in Sn-1 and are distinct, the point Foi lies in Sn__x also, and the theorem is proved. It is easily seen that the points Ftj cannot all lie in an Sk (k < n — 1). For if they do, and A0 is a vertex of the simplex which does not lie
248 VI. projective space: synthetic definition in this Sk, the vertex A0 and Sk determine an Sk+V This Sk+t contains A0 and Foi, which is distinct from A0. Hence Sk+1 contains the line A0Ai9 and in particular it contains At. Therefore Sk+1 (k+l<n) contains the set A0i..., An, contradicting our assumption that these points are linearly independent. The above results still hold in the special case. The points Eio ik were constructed by means of the vertices A0,..., A n and the point E. We now prove, as a converse to the above result, that if we are given any Sn_1 not containing any vertices Ai9 we can find a point E such that all the F{j constructed from E he' in this Sn_v Let AtAj meet this Sn_1 in the point Ftj. If we show that the primes A0, ..., Aim_l9 Ai+li ..., Aj_v Aj+l9 ..., Ani Ey all meet in a unique point E. (In the special case we define Etj as the point where A^^ meets Sn_v) First consider the case n = 2. If we are in the special case, for which Etj = (AtAfilFy = Fy, we construct Ey by drawing A^^ AjAk, FtjFjkFki through Ai9 Ajf F{j respectively to form the triangle AkFkiFjk. lfAiFkj, AjFki meet in Eijk then AkEijk passes through (AtAJIFy = Etj. Since Fki = Eki, Fkj = Ekji Eijk clearly satisfies the requirements of our theorem. If we are not in the special case, Ei^Fii, and we have Ak = (44)/4 Ak = (EikFih)IAs. Hence it follows that the lines A^p EikEjk) FikFjk are concurrent, that is, EikEjk passes through Fy. Consider the triangles A^jA^ EikEkiEijt Corresponding sides meet in the collinear points Fjki Fki, Ftj. Hence, by the converse of Desargues' Theorem the triangles are in perspective from a point Eiik, and this point satisfies the requirements of the theorem. It is easily verified that if Eijk lay on any side of the triangle A^A^ the line F^F^F^ would contain a vertex of the triangle, contrary to hypothesis. We now suppose that the theorem is true for each face Sl and construct the points ^0...i-n+i...rr If * *■?"> *^e ^nes AjE0 A_1M n and ^A.-j-ii+i..." have the Point ^o...<-i«+i...j-u+i...n in com" mon. Hence the lines AiEQ *_n+1... n and Aj -¾¾... j-u+i... n intersect. This is true for all values of i and j. Since all these lines cannot Ue in a plane they must all pass through a point E. This point does not
6. REFERENCE SYSTEMS 249 lie in any of the Eiy and from E and the vertices A0,..., An we return to the points Fij9 if we carry out the constructions given at the beginning of the section. The configuration discussed above will be called a system of reference in the space. We have seen [§ 3, Th. I] that a projectivity can be established between two ranges on lines I, V so that three distinct assigned points on I correspond to three distinct assigned points on V. We also saw, later on, that this theorem cannot be extended to any four pairs of corresponding points. But we do not know, as yet, whether the projectivity is uniquely determined by the assignment of three pairs of corresponding points. We may therefore establish a correspondence between ranges on any two fines AiAi and AhAk of our reference system so that Ai9 Ap Etj correspond to Ahi Ak) Ehk respectively. We wish to do this, however, in such a way that if AtAjt AhAk and AxAm are three lines between which such correspondences are established, and Py corresponds to Phki and also Phk corresponds to Plm) then Ptj also corresponds to Plm (Pijy Phk, Plm being, of course, points on A{Ajy AhAk and AxAm respectively). We do this as follows. Let T\k denote the operation of projecting AiAi into AiAk from the vertex Fjki which is Ejk in the special case. The result of following the projection T)k by the projection 7¾ or by the projection T\ will be denoted by T^T^, or by T\kT\h respectively. We note that T\k T%c is not defined unless (i, k) = (a, b) or (i, k) = (6, a). Since the operation T\kTlki consists of projecting A{Aj into AiAk and back again, we write TjkT^ = 1, where, as is usual, 1 denotes the identical transformation. Now, we can pass from the points (A^ApE^) to the points (Ah) Ak) Ehk) by a series of projections Tgc, and we wish to show that the correspondence between the ranges does not depend on the series of projections chosen. We first consider the transformation between points on AtAj given by (i) TfaTfyTh^Hy. Let P{j be any point on AtAp and let FjkP{j meet AiAk in Pki\ let FiiPki meet AjAk in Pjk; and let FikPjk meet A^^ in P'tj. H{j is a transformation of the points of AiAi into points of AtAj9 and we see at once that RiiAi = Ap and HyAj = Ai; also H^E^ = Etj and H^F^ = F{j. We now consider
250 VI. projective space: synthetic definition the case of any other point P{j of At Ay It is verified at once that PijfPkiiPjktP'a are ^611 aH distinct. Fig. 29. Let PjciPy meet P^Py in K. We first prove that AkK passes through E{j. We may construct the harmonic conjugate of F{j with respect to Ai)Aj by taking the Unes AtAk) A^A^ F^P^P^ through Ait Ajt Ftj. It follows that AiPjk and A^PU meet on the line joining Ak to (Ai9 A^/F^, that is, to E{j. Now consider the triangles AiAjAk and PjkPkiK (which are proper triangles, even in the special case). Corresponding sides meet in Fjk> Fki9 Fijf and hence, by the converse of Desargues' Theorem, ^2¾ and A^Pki meet on AkK, which is therefore the same as AkEij. The triangles EikEjkAk and P'yPyK are in perspective, since corresponding sides intersect in the collinear points Fjki Fkii 2¾. Hence EikPq and EjkP{j meet on KAk = E{jAk. Also, the triangles EikEjkAk and PyP'yK are in perspective, since corresponding sides meet in Pjki Pki) Fip and hence E^Py meets EjkP'{j on KAk = EtjAk. From these results we deduce a simple construction for the transformation Hy: let the join of Ptj to Ejk meet AkEij in Q; then EikQ meets A{j in P^. We also deduce the following properties of Hti\ (a) H{j = Hn\ (b) (/¾)2 =1. Further, we have mi tj — mi tj —mi mk mi mi _ 17 77¾ •* jfc^tfc — i iknki — 2 jk1 ijIki1 jk — nij x iky that is, (c) ^ = ¾¾.
6. REFERENCE SYSTEMS 251 Finally, if Ah is any other vertex distinct from ^, Aj9 Aky and the plane AiAiAk is projected from the point Fhk on to the plane AiA)A> the points Ai9 Ajy Aki Pip Fjk) Fip Fik project, respectively, into the points, Ai9 Ap Ahi Pip Fjhi Fip Fih. Hence the configuration of lines in AiAjAk which determines the transformation Hy is projected into the configuration of lines in A^^A^ which determines the transformation in that plane. Since P^ is projected into P^, the transformation Hi:f is independent of the third vertex Ak by which it is defined. We may note, in passing, that if we are not in the special case, so that E{j 4= Fij9 P^ can also be defined as (Eij9 F^jPy. A second type of transformation is T)k Tkh. We prove that (ii) T\kT\n=T)h. Let Tk(P<,) = Ptt, and T{h{Pik) = Pih. We wish to prove that T^P^) = Pih. We assume that P{j #= Eijf since in that case the result is trivial. Now, the triangles P^P^P^ and E^EikEiU are in perspective from A{. Hence corresponding sides meet in collinear points. Two of these points are Fjki Fkh. Hence FjkFkh passes through the intersection of PijPih and Ei;jEih. But E{jEih and FjkFkh meet only in Fjh. Therefore PijPih passes through Fjh. That is Fig. 30. Finally, if i, j> h, k are all different, we prove that (iii) T)kT^h^ T{hT^k. Both transformations in question take ranges on A{Aj into ranges on AhAk so that the points Ai9 Eijy A^
252 VI. PROJECTIVE space: synthetic definition are transformed into the points Ah, Ehk9 Ak respectively. Let Py be any other point of A^p and let Tjk(Pij) = Poc> Tih(Pik) = Phk. If T\K(Py) = Pih9 we wish to prove that T%(Pjh) = Phk. Now, P{jPjh passes through Fih, and PikPhk passes through Fih. Hence these two lines (which are distinct if P{j^Aiy as we have supposed) lie in a plane. Therefore -¾¾ and P^hk m&&t. Since one of these lines lies in the plane A^jA^ and the other in AhAjAk, they can only meet on the line of intersection, which is given by AjAk. But since P^^Aj the line /¾¾ meets AiAk only in Fjk. Hence PjhPhk passes through Fjk. That is, T%{Pjh) = Phk. We use these three types of transformation to prove the main theorem of this section. Consider two series of T\k transformations which transform a range on ApAq into a range on ArA8 so that Apy Epv Aq transform into Ar, Er8, A8 respectively. We wish to prove that the resultant projectivities are identical. Let one series of transformations be mp ml ■*■ ql-L mn ■ • • and the other T\v Tl^n Since m>TlM,n....)-* = (... T^Tfq) it is sufficient to prove Theorem I. The resultant of any series of T)k transformations which transforms the line AaAb into itself in such a way that Aa9 Ah {and therefore E^) correspond to themselves must be the identity transformation. If there are only two transformations in the series, they must be either T%CT% or TbacTbca in order that Aa and Ab should be self- corresponding. As we have seen, each of these transformations is the identity transformation. If there are more than two transformations in the series, consider any sequence of three transformations. If two consecutive transformations in the selected three involve the same upper index we can substitute one transformation for the pair, using the result of (ii) above. Hence we need only consider a sequence such as mi rpk ml x jk x ilx km> and consider the various possibilities. (a) I = j, m = i. By (i) above the sequence can be replaced by H{i. If this is preceded by any transformation of the series, this
6. REFERENCE SYSTEMS 253 transformation must be either Tf, or 7¾. In either case we may use the result of (i) (c) to move the Hy transformation towards the head of the series of transformations. When it reaches this position it must be an Hab transformation. (6) l=j,m* i. Then T\k T% T{m = T)k T\m T% (by (iii)) = T^T?, (by (ii)). (c) If 14= j, we have, similarly, Tjk n nm=n i% nM (by m = T^T)m, (by (ii)). Hence, in all three cases, we can reduce the number of transformations in the sequence. Finally, the whole series of transformations must be reducible to one of the following forms: (H^r, (H^rTZ,, {H^YT^Tlb (<r = 0orl). Only the first of these transformations transforms Aa into Aa and Ab into Ab, and then only when a = 0. Hence we can reduce the series of transformations to the identity transformation, and the theorem is proved. This theorem enables us to establish a unique projectivity between ranges on any two lines AaAb and AtAj9 in which Aa, E^, Ab correspond respectively to Ait Eii9 A,. Such a projectivity will be used later and we shall denote it by T%. Then Tfb = 1, and Tfa = H^. 7. The algebra of points on a line. In § 5 we discussed two projectively invariant constructions. We now use these to construct a field, which will be used in a later section to introduce coordinates into our space. The constructions depend on the choice of three distinct points 0, E, U on a line I. We call the process of choosing these points the establishing of a scale on I. Our constructions determine points on I from any two points Ay J5, neither of which coincides with U. The points determined are always distinct from U. We have, therefore, two laws of composition for the points of I other than U. We consider these laws of composition separately. Construction I is symmetric in A and JS, and we denote the derived point T by A + B. We show that this law of addition is associative.
254 VI. PROJECTIVE space: synthetic definition Let A, B, Cbe three points on I. The result is evident if any one of these points is at 0, so we confine ourselves to the case in which they are all distinct from 0. Perform the usual construction for o A B TG FT u Fig. 81. T = A + B, using the notation of § 5. To construct T+C we draw TSP through T and let OS meet UQ in Q\ which is necessarily distinct from S. If OQ' meets US in S', then PS' meets I in T, Using the same figure, QS' meets I in Tf = B+C. We construct A + T' by means of the KnesARP, T'S'Q, and Oii!g. Then PS' meets i in A + T'. But this is T. Hence A + (B+C) = (A + B) + C. We have already seen that if A is any point on I distinct from U, A + 0 = A. We show now that given any two points A.BoU (distinct from U), there is a point X of I, also distinct from £7, such that A + X = B. It is sufficient to show that there is a point Y such that A+Y = 0. For if Y has this property, the point Z= Y + B satisfies our requirements. Let A' be the point (0, U)/A. Then, in the usual construction for harmonic conjugate, as in Fig. 32, the lines ARS9 A'PQ are drawn through A and A', and ORQ meets
7. THE ALGEBRA OF POINTS ON A LINE 255 them in i?, Q respectively. Also, UQ9 UB meet them in 8, P respectively, and PS meets I in A + A'. Hence A+A' = 0, and A' is the required point F. Since A + T = 0, we shall usually denote the point F by —A, and write B — A for JB + { — A). We have proved Theorem I. The points of the line I other than U form a com- mutative group under the addition law. We now consider Construction II of § 5, which also gives us a law of composition for the points of I other than U. We regard this
256 VI. projective space: synthetic definition as a multiplication, and write T = A-B. Let us prove that multiplication is associative. It is easily verified that if any one of the points A, J3, C coincides with 0 or E9 and we therefore suppose that none of the points A, B, C coincides with 0 or with E. Now, draw two lines OU' > UU' through 0 and U, intersecting in U'. Through A draw ARPy meeting OU' > UU' in R and P respectively, and let ER meet UU' in Q, QB meet OU' in S9 ES meet UU' in Q'y and CQ' meet OU'in S'. Let PS meet I in T, QS' meet Z in T'y and let PS' meet I in T. Considering the quadrangle PQSR, we see that T = ^4 • J5, and from the quadrangle PQ'S'S that T = T-C. On the other hand, the quadrangle QQ'S'S shows that T' = J5* C, and the quadrangle PgS'jR that T = ^4 • T'. Hence Multiplication is also distributive over addition. We prove that C-(A + B) = C-A + C-B. Fig. 34. In a plane through I take a point Z, and then a point Y on the line ZU. Join YB, YA and ZC. Let YE meet ZC in C", and let OC meet 75 and YA respectively in P' and P. Let ZP' and ZP respectively meet OU in B' and .4', and let UP and £/P' meet ZP' and ZP respectively in R' and R, these same lines meeting YP' and FP respectively in Q' and Q.
7. THE ALGEBRA OP POINTS ON A LINE 257 Since corresponding sides of the triangles PQRy P'Q'R' meet in the collinear points U, Z, Y, the converse of Desargues' Theorem shows that QQ' passes through the intersection F of PP' and RR'. Also, if CT, C'Z meet QQ' and RR' respectively in H and K, then since corresponding sides of the triangles YHQ' and ZKR' meet in the collinear points F> P'y C, it follows that KH passes through the intersection U of YZ and Q'R'. We can now prove our theorem. Denoting the intersections of QQ' and RR' respectively with I by S and My the quadrangle QPQ'P' shows that S = A + B, and the quadrangle G'FHK that M = C-jS = C-(A + B). On the other hand, consideration of the quadrangles C'PYZy C'P'YZ and P'RPR' respectively tells us that A' = C-Ay B' = C-B and Jf = ^' + B' respectively. We have therefore shown that C-(A + B) = C-A + C-B. In a similar fashion it may be proved that (A + B)-C = A-C + B-C. We have already noted in § 5 that AE = A = E-A. Hence the point E acts as unity under multiplication. Finally, if A is any point on I distinct from 0 (and U) we can construct an inverse A * such that The construction of A* is evident from the diagram, as is the proof of the properties of j4*. HP I 17
258 VI, PROJECTIVE space: synthetic definition These results show that the set of points on Z, excluding U> can have two laws of composition associated with them, one a law of addition, the other a multiplication law, and that under these laws the points form afield. In this field 0 is the zero, E the unity. We sum these results in Theorem II. Under the two laws of composition defined by Constructions I and II the points of the line I other than U form afield. Before passing on it is convenient to prove here a result, which we shall use later, concerning the construction of the inverse A* of A. Let RS and QP meet in Z, QS and RP in X, and let ZX meet Z in F. Since ZE is the harmonic conjugate of ZX with respect to ZP and ZRy F = (0, U)/E. Again, XS, XR are harmonically conjugate with respect to XE, XZ. Hence A* and A are harmonically conjugate with respect to E and F. In other words, we have (if we are not in the special case) Theorem III. If F is the harmonic conjugate of E with respect to 0 and U> the inverse of A may be defined as the harmonic conjugate of A with respect to E and F. Returning now to the consideration of the field we have constructed, we note that it has been defined by selecting on a line I an ordered set of three points 0, E, U. Let us denote the field which is then defined by K(0, E, U). If 0\ E\ U' be any other ordered set of points on a line Z', we may similarly construct a field K(0\ E\ U'). Now, there is a projectivity which relates a range on
7. THE ALGEBRA OF POINTS ON A LINE 259 I to a range on l\ 0, E, U corresponding to 0', E', W respectively. If points P, Q, ... of the first range (other than U) correspond to points P', Q', ... of the second range, then P + Q corresponds to P' + Q', and P* Q corresponds to P' • Q''. For we have proved that Constructions I and II yield projectively invariant points. Hence, there is a one-to-one correspondence between the elements of the fields K(0, E, U) and K{0\ E\ U') which is, in fact, an isomorphism. In particular, there is a projectivity which transforms I into itself, 0, E9 U being self-corresponding points. This defines an automorphism ofK(0, Ey U) which is the identical correspondence only if the projectivity is the identical projectivity. Since all the fields we can construct by establishing a scale on any line I in Sn are isomorphic, we can call any field isomorphic with them the field of the geometry. There is, of course, another field, for which addition is defined, as above, by Construction I, and multiplication by interchanging the roles of A and B in Construction II. This second field is the skew-isomorph [V, § 1] of the field of the geometry. 8. The representation of the incidence space as a Pn(K). We are now in a position to introduce coordinate systems into our space Sn. We first choose a reference system [§6], and use the notation already introduced. On any line AiAi there are three points Aif Aiy Ey. We choose these as 0, U, E, and define, as in the preceding section, a field by means of these points. All the fields obtained in this way are isomorphic, and the transformations T*\ defined in § 6 estabUsh, in a unique fashion, an isomorphism between the fields K(Aa, E^ Ab) and K(AC, Ecd, Ad). This means that if K denotes the field of our geometry, we have established without ambiguity, for every choice of a, 6, a correspondence between the points of AaAh (other than Ab) and the elements of K in which Aa corresponds to zero. The element of if which corresponds to the point P on AaAb will be called the coordinate of P in the scale (Aa, 2^,, Ab); or more briefly, when there is no risk of ambiguity, the coordinate of P. The coordinate of E^ is always 1. If P is distinct from both Aa and Ab, and a is the coordinate of P in the scale (Aa, Eab, Ab), the coordinate a* of P in the scale (Ab, Eba, Aa) is found as follows, when we are not in the special case. 17-2
260 VI. projective space: synthetic definition The transformation T%\ takes P into P*, where P and P* are harmonically conjugate with respect to Eab and F& [§6], and of course this transformation takes P* into P. But [§ 7, Th. Ill], if P and P* are harmonically conjugate with respect to E^ and P^, the corresponding elements of K are inverses. Since the coordinate of P* in the scale (Aai E^, Ab) is the coordinate of P in the scale (Ab9 Eba> Aa), the coordinate of P in this last scale is a* = a"1. Hence Theorem I. If P is any point of AaAb distinct from Aa and Ab, and a> a* are its coordinates in the respective scales (Aa, 25^, Ab) and (Ab9Eba9Aa)9 then act* = 1. This theorem is also true in the special case, which has been excepted from the foregoing proof. To see this we first note that in the special case the configuration of Fig. 36 has the following properties: P, E, S are collinear points, and lie on the sides of the triangle OZU, therefore OP, ZE, US are concurrent. Let the point of concurrence be T. Also, Z, Ey X are collinear. Now identify this configuration with that of Fig. 29, identifying 0, U, Z> T with Ai9 Ap Ak, Eijk respectively. Then E, P, S are identified with Eijy Ejk, Eki respectively. We saw in § 6 that if Py is any point on AtAp the point P^ which corresponds to P{j in the transformation Hy which makes Ai9 Eij9 Aj correspond to Ap Ej{, Ai respectively is obtained thus: E^P^ meets EikP'{j on AkEtp Now, returning to Fig. 36, with the above identification, the point Af corresponding to A is constructed as follows: let AP meet ZE in X> then XS meets I in A'. But this shows that A* = A'. Hence it follows that Theorem I remains true in the special case. We now consider any point P of the given space, and we suppose, in the first instance, that P is not in any face of the simplex A0,..., A n. If we use P as we used E in the constructions of § 6 we obtain a series of points P P P where Pio„.ik lies in the space of k dimensions determined by Ai9...,Aik> but not in any face of the simplex formed by these k + 1 points. Since P^ = P^ is on AiAj and distinct from both Ai and Ap it has a coordinate in each of the scales (Aiy Ey, Aj) and (Ap Ejiy A^. Let these coordinates be^ andj?# respectively. Then, by the above theorem, ViiVa = 1.
8. THE INCIDENCE SPACE AS A P^(x) 261 Theorem II.* If i, j, k are all distinct, PikPkjPji = !• Aj RH /Ejle Qij ?jk Ak Fig. 37. Let FjkPki meet A{Aj in Qkii and let FkiQki meet A,Ak in Rki. Clearly T$Pki = 2¾.^ and therefore the coordinate of Rki in the scale {ApEjk,Ak) is pik. Similarly, if P{jFki meets A^ in Q^, then TfyPij = Q#, and the coordinate of Q{j in the scale (Ap Ejk> Ak) is Pa- With these prehminary remarks we prove the Lemma. Qki Ejk intersects P{j Pjk on Ai Ak. The harmonic conjugate Ejk oiFik with respect to Ai and Ak can be constructed by drawing the lines AjAi9 AkAi and FjkQkiPki. Hence AjPki and AkQki intersect in a point X of AiEik. Now consider the triangles EjkPjkAi and Q^P^A^ Corresponding sides intersect at the points Pijk) X, A^ These points are collinear, and the triangles are therefore in perspective. The vertex of perspective Y lies on Ai Ak> and is the intersection of QkiEjk and Pi j Pjk- This proves the lemma. We now take (Ajy Ejk, Ak) as our scale on AjAk, and construct the product Rki*Pjk> as in Construction II. We use the quadrangle Fki YQkiPi:f. The line AjAk meets the pairs of opposite sides of this * We are indebted to Mr G. Kreisel of Trinity College, Cambridge, for assistance with the proofs of Theorems II and III, which do not involve the assumption of Pappus' Theorem.
262 VI. PROJECTIVE space: synthetic definition quadrangle in the pairs of points Ajy Ak; Rki> Pik\ and Eik, Q^\ and we find that consequently Therefore pikPkj = pij9 or, by Th. I, PikPkjPji x PijPji = 1. This proves the theorem. Now, if x0 is any non-zero element of Ky let ^i^Pioxo (i = !>•••> n)> («**0). PijPjoPoi = l> Pij = PioiPjo)-1* Pa = w1- The set of right-hand equivalent (n+ l)-tuples (x0,xv ...,xn) determined by different x0's are the coordinates of P with regard to the given reference system. We now suppose that P lies in the space defined by Aio,..., Aiv but does not lie in any face of the ^-simplex defined by these points. If we complete the reference system in this space by means of the point E^ Aky we obtain, as above, a set of equivalent (k+ l)-tuples {xioy ...9xik) as coordinates of P with respect to the system of reference in the fc-space. The coordinates (y0i..., yn) of P with regard to the system of reference in 8n are defined as follows: yi = xi (i = i0,...,ik), We see that this definition is consistent with our definition of the coordinates of P when P does not he in any face of the simplex AQy..., An. In fact, if the coordinates of any point P are (yQy..., yn)> a necessary and sufficient condition that y{ = 0 is that, for each j #= i, the space defined by the set of points Aq, ..., A^i, Ai+ly ..., Aj_i> Aj+1> ..., An> r intersects the line AtAj at the point Ajy or that this set is linearly Then, since that is, we see that
8. THE INCIDENCE SPACE AS A P^k) 263 dependent. For the intersebtion is the point Pijy and in the scale (Ajy Eijy AJ the coordinate of P{j is p{j = y^J1. If Pti = AJy Pij = ViVT1 = °> and therefore yi = 0. Also, a necessary and sufficient condition that P lies in the space defined by Aio,..., Aiv and not in any face of this simplex, is that the points AQy ..., A^ly Ai+i> ..., Aj_ly Aj+i> ..., An> P (^4=*o> •••>*&) are linearly independent and determine a space which intersects A{Aj in Ap where j takes the values (i0,...,ik). Also, if P is not in the space Aik^...Ain the space Aik+i ...AinP meets the space Aio...Aik in a point /^..,^ whose coordinates are (z0,..., zn)y W ^6 % = Vij U = 0,...,^), z^ = 0 0' = fc+l,...,n). We leave the proofs of these last results to the reader. We show now that any set of right-hand equivalent (n+l)4uples corresponds to one and only one point P of the space. Let (xQy..., xn) be any (n + 1 )-tuple of the set. If Xik+i = ••• = Xin = °> P must lie in the fe-space defined by Aio,..., Aik. We may therefore suppose that no xi = 0. On A^sAi take the scale (A^E^A^, and let Pi be the point on this Une which has the coordinate xtx^x in this scale. Then, by hypothesis, Pi^Ai (i = 0,...,n). The n primes defined by the points Av ...,^<-i>^i+i> ...yAn,Pi have a single point P in common. We prove this by induction on n. When n = 2, the result is trivial, since two lines intersect in a point. We assume that the theorem is true for an 8n_v in particular for that defined by AQy ...,An__v Then the n— 1 spaces S^^x defined Aly ..., A{_i, Ai+V ..., An> Pi (¾ = 1,...,tt—1) intersect in a space 8r which, by hypothesis, meets the 8n_x defined by A0y..., An_x in a single point R; for Si^x meets the Sn_x defined by^0,..., An_x in the £n_2 defined by Aly...yA^vAi+ly ...yAn_vPi. This point R does not lie in any face of the simplex AQy ...yAn_v Again, AnR does not lie in the S^t defined by Al9 ...yAn_lyPn. It therefore meets it in a single point, and this is the required point P. It is evident that the coordinates of P are (xQy ...yxn).
264 VI. projective space: synthetic definition The condition for the collinearity of three points is our next subject. We prove Theorem III. A necessary and sufficient condition that the points P = (x0,...,xn), Q = (y0> ...,yn) and R = (z0, ...,¾) be collinear is that the rank of the matrix (*o 2/o *o\ *i Vi %| (1) be less than three. If the rank is one, the three (n+1)-tuples given by the columns of this matrix are necessarily equivalent, and the three points P, Qy R coincide. If, conversely, they coincide, the rank is clearly one. We may therefore omit this case. Similarly, in proving the necessity of the condition, we may omit the case when two of the points coincide. Since the rank of the matrix is now two, at least, there are at least two linearly independent rows. Without loss of generality we may assume that the first two rows are linearly independent. A necessary and sufficient condition that the matrix is of rank two is that every other row of the matrix is linearly dependent on the first two rows. (1) We first prove that the condition in the enunciation of Theorem III is necessary. By what we have just said, it is enough to prove that if P, Q, R are distinct collinear points such that the left-hand vectors {x0,y0,z0), (xvyvz1) are linearly independent, then any other row of the matrix (1), regarded as a left-hand vector, is linearly dependent on these two. The first step is to show that this simplified form of the hypothesis implies that the line PQR does not meet the space determined by A2,...,Ai__1,AM>...yAn (i > 2). If P, Qy R all lie in this space, then *o = xi = xi = °; Vo = Vi = Vi = °; zo = *i = ** = o, and therefore the first two rows of the matrix (1) consist of zeros, which is contrary to hypothesis. Next, suppose that the line PQR meets the space determined by A2i ...,ili-i>2i<+i» ...9An in a single point. Then the Sn_2 containing this space and the line PQR meets the plane A0A1Ai in a single point, S, say. If P lies in the space A2... Ai_1 AM... An, X(i = X\ ^ x^ = u,
8. THE INCIDENCE SPACE AS A^) 265 and if it does not, P01i is determinate, and equal to 8. A similar argument can be applied to Q and to JR. Hence in all cases we have where 8 is (t0, tly 0,..., 0, tit 0,..., 0). Therefore /*o V*< and is therefore a matrix of rank one. We conclude that the first two rows of the matrix (1) are linearly dependent, which is contrary to hypothesis. Since the linear independence of (x0)yQ)zQ) and (xvyvz1) has been shown to imply that PQR does not meet the space determined by A2,A39 ...,^4i_1,^4i+1,..., Ani it follows that the collinear points P01i, Qm, R01i are distinct. To simplify the notation we denote these points by P', Q\ R'. We have now to show that the collinearity of these three points implies that the ith row of the matrix (1) is linearly dependent on the first two rows. If the line P'Q'R' lies on a side of the triangle A^A^A^ *l = Vj = ^= 0 (j = 0, or j = 1, or j = i), and the matrix /xo Vis zo\ (2) is of rank two. If this is not the case, but P'Q'R' passes through one vertex only of the simplex A^A^^A^ say the vertex Ap let us first suppose that P', Q\ R' are all different from Aj. Then Wit1 = VhVk1 = ZftZfc *> where (j9hyk) is some derangement of (0, l,i); for we obtain the same point on the side AkAh if we project P', Q' or R' from Ay Thus the matrix XjVk1 VjVk1 Zjtk1^ Wk1 VhVk1 ^1 '** xx Pi Vo Vi Vi *o «1 %ii has two rows which are linearly dependent, as left-hand vectors;
266 VI. projective space: synthetic definition its rank is therefore less than three. But this matrix is the product '*i *h <xk Vi Vn Vk zi\ zh\ ZJ fa1 P \o 0 Vk1 0 0' 0 zfc* that is, the product of the matrix (2), with its rows possibly deranged, by a regular matrix. Hence the rank of the matrix (2) must be less than three. If, on the other hand, P' = Aj9 % *h 1¾ Vi Vh Vk z\ Zh\ zj VhVk1 = /1 ° \o = zhzk Vi Vh Vk 1 > 2A «»1» **/ where and we prove as before that the matrix (2) is of rank two. We are therefore left with the case in which the line P'Q'R' does not pass through a vertex of A^AxAi9 Let V = P'Q'R meet the sides of A^AxAi in the points UQy Uv Uiy and let X be any point on V. In the first instance let X be distinct from UQi Ux and 17+. Let AQX Fig. 38. meet A1Ai in Xxi9 and AtX meet AQAX in X01. Let XliU1 meet A0At in N, EnUi meet A0AX in L, and FoiL meet AxAi in M. Then, in this construction, the points UQ, Uly Ui9 L> M depend only on V, and not on the point X chosen on it.
8. THE INCIDENCE SPACE AS A P^K) 267 Taking the scale (Av EU,A^ on A1Ai we construct the product Xu as follows. Through M, Xu, Eu we draw the lines MF^, XliU1 and EliU1 respectively. Then, by the usual construction, F^N meets Ax At in the required point S = M #XU. If we denote the elements of the ground field K corresponding to the points Y of A1Ai by y we have _ Passing from AXAX to AXA0 by the transformation T{q> it follows that in the scale (AVE109A0) the product L*X'U = N. If, in the same scale, we construct N + X01, using the quadrangle Xf71u4iXli, we see that JV + X01 = Ut. That is It follows, that if X is the point (£0,..., £n), wg< + £0 = *<£i- (3) We note that this formula holds when X is at Z70, C^ or C^. For, if X — U0) N = U{ and £#X^ = Ui9 so that m^ = u^l9 and £0 = 0. If X = Uiy £< = 0, and ut = x01. Finally, if X = Ul9 let ^¾ meet AqAx in £7^. Then L and U'x are harmonically conjugate with respect to AQ and Av But the elements of K which correspond to L and to U[ in the scale (Al9 E1Q, AQ) are m and wx = xoi respectively. Hence m = — a^, and £x = 0, so that once again the formula is satisfied. Since the formula (3) is true for all points on Z', it is true, in particular, when X coincides with P', Q' and R' respectively. The rows of the matrix (2) are therefore linearly dependent, and so once again it is of rank two. Applying this result for i = 2,..., n we deduce that if P, Q9 R are collinear the rows of the matrix (*o Vo *o\ «i Vi *i I *n Vn Zn/ are all linearly dependent on the first two rows. The rank of this matrix is therefore two. We have proved the necessity of our theorem, and we now prove its sufficiency. (2) Let us suppose then that the rank of the matrix is less than three. We must show that P, Q, R are collinear. If the (n -f l)-tuple defined by any column is equivalent to that defined by any other
268 VI. projective space: synthetic definition column the points P, Q, R are not distinct, and the theorem is trivial. We need only consider, therefore, the cases in which P, Q, R are distinct. The point P = (x0,..., xn) cannot lie in every face of the simplex. Suppose, for instance, that it does not lie in the face A1...An. Then x04=0. A necessary and sufficient condition that P, Q9 R are collinear is that PQ and PR meet the face At... An in the same point. If PQ, PR meet this face in Q'y R', then by the first part of the theorem, proved above, Qf is linearly dependent on P, Q and satisfies the equation y'0 = 0, if we suppose that Q' = (y'0,...,y'n). Hence, by (1) above, Q' = (0,2/i, ..-,2/^) = (Ofy1-x1x^1y09...fyn-xnx^1y0). Similarly, R = (0,2^...,2^) = (0,z1 — x1x0 z0> ...,zn — xnx0 z0). Now, the matrix 0 0\ = /x0 yQ z0\ /1 -Xoly0 -x^z, Vi V'n 0\ 1 0 0 1 is the product of the original matrix by a non-singular matrix, and therefore its rank is equal to that of the original matrix. By hypothesis, this rank is less than three. The columns of the first matrix are therefore linearly dependent. Since x0 4= 0 this can only be the case if the submatrix is of rank one, that is, if Q' = R'. Therefore P, Qy R are collinear, and both parts of Theorem III are proved. Theorem IV. A necessary and sufficient condition that the points Pi = (*o**i* •••fSn*) (* = 0,...,r) be linearly dependent is that the rank of the matrix (4) I * f \xn0 . xnr/ be less than r+1. xoo #10 iXn0 • • • • #0r x1r • xnrt
8. THE INCIDENCE SPACE AS A i^(Jl) 269 We have proved this theorem when r = 2, and we therefore assume that it is true for r points, and use the principle of induction. (1) Necessity. Suppose that the r+ 1 points are linearly dependent. If any set of r of these points is also linearly dependent then, by the hypothesis of induction, the corresponding columns of the matrix (4) are linearly dependent. The rank of this matrix must then be less than r+ 1. If no set of r points is linearly dependent, the first r points of the set of r + 1 points determine a space Sr_x which contains the remaining point Pr. In this #,._! the points Pv ...,Pr_x define an #r_a which is met by the line P0Pr in a point jB, say. By the induction hypothesis we may write, with the usual symbolic notation [V, § 4], £=¾^ and also R = — P0 A0 — Pr A,. r Hence, 2 JFJ A^ = 0, and this means that the columns of the matrix o (4) are linearly dependent, and therefore the rank of (4) is less than r+l. (2) Sufficiency. Let the symbolic relation 2^^- = 0 express o the linear dependence of the columns of the matrix (4). There must be at least two A^ which are not zero. If there are only two, the corresponding Pt coincide. The theorem is evident if any of the Pi coincide. If more than two Xi are not zero, say A0, A1? A2 are not r zero, then the points P^k^PxXXy — S^A^ are coincident. Hence 2 the points P0, Px and P2, ..., Pr define intersecting spaces, and therefore by § 1, (6), p. 210, the pointsP0,Pl9 ...,Pr, are linearly dependent. Just as in Chapter V we deduce that the points of a A;-space Sk are precisely those which satisfy n — k linearly independent left- hand linear equations. We have now proved that an incidence space Sn has the property that its points can be put in one-to-one correspondence with the points of a projective number space PNrn(K) in such a way that the points of Sn lying in an Sa correspond to the points of PNrn(K) which satisfy n — a linearly independent left-hand linear equations. Nowlet y = Ax, x = By (AB = /w+1) be a projective transformation of PNrn(K). This gives us a new
270 VI. projective space: synthetic definition correspondence between the points of Sn and the points (y0>...,yn) of PNrn(K). A necessary and sufi&cient condition for the collinearity of three points of Zn is still that the corresponding matrix of the y coordinates be of rank less than three, for the rank of a matrix is invariant under a non-singular transformation. Now let A\ (i = 0,...,n) be the points of Sn corresponding to those of PNrn(K) with coordinates (8i0,..., Sin), and let E' be the point in En which corresponds to (1,1,..., 1) in PN?(K). Then we can construct a new reference system in En, using A'0>..., A'n, Ef in place of A0i..., An> E. Let P' be any point of Sn which is not in any face of the new simplex, and let the corresponding point in PNrn(K) (more briefly, its y-coordinates) be (y0,...,yn). We show that we can oonstruct a coordinate system for Zni using the methods described above, in which the coordinates of P' are also (y0,..., yn). First of all it is evident that if the prime A' A* A* A' A' A' P' meets the line A \ A J in P'ijy then the y-coordinates of Py are (0,...,0,^,0,...,^,0,...,0) = (0,...,0,^^0,...,1,0,...,0). There is, indeed, a one-to-one correspondence between the points of A'iA'j (other than A^) and the field Ky given by the y-coordinates of the points, in which A\ corresponds to 0, and Ey corresponds to 1. Let us now consider the efifect of the fundamental constructions on the points of A^A'p choosing (A'i9 E'ij9 Aj) as the scale. By V, §7, we see that if the points Q> R of the line correspond respectively to the elements q> r of K, the points Q + R and Q • R correspond to the elements q + r and qr of K. Hence we can set up an isomorphism between the field determined by choosing the scale (A'i9 Ey, Aj) on A\AJ and the field K. Now, if the y-coordinates of any point Q' on A\A^ are (0,...,0,^,0,...,1, 0, ..., 0), let R' be the point on A\A'k with ^-coordinates (0,...,0,^, 0,...,0,..., 1,...,0). Then Q'R' meets A\A'k in the point F'ik with coordinates (0,...,0,0,..., 1,...,-1,...,0)#. From these results it follows that if we choose the scale (A'0, E'01i A[)
8. THE INCIDENCE SPACE AS A ^(k) 271 on one line, A'0A[ say, and construct our field as in §7, the isomorphism between the field so constructed and the field K of the geometry can be set up so that the constructed coordinates of P'ol are the same as its ^-coordinates. We may then use the transformations T\k to assign coordinates to points on any edge A\A'm of the simplex of reference, and the constructed coordinates of these points are also the same as their y-coordinates. Hence the constructed coordinates of P' are the y-coordinates of P' for all points P'. Our method of introducing coordinate systems into Sn allows us to establish a one-to-one correspondence between the points of En and those of a right-hand projective number space PNrn(K)> and, furthermore, as we have just seen, we can construct a new coordinate system corresponding to each projective transformation in PNrn(K). But we cannot say, conversely, that every coordinate system obtained by the methods of this section necessarily arises from the original one by means of a projective transformation of PNrn(K). In fact, let us consider the coordinate system (xQ)..., #n) constructed at the beginning of this section. This gives, of course, an isomorphism between the points P of AQAX other than Ax and the elements p of the field of the geometry, in which AQ and EQl correspond to 0 and 1 respectively. There may, however, exist other isomorphisms between the points of AQAX other than Ax and the elements of K satisfying the same conditions, in which the point P corresponds to the element p of K, where p is not always the same as p. This isomorphism maps AQ on the element 0, EQ1 on the element 1, the point P + Q on p + q, and P-Q onpq. If we now construct a coordinate system as above, using the T)k transformations, we may obtain a coordinate system for the incidence space which is not obtainable from PNrn(K) by a projective transformation. For example, if K is the field of complex numbers, and p denotes the conjugate complex of p> the coordinates of a point (xQi...,xn) are (xQ,...,xn) in the new coordinate system. This transformation is not a projectivity, since it transforms the line x2 = xz =... = xn = 0 into itself and every real point of this line is a united point, whereas a projectivity on a line has only a finite number of united points. Thus the space Zn can be represented as a projective space PJ^(X), and our method of establishing coordinates in Zn yields all the allowable coordinate systems of P^iK), but, possibly, other coordinate systems as well.
272 VI. projective space: synthetic definition The reader may ask how, starting from the Propositions of Incidence, in which there is no mention of right or left, we arrive at a right-hand projective space, rather than a left-hand space. In particular, how is it that a left-hand projective space Pln(K), which satisfies the Propositions of Incidence, can be represented as a right-hand projective space? The answer lies in our choice of the definition of A-B in §7. We might have interchanged the parts played by A and B in the construction, and so arrived at the skew- isomorph of the field of the geometry. Using this field, we should have obtained a representation of En as a Ze/f-hand projective space. 9. Restrictions on the geometry. We have shown that a space defined by the Propositions of Incidence can be represented as a projective space, as defined in Chapter V, and have, in fact, shown that the definitions of that chapter are identical with those of our present chapter. For the further development of the properties of a projective space it is now convenient to impose certain further conditions on the space. The imposition of these conditions is equivalent to restrictions being imposed on the field of the geometry. A. The field K is to be commutative. There are several ways in which this condition can be expressed. By V, § 8, we see that an equivalent form of the restriction is to require that Pappus' Theorem is true in the geometry. In our next section we shall discuss alternative forms of the same restriction. In the meantime we prove Theorem I. If Pappus' Theorem holds in an incidence space of two dimensions, then Desargues9 Theorem is a true theorem in that space. After the discussion in V, § 6, it is sufficient to consider the' general' case, in which the vertices of the two triangles ABC, A'B'C and the centre of perspective 0 are all distinct points. Let the intersection of corresponding sides of ABC and A'B'C be P, Q, B. Let QR cut OB in P', and let AP' meet A'C in L and OC in M. Also, let LB' cut AB in N. Now consider the two triads of colHnear points OAA' and LB'N. By Pappus' Theorem ON and A'L intersect on P'R = QR. Hence 0, Q, N are collinear. Next, consider the triads OMC, ABN. By Pappus' Theorem MN and BC meet on QR. Finally, consider the triads OMC, LB'N. By Pappus'
9. RESTRICTIONS ON THE GEOMETRY 273 Theorem B'C and MN meet on QR. Therefore BG and B'C pass through the point common to MN and QR. That is, Desargues' Fig. 39. Theorem is true. In place of VIII (a) (p. 216), which postulates the truth of Desargues' Theorem forn = 2 (in the plane), we substitute a new postulate: IX. Pappus' Theorem is true. B. The field K is to be without characteristic. Let (0, E, U) be a scale on a line I, and define the set of points El9 E2,... (Fig. 40) by the relations ^ = E^± + E^ Ei = Em Fig. 40. If the field is of characteristic p, the points 0,EV ...,Ep_x are all distinct, but Ep = 0. HP I 18
274 VI. projective space: synthetic definition In particular, if p = 2, E% = 0, and therefore -E ~ E. That is, E = -E = (09U)IE, (1) and we have the special case discussed in § 4, in which the diagonal points of a complete quadrilateral are collinear. Conversely, if we are in the special case the relation (1) holds, and E2 = 0, so that the field K is of characteristic 2. We now add the postulate X. The points Ey E2iEz>... are all distinct. Finally, we sometimes wish to impose the condition that the field K be algebraically closed. The incidence space Zn can be represented as a projective space Pn(K). We know [III, §4, Th. II] that there is a unique field K which is the intersection of all algebraically closed fields containing K. We can embed Pn(K) in Pn(K) [V, §3]. Hence En can be embedded in an essentially unique space En whose field is K. We call Tn the algebraic closure of En. Our final postulate may then be stated as: The space Zn coincides with its algebraic closure. More simply we shall usually say XI. The field of the geometry (or ground field) is algebraically closed. In the rest of this work we shall always assume Postulates I-X. We shall state at the beginning of each chapter whether Postulate XI is to be assumed. 10. Consequences of assuming Pappus5 Theorem. It has already been remarked that there are several geometrical assumptions which are equivalent to the assumption that the field of the geometry is commutative. Our choice fell on Pappus' Theorem because it is possible to state it without a considerable amount of preliminary explanation. A more fundamental result, which is equivalent to the condition that the field is commutative (and hence to Pappus' Theorem), concerns projectivities between ranges. We have seen [§ 3] that it is possible to establish a projectivity between two ranges so that three distinct points of the one range correspond to three distinct points of the other. On the other hand, if we are given four points Ay By Cy D Of 0^ TBJigOy SiTid fOXXT pOUltS A'y B'y C'y D' Of tlte Otl^,
10. CONSEQUENCES OF ASSUMING PAPPUS* THEOREM 275 there may be no projectivity in which A, B, 0, D correspond, respectively, to A', B', C, D'. We now prove Theorem I. A necessary and sufficient condition that a projectivity between two ranges be uniquely determined by the assignment of three pairs of corresponding points is that Pappus' Theorem be true. (1) Necessity. Suppose that a projectivity is uniquely determined when three pairs of corresponding points are given. Let ABC, A'BfC be two triads of points on two lines I, V which intersect at 0. Let B'C, BC meet in P, CA', G'A meet in Q, and A'B, AB' meet in R. (a) If the triads are in perspective from a point V then clearly the lines OP, OQ, OR all coincide with the harmonic conjugate of OV with respect to I and V. Then P, Q, R are collinear, and Pappus' Theorem is true without any further condition. Suppose, next, that the line joining two of the points P, Q, R, say Q and R, passes through 0. Then if BB' meets A A' in V3, OV3 is the harmonic conjugate of OR with respect to I and V. If CC meets AAf in V2, OV2 is the harmonic conjugate of OQ with respect to I and I'. Hence OVz = OV2, and therefore Vz = V2, the triads are in perspective, and P lies on QR. (b) If the triads are not in perspective, we use the hypothesis that a projectivity is determined by the assignment of three corresponding pairs of points to deduce Pappus' Theorem. Let QR meet I in L, and V in M'. Then the points L, Mf, 0 are all distinct, by what we have just proved. Let A A' meet QR in S. We denote the line LM'QR by I*. By hypothesis there is a unique projectivity between the ranges on I and V in which A, B,G correspond, respectively, to A', B', C. This projectivity may therefore be defined by the perspectivities (A,B,C,...)^(S,R,Q,...)^(A',B',C,...). Now, regarding 0 as a point of I, there is a unique point of V corresponding to it in the range on V. Passing from I to V via I* by the above perspectivities, 0 becomes M', then M'. Similarly, the point L of I corresponds to 0 of V. Since, by hypothesis, the projectivity is uniquely determined by the pairs A, A'; B,B' and C,C, the points L, Mf are uniquely determined by the projectivity. The line Z* is therefore uniquely determined by the projectivity, and this line contains the points Q, R. 18-3
276 VI. projective space: synthetic definition Now, if we choose B> B' as centres of perspective, we find that the line RP cuts Z in the point L and V in the point M'. That is, R and P also lie on Z*. Hence, P, Q, R are collinear, and Pappus' Theorem is true. Fig. 41. (2) Sufficiency. We now suppose that Pappus' Theorem is true. Before proving our theorem we must prove a Lemma. If Z, V are distinct lines intersecting in O, and if ranges (P, Q, R, 0,...), (P', Q', R',0,...) on I and V are protectively related so that 0 on I corresponds to 0 onl', then the ranges are in perspective. It is evident that if the ranges are in perspective 0 on Z corresponds to 0 on V. Now, we can pass from the range on Z to that on V by two perspectivities at most [§3]. There is nothing to prove if we can pass from one range to the other by a single perspectivity. Let there be two, then, and denote the intermediary line by Z*. If Z* passes through 0, the ranges on Z and V are in perspective [§ 3, Th. III]. We suppose, then, that Z* meets Z in P, V in Q\ and that P, Q', 0 are distinct points. We pass from the range on Z to a range on Z* by a perspectivity with vertex L, and then from the range on Z* to that on V by a perspectivity, vertex N. It is easily verified that P and Q' do not correspond in the projectivity. If L = N the ranges on Z, V are in perspective, vertex L = N. If L #= N, the line LN must pass through 0, since 0 on Z corresponds to 0 on V. Let LQ9 meet lmQyNP meet V in P\ and let LQ' meet NP in M. If R is any other point of Z, Rf the corresponding point of the range
10. CONSEQUENCES OF ASSUMING PAPPUS* THEOREM 277 on V, LR and NR' must meet in a point JS* of Z*. Now, applying Pappus' Theorem to the triads LON, PR*Q', we see that RR' Fig. 42. passes through M. Since R, R' is any pair of corresponding points, it follows that the ranges are in perspective from My and the lemma is proved. We now return to our main theorem. Let (P,Q,R,...), (P', Q'y R',...) be ranges on two distinct lines ly V, and suppose that there are two projectivities 77! and 772 between the ranges which Fig. 43. make P, Q, R correspond to P', Q', R' respectively. If a pair of corresponding points, say P and P', coincide at the intersection oi I and l\ then, by the lemma, 77! and 772 are both perspectives with the intersection of QQ' and RR' as vertex of perspective. In thig case nx = 772. Since the three pairs of corresponding points are
278 VI. projective space: synthetic definition assumed to be distinct pairs we can select points on I and V, say Q and P', which are both distinct from the possible intersection of I and Vy and do not correspond. We can now pass from the range on I to the range on V by two perspectivities in which QR' is the intermediary line I* for both projectivities II1 and 772 [§ 3, Th. IV]. Let the vertices of perspective for TI1 be 0 and 0'. If X is any other point of I, X' the point on V corresponding to X in IIl9 then (P, Q, P, X) ° (P*, Q\ B\ X*) ^ (P\ Q\ R\ X'), where #* = Q and R* = R'. Now let -X" be the point of V corresponding to X in 772, and let O'X meet Z* in Z*. Then (P*, Q*f i2*, X*) ^ (P', Q', P', X) A (P, Q, B, X) (in 772). Since Q* = Q, the ranges (P*, Q*,P*,X*), (P,Q,P,X), by the lemma, must be in perspective. The vertex must be 0. Hence 0, X, X* are collinear points. That is, X* = X*, therefore X' = X, and hence nx = 772. If the ranges (P, Q, P,...), (P', Q', P',...) are on the same line lf we project the second range on to a line V, obtaining a range (PvQvRl9...). If there were two distinct projectivities from (P, Q, P,...) to (P', Q\ R\ ...) there would be two distinct projectivities from (P, Q, P,...) to (Px, QVRV ...)> and we have just shown that this is not so. This concludes the proof of Theorem I. We now consider certain particular projectivities on a line I. In Construction I of § 5 for A + B let us keep A, 0, U fixed, and also the lines a, u. This fixes P, and we suppose that Q is also fixed.
10. CONSEQUENCES OF ASSUMING PAPPUS' THEOREM 279 R is then fixed, and b is defined as QB. As B describes a range (B) on OU, S describes a range (S) on UR such that (S) % (B). Similarly, (T) £ {S)f and therefore (T) A (J3). If, in the usual way, we introduce coordinates on Z, taking (0, E, U) as our scale, and if A has the coordinate a, B the coordinate #, and T the coordinate a?', then x' = x + a. (1) We have shown that a transformation of this type between the points of I is a projectivity. Next we consider Construction II, keeping 0, E,U,A,P,Q and 0 E A B T U Fig. 45. hence R fixed, and confine our attention to the case A #= 0. We take QB as 6. Then J5, S> T describe ranges on l> OR, I respectively, such that (B) % (S) * (T), so that (B) A (37). Defining coordinates by the same scale, if the coordinates of A, B, T are a, x, xf respectively, then x' = ocxy (2) and this transformation is also a projectivity.
280 VI. PROJECTIVE space: synthetic definition Finally, in the construction [§7], p. 257, for the inverse A* of A, Fig. 46. we keep 0, E, U, Q> R fixed. Then A*, S, P, A describe ranges on I, OR, UQy I, so that (A*) % (S) * (P) * (A), and (A*) A (A). If the coordinates of A, A* are x, x' respectively, then xx' = 1, (3) and this transformation is a projectivity. These three projectivities are of fundamental importance. If we consider the transform of U under the projectivities, given respectively by (1), (2), (3), we see that under (1) U is transformed into itself; under (2) U is transformed into itself when a #=0, and under (3) U is transformed into 0. There is no element of K corresponding to U> but it is convenient to attach the symbol oo to U. The properties of this symbol are (i) oo + a = oo, (ii) aoo = oo (a 4=0), (iii) l/oo = 0, 1/0 = oo. These are the usual properties of ' infinity' assumed in elementary algebra. We consider now, with reference to the coordinate system set up on I, the bilinear transformation x' ~ ^^ (4)
10. CONSEQUENCES OP ASSUMING PAPPUS' THEOREM 281 where a, 6, c, d are in K, and are not all zero. If ad = be this transformation reduces to x' = constant, where the constant is in K, or is oo. Then x' does not depend on x> and we exclude this case. We now consider two cases. (i) c = 0. Neither a nor d can be zero, and , a b X — ~z X + -3« d d Writing xx = axjd, this becomes x' = xt + bjdy xt = oaj/dL We can therefore pass from x to x' by a transformation of type (2), which takes x into xv followed by a transformation of type (1), which takes xx into x'. Since both transformations define pro- jectivities, it is clear that (4) defines a projectivity when c = 0, ad + 0. (ii) c=#0, ad 4= be. We can pass from x to x' by the series of transformations _ _ ,7 __ * __ fa —ad) , __ a X\ == eXy Xc% — X-t ~t~ Cv, Xo z=z y X^ — ^a* •*> == ^4 n— • Xo c e These are of types (2), (1), (3), (2), (1) respectively. Since each transformation represents a projectivity it follows that (4) represents a projectivity. Hence we have Theorem II. A transformation given in a coordinate system on a line I by the equation ax + b x = cx + d9 where a, 6, c, d are in K and ad #= 6c, is a projectivity. We now show that every projectivity on I can indeed be represented by an equation of the form given in the above theorem. We first see what happens to the point x = oo under the five projecti- vities into which x' = (ax + b)/(cx + d) has been resolved. The first two leave it invariant, the third transforms it into 0, the fourth leaves 0 invariant, and the fifth transforms it into aje. The point oo is therefore transformed into the point x' = a/c. Hence we derive a further property of the symbol oo: ax + b a . = = -, when x =; oo. cx + d c
282 VI. projective space: synthetic definition Now let us suppose that a given projectivity transforms the points 0, 1, oo into the points p, q, r respectively. The transformation *> = «*-*)*+&-*) (5) (V-p)*+(r-q) clearly transforms 0 into p, 1 into q> and, by virtue of the relation above, it also transforms oo into r. This transformation is also, by Theorem I, a projectivity, since ad-bc = (q-p)(r-q)(r-p), which is not zero if p, q, r are all distinct. But a projectivity is uniquely defined if we are given three pairs of corresponding points. The transformation (5) therefore represents the given projectivity, and we have proved Theorem III. Any projectivity on I is given by an equation , ax + b cx + d where a, 6, c, d are in K, and ad 4= be. As an immediate corollary to this theorem we see that the projectivity in which (xvx[)9 (x2> x'2)y (x3> x'3) are corresponding pairs is given by \X — Xi) (#2 -" 3¾) __ (# -"*^l) (^2 ~" ^3/ {x-x^){x2-xx) ~ (a'-a^Mai-ai)' Let us now consider a change of coordinate system on I. Let x be the coordinate of a variable point in the original coordinate system for which (0, E, U) is a scale, so that these points are given by x = 0, l,oo respectively. Let (0'y E'y U') be a new scale on I. This does not of itself fix a new coordinate system, since we still have to set up an isomorphism between the points on I other than TJ' and the elements of the ground field K so that 0', E' correspond to 0, 1 respectively. There may be several ways of doing this. For example, if K is the field of complex numbers and we replace every element of K by the conjugate complex element we satisfy all the conditions and obtain a new coordinate system. But we can determine a particular isomorphism as follows. There exists a unique projectivity operating on the points of I which takes 0, E, U into 0', E'y V respectively. If this takes P into P' we associate with P' the element x which is the coordinate of P in the original coordinate
10. CONSEQUENCES OF ASSUMING PAPPUS* THEOREM 283 system. This correspondence between the points of I and the elements of K is clearly an isomorphism, and the coordinates 0, 1 are associated with 0'9 W respectively. We now find the coordinate y of a point X in the new coordinate system in terms of its coordinate x in the original system. If Y is the unique point of I such that (09E9U9Y)7{(0'9E'9U'9X)9 then y is the coordinate of Y in the original coordinate system. Since the correspondence between X and Y is a projectivity, ax + b y ** cx + d' by Theorem III, where a, 6, c, d are in Ky and ad #= be. Next, consider a projectivity between two distinct lines I and V, such that (O,#,tf,P,...)fi(0',#',t7',P',...). If a coordinate system exists on I for which (0, E9 U) is a scale, then the projectivity between I and V sets up a coordinate system for which (0', W 9 Uf) is a scale, corresponding points of the two ranges having equal coordinates. The projectivity is given by x' = x in the two coordinate systems chosen. We have stressed the point that the selection of a scale (0, E9 U) on a line does not, of itself, fix the coordinate system. But, having determined a coordinate system on any line I with respect to the scale (09 E9 U)9 we can determine a unique coordinate system on any line V with respect to any given scale (0\ E'9 U') on that line, by using the unique projectivity which makes (0, E9 U) correspond to (0\ E\ V) respectively, as above. The coordinate systems on I and V related in this way may be described as related coordinate systems. Let us fix, once and for all, a coordinate system on a definite fine I of the given space. For brevity we shall call any coordinate system related to this chosen one on I a related system. If AQ) ...9Ani E is a reference system, and we use the related coordinate system on A^ for which the scale is (Ai9 Eij9 Aj)9 we can introduce coordinates into the space as in §8. We call this coordinate system a scale coordinate system of the space.
284 VI. PROJECTIVE space: synthetic definition Now let m be any line of the space. It cannot meet all the $n_2 spaces A^ Ain%y spanned by vertices of the simplex of reference, and we may suppose that it does not meet that determined by A2,..., An. Then if it meets x0 = 0 in P, xx = 0 in Q> the points P, Q are distinct. By suitably choosing the multipliers, we may take P =(0,1, ^)2,...,pn), and Q = (l,0,ga,...,gn). Let R = (1, l,p2 + q2, ...,2>n + ?n). This point is on PQ and in the Sn_1 determined by A2,..., An> E. Now let „ , x X = {x0,...,zn) be any point of m other than P. Then xQ =f= 0, and we can write X^Q + Px^^Q + PZ. We now have a one-to-one correspondence between the points X of PQ other than P and the elements £ of K. We show that this correspondence is that determined by the related coordinate system on m with scale (Q, M, P). Project the range on m from the Sn_2 formed by A2,..., An on to the line AQAV The points Q, P, P project, respectively, into AQ9 EQli Av and X projects into the point (1, £, 0,..., 0), which has £ as its scale coordinate. This proves the result. Now let Ly My N be any three distinct points on m, and choose the multipliers of their coordinates so that M = L + N. Then, if Q = La + Ny and P = L/3+NS, then <x8-fiy*0y since P, Q are distinct. Also, the point X onm can be represented by the symbol X = L + NV, where rj a+M' Since this last relation is of type (4), it follows that the correspondence between the points Xofm and the elements rj of K (and oo) is that defined by the related coordinate system on mfor which the scale is (L, M, N).
10. CONSEQUENCES OF ASSUMING PAPPUS* THEOREM 285 Now carry out an allowable transformation of coordinates x' = Ax {A = (a#)). If, in the original coordinate system, L = (l0y...,ln)y N = (n0,...,nn), and Z = (x0,...9zn), where xi — li + ntf (i = 0,...,n), the new coordinates of Ly N9 X are L = (l'0> ...,l'n), N = (n'Q,...,n'n), X == \XQ> •••yXn), n where l\ = J^aylp n and therefore Xi — Yt&ijXj, ^i = Zi + ^i^ (i = 0,...,n). Hence, the correspondence between the points X of m and the elements 7/ of if (and oo) is unaltered by an allowable transformation of coordinates. Since this result is true for every line of the space, it follows that an allowable transformation of the coordinate system merely changes the simplex of reference used to construct the scale coordinate system. Finally, since a scale coordinate system is uniquely defined when we are given the reference system AQ,...,An,E, and an allowable transformation is uniquely defined (save for a common multiplier) when the points with new coordinates (1,0,...,0), (0,1,...,0), ..., (0,0,...,1), (1,1,...,1) are given, it follows that the set of transformations of coordinate systems from one related system to another coincides with the set of allowable transformations of coordinates. *
* CHAPTER VII GRASSMANN COORDINATES We now return to the algebraic methods used in Chapter V. From now on we shall assume that the ground field K is commutative, and without characteristic. Indeed, the main reason for considering more general fields in Chapter V was that this cleared the way for the detailed analysis of the Propositions of Incidence which was made in Chapter VI. It is not necessary to assume in this chapter that K is algebraically closed, since all the algebraic operations which we shall use are rational. 1. Linear spaces. In V, § 4, we defined a linear space of k dimensions in P'niK), and in the subsequent sections we obtained certain elementary properties of linear spaces. We indicated how similar definitions and properties could be obtained for Pln(K)9 and it was clear that the two sets of definitions and properties coincide when K is commutative. In this chapter it will not be necessary to indicate the ground field explicitly. We shall denote the space with which we are concerned by [n]. Linear spaces in it will be denoted by Sa,Sb>..., the suffix denoting the dimension. We recall that if 4< = («$,...,a£) (* = 0,...,fc) in a given allowable coordinate system, the points A0,..., Ak are linearly dependent if and only if the matrix is of rank less than k + 1. By definition, any point (a0,...,an) is linearly dependent on A0,..., Ak if and only if there exist elements ^o> • • • > ^k °f K such *hat k <*>i = S^al (i = 0,...,7i). We saw also that in a linear space of k dimensions there exist sets of k + 1 independent points, but no set of h independent points if h>k+l.
1. LINEAR SPACES 287 Now let 4°,..., Ak be a basis for Sk. The points 5°,..., Bk of Sk can be expressed in the form # = EALi' (i = 0,...,¾). As we saw, the set of points JS°,..., Bk forms a basis for Sk if and only if the matrix A , w. J A = (Aj) is non-singular. If 2^(^,...,^), and £ = (/?}), the space $fc spanned by ^4 °,...,-4 fc is determined, in an obvious sense, by the matrix Ay or, when A is non-singular, by the matrix B = AA. Thus Sk is determined by a set of equivalent (¾ + 1) x (n 4- 1) matrices Ay By..., equivalent matrices being such that we can pass from any one to any other by premultiplication by a non-singular square matrix. Let us consider an allowable transformation of the coordinate system n #* = 2^¾ (t = 0,..., n), i-o or x* = Px, in matrix notation. If, in the new coordinate system, ^• = (a*%...,a**), n then af = 2 riA.a|., and hence in the new coordinate system Sk is determined by the set of matrices equivalent to A* = {af*) = AP'. (1) Thus in [n] a linear space 8k is determined by a set of (k 4-1) x (n -f 1) matrices which are equivalent (in the sense described) in a given allowable coordinate system. The law of transformation for allowable transformations of coordinates is given by (1). However, for many purposes the identification of a space by means of a set of equivalent matrices is not the most convenient method, and
288 VH. GRASSMANN COORDINATES it is our purpose in this chapter to consider a more suitable system of coordinates for the ^-spaces of [n]. These coordinates were first introduced by Grassmann, and they coincide, when k = 0, with the coordinates already defined. 2. Grassmann coordinates. We choose an allowable coordinate system and keep it fixed. Then a given Sk of [n] is determined by a set of equivalent (k+ 1) x (n-f 1) matrices A9B>... which are all, of course, of rank k+l. Suppose that B = AA, where A is a non-singular (k+ 1) x (k+1) matrix. We consider the (k+ l)x (fc-f 1) submatrix of A formed by selecting the columns with suffixes i0>iv ...,^, and the corresponding submatrix of B. We have and taking determinants of both sides, we find that fil fit • n> ■ fit -Ml a? "■to where | A \ = detA is different from zero, and does not depend on the suffixes i0,...,ik chosen. We write this equation in the form &*o...^ = MK0...*r Since A is of rank k + 1, for at least one set of suffixes i0,..., ik, «<•...<* *0, and then J<o».ik 4=0. Again, from the elementary properties of determinants, ^0...^ = 0 if the integers i0, ...,¾ are not all distinct; and if j0,...,jk is any derangement of the set i0>..., ik> a io.>.ik ea. U-.iv where e = + 1 or — 1 according as the derangement is even or odd.
2. ORASSMANN COORDINATES 289 From these considerations it is clear that the set of equivalent matrices A,B9...y which determines Sk also determines a set of equivalent / *+ x j-tuples (...,¾...¾....), (..,^0...^, ...)• We define the Orassmann coordinates of Sk to be any one of the equivalent I, I-tuples. Hitherto, in using the term coordinates in connection with points we have implied (i) that each point has a set of coordinates which is unique except for a common multiplier, and (ii) that any (n + l)-tuple gives the coordinates of some point. This property (ii) is not characteristic of Grassmann coordinates, and we shall see later that except when & = 0or& = w — Ian 1-tuple (...,¾...^ ...) can only give the coordinates of an Sk when the ai0...ik sa*isfy certain algebraic relations. The property (i) of coordinates described above is, however, an essential one, and, before proceeding, it is of fundamental importance to show that no two distinct spaces Sk, S'k can have the same coordinates. We do this by showing how to write down the equations of n — k linearly independent primes through a given Sk> and remark that these equations depend only on the coordinates of Sk. Suppose that Sk is determined by the basis 4* = (aJ,...,o£) (* = 0,...,^), and let the Grassmann coordinates of Sk be (...,^,...^, •••)• One at least of these coordinates is not zero. Suppose that The condition that any point (x0,...,xn) should lie in Sk is that the matrix '«? <4 <xo • . • • < CCn xn.t should be of rank k+ 1. In this case the determinant of the sub- matrix formed from the columns with suffixes i0, ...,ifc+1 must be zero. Expanding this determinant in terms of the last row, we get the equations ft+i RPI 19
290 VH. GRASSMANN COORDINATES These equations, for all possible choices of i0,...,H+i> are either identically true, or define primes which contain all points of Sk. Jlence the aggregate of these L I equations defines a linear space which contains Sk. In order to show that these equations actually define Sk itself it is therefore sufficient to show that amongst them there is a set of n — k linearly independent equations. We consider the equation of the set for which where j is any integer 0 ^ j ^ n distinct from j0,..., jk. This equation is /1 = 0 Since each Grassmann coordinate is skew-symmetric in its suffixes we may write the equation in the form 2(-1^,(-1^...^^..1,+(-1^%...¾ = o, k /1=0 If 0'o> -~>jk>jk+i>-~>jn) is a derangement of (0,...,n), we consider the equations (2) for the values of j given by The matrix of coefficients of these equations is an (n — k) x (n+1) matrix. It is of rank n — k> for the determinant of the submatrix corresponding to the columns with suffixes jk+l,..., jn is equal to (Pj ...ta)n~"&> anc* *k*s *s n0* zero- I* follows that the set of equations (2) defines n — k linearly independent primes, and therefore defines Sk. The equations of Sk are therefore determined by the Grassmann coordinates of Sk, and it follows that no two distinct ^-spaces can have the same coordinates. We conclude this section by considering the effect of an allowable transformation of coordinates in [n] on the Grassmann coordinates of Sk. We saw in § 1 that the transformation x* = Px (3) transforms the matrix A which defines 8k into the matrix A* = AP'.
2. GRASSMANN COORDINATES 291 Hence Pt...ik oc *0 a*o a; *fc a ft h = s 11*0 fy> Ti0hk Tikho 'ikhic < < rtk by II, § 8, Th. V, and this is equal to S r<o** rikho rikhk Ph0...hk- (4) Hence an allowable transformation of coordinates in [n] induces a linear transformation of the Grassmann coordinates of a &-space. The matrix of this transformation has as its elements the minors of order k +1 which can be extracted from P. Let all minors which come from the same set of Jc +1 rows (or columns) of P be placed in the same row (or column), and let the priority of elements in rows or columns be decided on the principle by which words are ordered in a dictionary; that is, let the order be lexicographical order. The resulting I. I x J. J matrix is called the (Jc + l)th compound of P, and it is denoted by pcfc+i). From the reasoning given above it follows easily that (PQ)(fc+l) = pifc+«QC*+l># (5) IfP = A+i (*ke un^ matrix), evidently 1**+» = Ln+1) Hence, taking Q = P-1, we see from (5) that i(n+1) = (/n+1yfc+« = **"■»#*+», and therefore the matrix is the inverse of P^+U. (p-l)(fc+D 19-2
292 VH. GRASSMANN COORDINATES These compound matrices are of importance later on, when we consider them as defining a particular type of allowable transformation in an associated JV-space, where ^+1 = 1, I. 3. Dual Grassmann coordinates. An Sk may be defined not only by Jc +1 linearly independent points which lie in it, but also by means of n — Jc linearly independent primes which pass through it. In a given allowable coordinate system let Sk have Grassmann coordinates and let 2^ = 0 (i = 1,...,n-&) (1) be the equations of n — Jc linearly independent primes which contain ** If U = («$), the equations (1) may be written in matrix form as Ux - 0. (1)' These equations are the equations of Sk, and the (n — Jc) x (n+1) matrix U is of rank n — Jc, since the primes are linearly independent. Any prime through Sk has for its equation a linear combination of the equations (1), and n — k such primes, given by the matrix equation yx = ^ are linearly independent primes through Sk if and only if there exists a non-singular (n — Jc) x (n — Jc) matrix M such that that is, in the terminology of § 2, if and only if U and V are equivalent matrices. Thus Sk can also be determined by a set of equivalent (n — Jc) x (n +1) matrices. If, as in § 2, we consider the transformation of coordinates x* = Px, (2) the equations of Sk become UP-ix* = 0, and the set of equivalent matrices U> V, ... becomes the set of equivalent matrices UP"1, FP"1, ....
3. DUAL GRASSMANN COORDINATES 293 Again, the determinants of the (n — k)x(n — k) submatrices of the set of equivalent matrices U> V, ... determine a class of equivalent I , J-tuples. A representative member of the class can be determined from the matrix U. We write P {ii»»in-k < U in-h U n—k U n—k in-k (3) Then Sk is determined by the set of I ,) -tuples (...,^1-^,...). We call these dual Orassmann coordinates. It is easy to show that no two distinct spaces Sk, Sk give rise to the same set of dual coordinates. If (x0,...,#n) is any point of Sk, then n—k n i-li~0 where Ax,..., An_& are any elements of K. Now choose A< = (-1)' < <-' <+1 1 ul~« ' <-*-! ui"1 vn-k and we obtain the equation J^ph-ln-k-ilzj^ 0. j-0 (4) If we give jx, ...,jn-k-ia^ possible values, (4) gives us a set of equations defining a linear space which contains 8k. As in the previous section we now show that there are n — k linearly independent equations in the set, and deduce that the linear space coincides with Sk.
294 Vn. ORASSMANN COORDINATES Since U is of rank n — k, there is at least one non-zero dual coordinate. Suppose that -pOi...«n-*4=o, and consider then —A; equations obtained from (4)whenj1, ...,in-*-i are selected from the set ax,..., an_k. If ai is the element omitted in the selection, (4) becomes n 2 pai — ai-\ai+i — an-k1x4 = 0, that is, ^ai—^-1^+1-an-fcai# 4- 2 pa\ — at-i<H+x — an-k1xA = 0, i*«i On-* Or pai"a*-*Zai+ S ^1-^-1^+1-^-^ = 0. i+cH a»_* If we take i = 1,..., n — k, we obtain n — k linearly independent equations, since the determinant of the submatrix of the (n — k) x (n + 1) matrix of coefficients which corresponds to the columns av ..., an_k is evidently and this is not zero. The dual coordinates of ^¾ therefore determine it uniquely. These dual coordinates are closely related to the ordinary Grass- mann coordinates defined in § 2. We now prove Theorem I. Let (i0,..., in) be any derangement of (0,..., n). Then if Sk has Orassmanncoordinates (..., #t-0...i*> •••)> anddval Grassmann coordinates (...,p<*+i,,,<»,...), there exists a non-zero element p in K such that This equation defines q^mmmir The proof is in two stages. We show, first of all, that the transformation (2) of coordinates in [n] induces a non-singular linear transformation in the coordinates &0...^, of the same kind as that induced in the coordinates pi9_ik [cf. §2]. We then show that P can be chosen so that in the new coordinate system it is evident that for some p. The theorem then follows from the fact that the co-
3. DUAL GRASSMANN OOORDINATES 205 ordinates (...,!>*,...<*>•••) and (...,#*...**>•••) ar© transformed in the same way for each allowable transformation in [ri\. Since the transformation (2) transforms the equations of Sk into if then UP-W = 0, P -(t,,), and P-i-(crtf)f 2> *ii —in-* 3 . 2^: n-fc, rtfn-» == s /ll ...,irt-Jfc W u: 'Jl W: 7*-* 7/n-fc = £ ph-U-k °Viil • ^1 in-* • • • • • • °Vn-*il * °"in-*in-* Hence = e.-... = ¾. .in S P**+t-f- /* + i,...,Jn ,.<n 2^0...^/1^0-..^ °V*+i<*+i 1 ^Vnfc+i °V* + X<*+1 °Vni*+i • • • • • °V*+i*» ' ^inin Vfr+iin Vjnin 1 = Z#:::fcffj....*i °V*+ii*+i ' °V*n<» °Vni*+i * °Vn<ii the summation being over all the 1, 1 distinct partitions (Jo> ~->Jk)Uk+v->Jn) of Oot—Jn). The sets (i0>...>%J> Oo»—»j») are, of course, derangements of (0,...,?&).
296 VH. GRASSMANN OOOEDINATBS But [II, § 8, Th. VIII], ^t+l^+l ^inU+l • • Jk+xin ^inin -^:::^1^1-1 TUh Tiote ' Tikh Tikik where I PI = det P. Hence Ttoh TUii Tiih Tik1k ¢/0.../4- (6) Comparing this with the result of § 2, (4) we see that the allowable transformation of coordinates in [n] given by (2) induces the same linear transformation on the coordinates (...,¾...^ ...) as it does on the coordinates (...,^...^,...), save for a common non-zero factor. Now, by V, § 4, Th. V, there is an allowable coordinate system in [n] in which the points A0,..., Ak, which form a basis for Sk> coincide with vertices X*°,..., X*k of the simplex of reference of an allowable coordinate system. Let us suppose that the transformation (2) is to this coordinate system. Then Sk is now given by the (k + 1) x (n +1) matrix 1 0 . . . 0> 0 1...0 <0 0 0> and therefore p$m k = 1, but p*mmmik = 0, unless the set (i0, ...,ik) is a derangement of (0, ...,&). Also, Sk is given by the equations xi = 0 (i = k + l,...,7i), that is, by the (n — k) x (n + 1) matrix . 1 0 . . 0 1 . . 0 . . and therefore #*...& = p*k+1-n = 1, but qf0_ik = ±p*i*+vi* = 0 if (*o» •••»**) *s no* a derangement of (0, ...,&). In this special system of coordinates * ^ &o...** — Pio...ik>
3. DUAL GRASSMANN COORDINATES 297 and therefore, as explained above, we conclude that in any allowable coordinate system the relations (5) hold for a suitable p. The Principle of Duality [V, §5] throws further light on the relation between the Grassmann coordinates of an Sk and its dual Grassmann coordinates. If the points Ai^((49...9ain) (i = 0,...,A?) form a basis for Sk, the equations n ^ajxi = 0 (i = 0,...,fc) define the dual space SnmmkmmV Hence, if (-~>Pi0...ik>---) and (...,^1-*»-*,...) are the Grassmann and dual Grassmann coordinates respectively of 8k> and (--1¾...¾.^...) and (...,s^-*'**+i, —) are the Grassmann and dual Grassmann coordinates of 2^.^, then evidently . , J ql*'~lk = PPi0 ...ik> and therefore, by our theorem, Using these results it is possible to read off a property of Grassmann coordinates dual to any known property. For instance, it follows at once from the fact that the equations of Sk are [§ 3, (4)] n 2 p*v1n-k-i1 Xj = 0, i-0 that if the vector is not zero, it gives us the coordinates of a point in the Sn_k_x whose Grassmann coordinates are (...,¾...^.^ ...)• By giving the set of integers (jv ...,jn_&_i) a^ possible values we obtain a set oi points in #n_fc_i which contains a basis for this space. 4. Elementary properties of Grassmann coordinates. The condition - Pa0...ak = °> for a given set of suffixes (a0,...,a&), is easily interpreted as a
298 VH. GRASSMANN COORDINATES geometric property of the 8k which has the Grassmann coordinates (-.^...^,...). For if Ai = (al...,ocin) (» = 0 k) is a basis for Sk, the condition is < • < I - o. • • • From this statement of the condition we deduce the existence of elements \$,..., Afc of K, not all zero, such that k £A^ = 0 (j = a0,...,a&). i-0 k It follows that S \Ai is a point in 8k whose coordinates satisfy i-0 the conditions _ _ _ _ These last equations define an Sn_k_ly a basis for which consists of n — k vertices of the simplex of references. The condition we are examining is a sufficient condition for Sk to meet this Sn_k_v Conversely, if Sk meets this Sn_k_ly one of the points, say Ak> of a basis for Sk can be chosen in Sn_k_v and then < = ... = <x£k = o, and therefore pao„,ak = 0. This result can be generalised: Theorem I. Let a0, ...,a8 be s+1 distinct integers (s+l^k+1) chosen from the set 0,...,n; and let #n_8_i be the linear space whose equations are xao == • • • == xas = 0. The necessary and sufficient conditions that Sk should meet Sn^8^ in a space of dimension at least k — s are Pao...asis+i...ik == ® for all possible choices of i8+v ..., ik. Clearly we need only consider sets (i8+v ••.,*&) which consist of distinct numbers, none of which is equal to any member of the set (a0,...,as). After what has been proved above we may also suppose that 8 < k.
4. PROPERTIES OF GRASSMANN COORDINATES 299 Now let us assume that Sk and Snr_8_1 have a linear space of at least k — 8 dimensions in common. Then we can find k — 8+1 linearly independent points of Sk in Sn_8„v and we may include these in a basis for Sk, say as the points A8, A8+1> ..., Ak. Then 0 (i = sy...yk)j = a0, ...,¾). It follows that in the matrix A = (aj) which defines Sk the columns corresponding to the suffixes a0,...,a5 have non-zero elements only in the first s rows. The rank of the submatrix formed by these columns is therefore not greater than s> and hence these columns, considered as vectors, are linearly dependent. Therefore the columns corresponding to suffixes (a0,...,ad, i8+1> ...,½) are also linearly dependent, and finally This proves the necessity of the condition. To prove its sufficiency we recall that any Sk intersects an Sn_8_1 in a linear space of at least k — 8—l dimensions. Hence we can choose the points A8+1, ...,Ak to lie in Sn_8_l9 so that a) = 0 (i = 8 +1,..., k;j = a0, ...,aj. Then evidently Pao...as{8+i...tk a «0 a a* a a0 a: as a: a: k a: I* a: t* using elementary properties of determinants. Not all the determinants a8-*1 a* a 'it+t < can be zero. If they were, the last k — s rows of A would be linearly dependent. Since k — s>0, this would imply that the rank of A was less than k+ 1, contrary to hypothesis. Hence the equations imply that Pao...asis+i...ik """ a" a! '«« = 0. o£ a ^
300 VH. GRASSMANN COORDINATES It follows that there exist elements A0,..., A& of K> which are not 8 all zero, such that the point 2 Ai-4i, which is a point of Sk linearly independent of A8+1y...yAk9 lies in Snr^9mmV Hence 8k and Snr_a_1 have at least k — s + 1 linearly independent points in common, and therefore intersect in a space of at least k — s dimensions. This completes the proof of the theorem. We now show how to determine a basis for an Sk when we are given its Grassmann coordinates (...,2>t-0...**> •••)• Let ^..,^ be a non-zero coordinate, and define ri __ >n). (1) x) =2>a0...ai-i^+1...a* (» = °> —>*J 3 = °> Since < =Pa....a**0, the (Ti+1 )-tuple Bi = (aj,..., x*n) represents a point, which we denote by Bl. Since «} = 0 (j = a0, ...,flM,aw, ...,^), the point Bl lies in the #„__£ whose equations are We show that J3* is the unique point of intersection of Sk with this Sn_k. Since (2) **; = a a! a a A; a ai-i ay a A; a a* A: = 2(- l)h+i cc cc cc «0 ft-1 a. a «0 a, o 'tzi-i a a a '«* a a; a»-i a a! ft+i A; a a! a a* a* a* a: a* /vfc -SAfo*, fc=»0 it follows that J3* = 2 Aj[-4* is a point of /¾. If Sk meets the j8fn_fc ft = 0 given by (2) in more than the point Biy it has at least a line in common with it, and this line, lying in #n_fc, must meet the Sn^k^1
4. PROPERTIES OF GRASSMANN COORDINATES 301 This S^^i would then meet Sk, and by Theorem I we should have Pa0...ak = Vy contrary to hypothesis. It follows that Bl is the unique point common to Sk and the Sn_k given by (2). We now show that the points B°,..., Bk defined in this way form a basis for Sk. Since each J3* lies in Sk it is sufficient to prove that the k + 1 points are linearly independent; that is, that the matrix 2L = /Xq • xn xn is of rank k +1. The submatrix of X whose columns correspond to the suffixes a0,..., ak is fPa0...ak 0 0 Pa0...ak 0 0 0 Pa0... a>ki (since #at-= jPa0...a*)> an(* **ris submatrix is non-singular, since Pa0...a**0- Hence X is of rank fc+1, and the points B°,...9Bk form a basis for Sk. An immediate consequence of this result is the fact that there exists a non-zero element p in K such that PPi0...ik = x\ lo X) %l Ik for all choices of (i0, ...,¾). If we take ij = aj U = °> this equation becomes PPao...ak=(P a0...ak ,k), and therefore P = (Pa*...ak)k- Again, suppose that ij = aj9 except when j = t, u> v>.. being 8 elements in this last set. Then PPi0...i* x) a. *4 x\ lu *l Xak }wy there < x\ < < "at
302 VII. GRASSMANN COORDINATES In any column corresponding to the suffix a{ every element other than xj^ is zero. Since a determinant can be expanded in terms of the elements of such a column, and since x^ = pao_ak is on the principal diagonal, PPu-i* = (2V..a*)fc~8+1 that is, a& vlw a£ (a Oo».a* )^^0...^ = XI *z (3) If we dualise these results we can obtain a basis for the primes which pass through Sk. But we may also use the results of § 2. If (a0,..., an) is an even derangement of (0,..., n)y by hypothesis. By § 2, (2) the equations k Pa0...akXi~~ 2 Pa0...ah-1iah+1...akXah = 0 (l = ak+l> •••» an) h = 0 are the Equations of a dual basis for Sk. Since for the given values of i we may write » = 0j 0' = &+l,...,n), and then the equations become k pak+l...OnXam+ 2 pak+i-aj-iah*j+i-<*nxah = 0 (j = fc+l,...,w), j fc-0 or, more simply, n 2 p«*+i-^-1^+1-^¾ = 0 (j = fc+1,...,n), <-0 on inserting terms with zero coefficients. We write ulj = pak+i'"aj-iiaj^van9 and the equations of Sk are then (4) S «***« = 0 (j = k+1, ...,n). <-o
4. PROPERTIES OF GRASSMANN COORDINATES 303 The matrix K+1 • «fc+i\ defines Sk, and pph+i'^in = <4*-tt • 4+i <*+i . ttJr • Since uf = jp«*+i-«» (j = k +1,...,w), ppak+i»*On = ^a/,+1...an 0 . . 0 p^+i-an 0 0 • • pak+1...an and therefore we may write our previous result in the form (pak+i — <*>n)n-k-l pik+1 •••*»! = <4*tt Mfc-i <*+l • «*• If i^ = ¢^, except when j = t,u,v,...,w, we find, as before, there being s elements in the set (t, u, v,..., w), that (pak+l — Ony-ipik+i — in = U}' U W U, iu, (5) Finally, we note the following relations between the uj defined in (4) and the x\ defined in (1) above, (i) If j > k, and i ^ k, (ii) if i>j >kyi 4=j, and Z, m ^ k, I #= m, «? = o = <*; (iii) ifi>fc,j<&, -rao...Gy-i<2ia/+i...<2n #i- We now use these results to obtain necessary and sufficient conditions for a given Sh to meet a given Sk in a space of at least t dimensions.
304 VII. GRASSMANN COORDINATES 5. Some results on intersections and joins. Let the coordinates of Sh be (.-.,jP<0...to> •••) and let the dual coordinates of Sk be (..., g^+i — S ...). We prove Theorem I. Necessary and sufficient conditions that Sh and Sk have a linear space of dimension not less than t in common are 2 ^•••""-'^••^...A.yv..^-, = © (1) for all choices of <xk+1,...,ocn_3; /?0,...,/3h__s, where s = h — t+l. A sufficient set of conditions is obtained from (1) if we restrict our choices of a&+1,..., an_8 to all possible selections from ak+v ..., an, and Ao> --->Ph-8 t° aM possible selections from 60, ...,6^, where afc+1,...,an and b0, ...>bh are sets of n — k and h+1 numbers respectively chosen from the set 0,...,n such that ^+1--^...6, + 0. (2) Clearly if ^--^0...^ = ° for all sets (an_a+1,..., a J, (&A_a+1,... A), condition (1) is satisfied. We need only consider sets ak+v ...yan\ 60, ...,6^, therefore, for which (2) is true. If we now write and xi = Pb,...bi-libi+l...bh> the equations of Sk can be written n S 1^ = 0 (j = *+l,...,n), and the points ^ = (4,...,4) (*' = 0,...,¾) form a basis for Sh. Any point of Sh can be written in the form h 2 Aaa^, and lies in 8k if and only if o S 2*4*?Aa = 0 (j = ^+l,...,?i). (3) These equations in (A0,..., A^) determine the (h + l)-tuples such that h 2Aaa^ lies in Sk. Hence a necessary and sufficient condition that o
6. SOME RESULTS ON INTERSECTIONS AND JOINS 305 Sh and 8k meet in an St (at least) is that the equations (3) should have at least t + 1 independent solutions. The rank of the matrix {zM must therefore be h — t, at most [II, § 6]. For this to be the case it is necessary and sufficient that every determinant of 8 = h — t + 1 rows and columns extracted from this matrix should be zero. A set of necessary and sufficient conditions is therefore given by the equations 2 uXiA • S <*£ A-0 A-0 2 mW1 A-0 2 <4' A«0 = o, (4) for all possible selections of iv...,i8 from (&+1,...,n) and of ji> '-->j8 from (0, ...,¾). The above determinant, by II, §8, Th. V, Ai,...,A* <* «2? <* tt: *« a& a* Ai ^A, Now, by equation (5) of the preceding section, u'l u u *1 u = e(qafc+i••• ^)8^1 ga*+i ••• an-« *i — **f where afc+1,...,ocn_8 is a selection of n — k — s numbers of the set a&+1,..., an which is uniquely determined by the set ^,..., i8, no two distinct sets iv ...,i8 giving the same selection ak+1, ..., a„_s, and e is +1 or —1 and is uniquely determined by iv...,i8. Similarly, using equation (3) of the last section, *% a* Ai - e'{Pbo...bh)* 1P*i...XMfio-fih-i> where /?0> •••>/?&-* is a selection from the set b0)...,bh which is uniquely determined by jl9..., jg, and e' = ± 1 is also uniquely determined by jl9...,j8. We can therefore write, instead of (4), Alt..., Aj where the factor ee', being independent of \lf..., As, may be ignored. HPI 20
306 VH. GRASSMANN COORDINATES Using (2) it follows that the necessary and sufficient conditions (5) are equivalent to equations (1) for all choices of a&+1,..., an_8 from a^+1,..., an, and of /?0, ...,fih-a from 60,..., bh. This proves the theorem. We now obtain some formulae connected with the joins and intersections of linear spaces. Let Sh have the coordinates (•••>Pi0...ih> •••)> sk the coordinates (...,&0...iife,...), and, as usual, denote the dual coordinates of Sh by (...,p*A+i•••*»,...) and those of flffc by (...,^+1-^,...). Let a** = (4>---^1) (i =-= 0,...,¾) be a basis for aS^, and y* = (yj,••.,!*) (^ = 0,...,¾) be a basis for #&. A necessary and sufficient condition that Sh and Sk meet is that there exist elements Aq,...,\h and /i0, ...,/£&, not all zero, such that h k 0 0 A necessary and sufficient condit: \y\ on, therefore, is that the matrix xn\ W y" 2/n/ should be of rank h + k +1 at most, since the rows are to be linearly dependent. Hence all determinants of h + k + 2 rows and columns extracted from this (h + k + 2) x {n+ 1) matrix must vanish. If we consider the determinant which corresponds to the columns with suffix i0,---,ih+k+v an<i expand it by Laplace's formula [II, §8, Th. II], we have n* ...»ja+*+i It can be verified that these conditions are equivalent to those given by Theorem I. If 8h and Sk do not meet, the points x°>..., xh, y°,...,yk are linearly independent [V, § 4], and determine the join of the two spaces.
5. SOME RESULTS ON INTERSECTIONS AND JOINS 307 Hence the determinants of h + k + 2 rows and columns extracted from the matrix (6) are the Grassmann coordinates of the join of the two spaces. We may write these coordinates as (..., rio ### ih+k+l,... )> where U* -»>lh+k+i Ignoring a common factor, we may also write these equations as r<0...ih+k+i = ? ±Pio...ihQih+i...ih+k+i> (8) where the summation is over all derangements Jo>--->ifc+*+i °^ i0, ...,*A+fc+i> the positive sign being taken with even, and the negative sign with odd derangements. Introducing the dual coordinates, we have where (j0, ...,jh+k+1, ih+k+2,...,»») is an even derangement of (to,...,tN). Hence i = ( _ 1 )(^+1)(15+1) 2 jp. j qlo-hh+k+i-in. Ignoring the common factor + 1, we may write this as i An alternative form is This follows from equation (8), the proof being left to the reader. By duality, or by an exactly similar argument, we show that if h + k > n, and th — iw-h-k = ^8U'--j™-h-*ph"-1n-hqin-h+i-"hn-h-k9 then fa- im-h-k = 0 (all iv ..., i2n-h-k) is the condition that 8h and 8k meet in a space of more than h + k — n dimensions, but if these spaces meet in a space of dimension h + k — n, the coordinates of their intersection are (...,**!••• *»-*-*,...). Equivalent formulae can also be found for tio_ih+k_n. MOM
308 VH. GRASSMANN COORDINATES We now obtain the coordinates of the projection, from a vertex Sh, of a space 8k on to a space $n_a_i. We assume that Sh does not meet Sk. The projection is then the intersection of the join of Sh and Sk with Sn_h_v The coordinates of the join of Sh and 8k are given by and if the coordinates of Sn_h_1 are (..., s1*--\ ...), the coordinates (--->hQ...iv •••) of the intersection of this join with #„_&_! are given by the equations j i m In the case h = 0 the coordinates of the projection of Sk from the n point 6 = (60,...,bn) on to the prime 2 -¾¾ = 0 can be obtained more simply. Let ° zi = (xl...yxin) (1 = 0,...,¾) be a basis for Sk. Since b does not lie in Sk, the points 6, x°, ..., xk are linearly independent. Hence the points t/°,...,yfc, where y* = ^-(s^)& (^= 60^0+.. . + 6n^n*0) are linearly independent. But since i i \ 3 1 the points y°,...,yk lie in the prime 2 -6* xt = 0. Hence these points, i which lie in the projection of Sk on to the prime, may be taken as a basis for the projection. Therefore U .** y\. yt y\ yl *<-*%„ M-^K where A* = S Brf.
5. SOME RESULTS ON INTERSECTIONS AND JOINS 309 Expanding this determinant we obtain £ V..** = Ak+1 4 4 • 4 • A -Ak^b v Xi, 4 • 4-. • 4-. a° 4+. ** 4+. x*- 1 • A\ and substituting for A* we find that i / We have therefore proved Theorem III. The coordinates (..., tft- ...^,...) o/ ^ projection of the k-space (...,2^...^,...) from the point (60,..., bn) onto the prime B0x0 + ... + Bnxn = 0 are given by the equations where k*...ik = ^0...^-22(-1)^¾¼...^^...^ i i A = B0b0 + ... +Bnbn. 6. Quadratic /j-relations. We have shown that any &-space Sk has a set of coordinates (...,^ ...**>•••) in a given allowable coordinate system, but it does not follow that a set of I, J elements Pi0...ik which are skew-symmetric in their sufi&xes, and are not all zero, are necessarily the coordinates of an Sk. Indeed, this is not the case, except when & = 0 or Ic = n—l. We must therefore consider what algebraic relations, if any, connect the coordinates of an Sk. We first show that there can be no linear relation. Suppose that 2 u, io...ikPi0...ik is a homogeneous linear form in the I, J independent indeter- minates i* .¾ (which are skew-symmetric in their sufi&xes) such m S.^...i*2V..i* = 0, (1) U ik whenever P{ ik is replaced by the coordinates pio _ ik of any fc-space. We show that each coefficient uio„tik must be zero. Let us show that ua*...av f°r instance, is zero.
310 Vn. GRASSMANN COORDINATES Consider the &-space defined by the k+1 points ^ = (^,...,¾4) {i = 0,...,¾). We see at once that unless i0>..., i& is a derangement of a0, ...,&*, whilst #ao...ait = 1. The relation (1) becomes <V..a* = 0. Thus we have Theorem I. The coordinates of the k-spaces in [n] do not satisfy any linear relation of the form 2 . ^i0:..ikPi0...ik = 0. There are, however, relations of higher degree satisfied by the coordinates of every Sk in [n]. We assume that k 4= 0, and k 4= n — 1, and denote by jP(P) a homogeneous polynomial in the I, , I independent indeterminates Pi(i_ik (which are skew-symmetric in their suffixes), and denote by F(p) the result of replacing Pi0_ik by the coordinates Pi0.,.ik of an Sk. We seek non-zero forms F(P) such that F(p) = 0 for every Sk in [n]. Let iv...,ik be k distinct numbers which are chosen from the set 0,..., n, and let j0>--->jft+i be k + 2 distinct numbers chosen from the same set. We define fc+i ^1...^0...^+1^)== S (- !) Pii...ikhP1o...h-ih+i...ik+i> (2) and prove that Fix„Akth...iit+1(p) = 0 for every Sk. Let 4< = (a*,...,a£) (1 = 0,...,¾) be a basis for any given Sk. Then fc+i = 2(-l)x A-0 /y9 /v° /y° 4 • < 4* aJ. • ajA-! ajA+i * aJ*+i ^o * aJA-i aJA+i • aJ*+i
6. QUADRATICS-RELATIONS 311 If A* is the cofactor of afA in the determinant < a: t* <4 we may write k fc+1 ,,-OA-O ' «<, «4 <*h a: '*. 4. a: ix- oc. ■h+i a! ato-, %+t i*+i ay*+i = S A" /.-0 <4 a. *A 4 a', JA 4*+l Since every determinant in the above sum has two rows equal, every term in the sum vanishes, and therefore ^...^....^) = 0- (3) This relation simplifies when the numbers iv ...,ik include some «fi»»..jw If ir=jr (f-1,...,A), all other terms vanishing. Since the indeterminates are assumed to be skew-symmetric in the suffixes we have ■^ii...ik>Jo>~Jk+i("' = -^1...^0^1-^^+1^^1-^0^1-^^+1 = In this case there is no relation. We obtain non-trivial relations containing three terms if respectively. Writing ik = h, we find that and therefore Pii...ik-ihlPi1...ik-1mn+Pii...ik-ihmP{1...iit-1nl + Pi1...ik-1hnPi1...ii-1hn = 0.
312 VH. GRASSMANN COORDINATES We now obtain an alternative form for the relations (3). If we write . . ... h> •••> V-i> Vn> • • • y H> h^Joy •••>,?& in place of il9..., i8, is+1,..., ik,j0Ji>..., ju+i for any 0 < s ^ k, the relation (3) becomes ^o.»<i-i<i+i«.<b<.fc...fc^) A esa 1 or (-i)fc-iV.<A.../* using the skew-symmetric property of the coordinates with respect to the interchange of suffixes. Hence (3) may always be written in the form k The relations (3), or (4), will be referred to as the quadratic p-relations. We now prove Theorem II. If (...,P*0...<fc>...) be I, J elements of K which are not all zero, which are skew-symmetric in the suffixes and which satisfy the quadratic p-relations, then there is a k-space which has coordinates (...,2V..fc>—)• If there is such a &-space it is, of course, unique. Since not all the coordinates are zero we may suppose that IV.. «* flit will simplify our notation if we assume that ai==i (i = 0, ...,&). It will be evident that the method of proof can be extended to any set a0,...,ak. We have, then, 2>o...**0,
6. QUADRATIC ^-RELATIONS 313 and since we are dealing with homogeneous coordinates we may assume thatp0##Jfc = 1. If b] = 2>o...i-i><-n...* (i = <>,...,&; j = 0,..., n), we shall prove that the k +1 points #-(¾....¼) (* = 0,...,fc) determine a fc-space with the prescribed coordinates. Since b\ = p0tk = 1 (i = 0, ...,&), the submatrix formed by the first k+ 1 columns of the matrix (6^) is the unit matrix. Hence this matrix has rank k + 1, and the points B°f..., Bk determine a &-space, which we call Sk. If &o...** - 6?. bl bi ut .), and all that the Grassmann coordinates of 8k are (...,0*o...<*> « we must prove is that Since the Pi0...ik are skew-symmetric in the suffixes, this result is independent of the order of the suffixes. Let the set iQ,..., ik contain the integers lv l2,...,l8 which are greater than &, and write the set Since &H*f 0'<*)> the determinant which gives #^...¾ can be simplified by expanding it in terms of the elements of any column whose suffix is in the natural order of the integers. This column contains only one nonzero term, and this is on the principal diagonal, and is unity. Hence, noting which row and column are removed by this simplification from the determinant we are considering, we eventually find that Qi<,...ii = H' • 6fc (6) Our theorem is now proved by induction on a. *o> •••> *ft — ^> •••> ^» • and &<,...<* = 1 =^0..^-
314 VH. GRASSMANN COORDINATES If * = 1, ft0...<* = b1l\ = ft...fc-lMi+l...*» so that the theorem is also true when $=1. We therefore assume that if s < t, jp^...fc = &0...i*; that is, I * * * Now suppose that * = t. Since the pio.„ik satisfy the quadratic ^-relations we have, from (4), where r is at our disposal (0 ^ r ^ k). Choose r so that ir = Z,. Then Pi0...ir-1Hr+i...ik = °> unless j = jl9..., j/, these being the only missing integers of the set 0,...,¾. If now j = jA, say, ^...^.^^...¾ has only <-l suffixes wof chosen from (0,..., &), and hence, by the hypothesis of induction Pio...ir-ihir+i'»it ~~ b\\ k 6JL, 6k C «U 6¾ &L. 'fc-i 6¾ &?>-• 6¾.. ^- *'- HA-+.1 fy+1 *t * 6¾ 6£* 6¾ «-, to to 6¾ 6* *t • to on expanding by the last column, the only non-zero element in
6. QUADRATIC ^RELATIONS 315 which is 6¾ = 1. Now jfy...* = l> an(l, since j = jx and ir = lt, we have 2>o...j-itri+i...& = HA> an(i so ^e quadratic p-relation gives us Pi0...ik t A-l 1)H-A6£ Mi - &£-. &£+» . *£ *L • bfcl • 6fc? i 6*J 6¾ • 6JL 6fc a* H- n thus proving the theorem for s = t. This completes our proof of Theorem II. We have thus obtained certain necessary and sufficient conditions to be satisfied in order that a set of elements (...,2V..**>...) of J? may be the coordinates of a &-space. In the course of doing this we have found certain conditions—the ^-relations—satisfied by the coordinates of a fc-space. This raises the question of other relations satisfied by the coordinates of a fc-space, and the complete answer to this question is the subject of the next section. 7. The basis theorem. This section is devoted to the following theorem: Theorem I. If F(P) is any homogeneous polynomial in P^.,,^ such that F(p)-0 for all Sk, then F(P)= 2-^...^,...^(^)¾...¾¾...¾^^). (1) is i where -^...^,^...^(^) is ^e quadratic form defined in §6, (2), and Aix tik Jo ...jk+l{P) is a homogeneous polynomial in Pio.,.ik- For the purpose of this proof we define a generic fc-space in [n to be the fc-space defined by k + 1 points (½....¾) (i = 0,...,^),
316 Vn. GRASSMANN COORDINATES where the bj are (n+ 1) (k+1) independent indeterminates over the ground field K. Clearly, if F( P) is any form such that '< bl • b% 6^ 6* r this relation remains true when the bj are specialised in any way; in particular, when the points (&£,..., bln) are specialised so that they become 1c +1 independent points of any assigned fc-space. Hence F(p) = 0 for every Sk. Conversely, if F(p) = 0 for every 8k9 F{ \ bl • 6?J • • • bl . bl ) vanishes for all specialisations of the b]. Hence, regarding this last expression as a polynomial in the (n+l)(fc-f 1) indeterminates bj> since it vanishes for all specialisations of these indeterminates, it vanishes identically [III, §8, Th. I]. Thus in order that F(P) should satisfy the conditions of the theorem, it is necessary and sufficient that F(p) = 0, where (...,piot ik,...) are the coordinates of a generic fc-space. Now let F(P) be a form satisfying the conditions of the theorem. We group together all the power products in F(P) which are isobaric in each of the suffixes 0,...,n; that is, we consider all the power products in which the suffix 0 appears A0 times, the suffix 1 appears Ax times, and so on. If the sum of such a set of terms is denoted by/Ao...AnCP), Now let bj (i = 0, ...,Jc;j = 0,...,n) be (n+l)(k+l) indeterminates, /90,...,/?n n+1 additional indeterminates, and let %...ik Piobl • Pikbl Piobl • PM = Pi*Pi1-PikPi*...ir
7. THE BASIS THEOREM 317 Then (...,2^...<*»•••) and (...,&0...^,...) are coordinates of generic ^-spaces. Hence 0 = F(q) = S/ao...a„(?) = Xpk-Pkf^.xnip)- A Since p0,..., pn are independent indeterminates over K(b)), it follows /A0...Anb) = 0 for each set A0,..., An. Hence Theorem II. Any form F(P) which satisfies the conditions of Theorem I is the sum of forms isobaric in each of the suffixes 0,..., n which satisfy the conditions of Theorem I. In proving our first theorem we may therefore confine ourselves to forms F(P) which are isobaric in each of the suffixes. We now proceed by double induction. First, we assume that k is fixed, and let n vary. If n = &4- 1, an Sk is a prime, given by a single linear equation , , n and its coordinates are Po...i-u+i...k+i = (-l)fc"i+1^ = (-l)k~i+1at. Since the ai can take arbitrary values independently, there is no non-zero form F(P) such that F(p) = 0 for every 8k in [k+ 1]. We may therefore assume, as hypothesis of induction, that Theorem I is true for the coordinates of fc-spaces in [n— 1]. Equi- valently, we may assume that the theorem is true for forms F(P) in which none of the Pio_ik actually present has n as a suffix. We now suppose that F{P)y which is isobaric in each suffix, is isobaric of weight r in the suffix n, and is of degree t in Pi0,„{k. By the hypothesis of induction made above, Theorem I is true if r = 0. We therefore make the further hypothesis of induction that the theorem is true for forms F(P) which are isobaric of weight less than r in the suffix n. We write dF(p) dPio~.tk to denote the result of replacing P{0.„ik by the coordinates pio„tik of an Sk in dF(P)
318 Vn. GRASSMANN COORDINATES Let bj (i = 0,...,k; j = 0,...,n) be independent indeterminates, and let .¾ bl b°- 1 uik 6* Then if O(P) is any form in Pio.„ik, {0(p) does not necessarily vanish), we write 9(b) = Q(P), and we find that where Mb) v w 8C(j)) 36j ^....^ = (-1)^ Hence Since *L. &L /(6) = *•(*) = 0, (2) and the 6- are independent indeterminates, /(6) is identically zero, and therefore & a/7A\ i=0 S6* By (2) we can write this in the form dF(p) Hence the n forms (¾ = 0,...,n—1) Lz0. ..^-1^ satisfy the conditions of our theorem. Since each of them is isobaric in the suffix n of weight r—l, we may write, by the hypothesis of induction, dF{P) l ujrlo...lk-\n where h = 0,...,?&—1. = 2^L.**.*o...i*+,(p)^...^o...*+1(n (3)
7. THE BASIS THEOREM 319 We propose to deduce (1) from (3) by operating on each side of (3) with • d -^h == 2j *mo...mk-in 3p » \ ) and then summing with respect to h from h = 0 to h = n—1. We first consider the effect of this operation on the left-hand side of (3). +nYYYP P d°F(P) A-OmJ °-rmo...mt_1At;-f7o — '*-i» d*F(P) "• 2j 2j 2j^m0...m;fe_in^/0.../^_1A ; _vyP' P *FW Z-» ^-4 mo...mi-!n-4/0.../|:-in op op m Z cy^wio...mjt_1nu^/0.../*-in = [nr-t(t-1)] jP(P) + .^ S S-^...mij-in-no...'*-i*ap jap > ( ) fc=0 m / Vjrm0...mjc^1hCI^l{i..Jic-in using the isobaric property of F(P), and Euler's theorem on homogeneous polynomials. Now let a0,..., afc be a set of numbers chosen from (0,...,n), and consider the terms of the summation in (5) corresponding to h = at% Then the corresponding terms in the summation in (5) are d*F(P) iZPa0...ak*Pi0 I v P P ...h-in Li=0 J v M(P) p p _ __J^(P)__ ^*2P r)P ^/0.../^-1,^00...^^/) / Ujrao...a*Ujt/o...^-in by §6, (2).
320 VH. GRASSMASTN COORDINATES Hence, = T(n-T+l)F(P) + WPa<t...akPl0...li-1n?p ^P d*F(P) a I °<ra0...akCjrl0...lk-in ^r(n-r+l)F(P) + TlPl^lk_l d (SP ff<£) I C'rl0...lk-in\a °^ra0...ak/ _vp dF^ 92i^(P) ~*2-i 2-i ;fp op -*V../*-i,nao...a*' "/ = T{7l-T+l)F(P)+TtF(P) -TF(P) v v WF(P) F (p, ^^7\p 2P jr/0.../Ar-i,na0...aifcVjr/ a Z ^-^0...^^-^0.../1.-^ = T(n + t-T)F{P) ""2 S ^p sp -^...fc-i.wao...a*(-P)- (6) Since f ^ r, and we are assuming that r > 0, the coefficient of F(P) in (6) is not zero. We now consider the effect of operating on the right-hand side of (3). We have (P)l From this equation, and equations (3) and (6), we derive (1) if we can finally prove that
7. THE BASIS THEOREM 321 is expressible as a linear combination of forms ^...^,60...6^(^) for all choices of ix,..., ik> j0,..., jk+1 and h. (i) If h is different from ^,..., ik and from j0>... ,jk+v clearly (ii) If h = il9 say, but h^jQ,..., jk+v it is easily seen that (iii) If h = j0, say, but h * tlf..., ^, ^i...<fcfc.../*+i(f)] = ^1...^^^...^+1(^)- (iv) Finally, if ft = i2 = j0, say, ^[^...^^0-..^+1(^ = ^^...^,^...^+1(^) + ^1...^^^...^+^^)- This completes the proof of Theorem I. *
* CHAPTER VIII COLLINEATIONS In this chapter the ground field K is again assumed to be commutative and without characteristic. It will not be necessary at first to assume that it is algebraically closed but, later, results will be obtained on the explicit assumption that K is algebraically closed. 1. Projective transformations. Let 8m, Sn be two distinct projective spaces over the field K, of dimensions ra, n respectively. Fix an allowable coordinate system x in 8n and y in Sm. Then consider the set of equations n Vi = E*«j*j (i = 0, ...,m; atj in K), which we usually write in matrix form as V = Ax. (1) We may assume that the (ra+ 1) x (n+1) matrix A is not the zero matrix. The equations (1) associate a point (y0>...,ym) of 8m with any point (x0,..., xn) of Sn which does not satisfy the equations n 2 ayZj^O (i = 0, ...,m). These points, which we call the exceptional points of Sn, fill a linear space of dimension n — r, where r is the rank of A. If r = n+ 1 there are no exceptional points, and (1) is said to define a non-singular projective transformation of Sn on 8m. In order that such a transformation should be possible, we must have If r<n, the equations (1) are said to define a singular projective transformation of 8n on Sm. Now let x = px. (2) define an allowable transformation of coordinates in 8n, and let y = Qy* (3) define an allowable transformation of coordinates in Sm. In
1. PROJECTIVE TRANSFORMATIONS 323 the new coordinate systems the projective transformation (1) becomes _ A _ Qy* = APx*, that is y* = Q~lAPx*. By II, § 4, Th. II, we can find P and Q so that (o o)' where r is the rank of A. Calling r the rank of the transformation, we have proved Theorem I. By a suitable choice of coordinate systems in Sn and in 8m any projective transformation of rank r from Sn to Sm can be taken in the form yi^xi (i = 0,...,r~l), Ui = 0 (i^r). We call this form of the equations (1) the canonical formy and in discussing the geometry of a projective transformation it is usually convenient to use the canonical form of the equations of the transformation. Before doing this, we consider the points of Sm which lie in the prime m 2^ = 0. o In the projective transformation (1) these are just the points which arise from the points of 8n lying in n ^UjXj = 0, 0 m where ut = £ v^ (j = 0,...,n). We write these last equations in matrix form as u = A'v, (4) where A' is the transpose of the matrix A. These equations define a projective transformation of the primes of Sm on the primes of Sn. We call this transformation the dual of the projective transformation (1). The transformations of coordinates (2) and (3) induce the dual transformations u* = P'u, v* = Q'vy 21-2
324 VIII. OOLLINBATIONS respectively. Therefore the dual transformation (4) becomes, in the new coordinate systems, u* = P'u = P'A'v = P'A'iQ'^v* ^(Q-iAPyv*. The relation between a projective transformation and its dual is therefore preserved under transformations of coordinates. This result is otherwise evident from the geometrical definition. If the point transformation is in canonical form, the dual transformation is in the canonical form ui = vi (i = 0, ...,r~l), ut = 0 (i^r). Before discussing the geometry of projective transformations for arbitrary m, n, r, we consider the simple case in which m = n = r—1. The canonical forms are then Vi = Xi {i = 0,...,n), ui^vi {i = 0,..., n). To each point of Sn (prime of Sm) there corresponds a point of 8m (prime of 8n), and conversely. The inverse correspondence is also a projective transformation of Sm on Sn. Linearly independent points or primes correspond to linearly independent points or primes, and, similarly, linearly dependent points or primes correspond to linearly dependent points or primes. Hence, any Sk of 8n corresponds to a unique fc-space S'k of Sm, and any S'k of Sm arises from a unique Sk of Sn. If Sk is determined by the points (bl..., b\) (i = 0,...,&) of Sny and has Grassmann coordinates (...,2¾ mmik,...), Sk is determined by the points (¾....½) (* = 0,...,¾) in Sm. Hence its Grassmann coordinates are (...,(7^...^,...), where Qi0...ik bi • % bl • bl = Pi0...iy It follows that the projective transformation carries the Sk with coordinates (--->jft0 ...<*>•••)> into the Sk of Sm with coordinates
1. PROJECTIVE TRANSFORMATIONS 325 (...,pio_<t,...), for any value of k. In particular, if fc = ra—1, n~x {^,..., un) of Sn is transformed into the S^^ (u0,...,un) of #m, and this transformation of the primes of Sn into the primes of 8m is the inverse of the dual projective transformation of the primes defined above. We now return to the study of the projective transformation of rank r of Sn on Sm, where m, n are arbitrary, and r > 0. We choose coordinate systems in both spaces so that the projective transformation is given in canonical form by the equations yi = xi (i = 0,...,7--1), (5) 2/i = 0 (%>r). \ v ' The dual transformation is then given by the equations ui = vi (i = 0,...,r~l), (6) Ui = 0 (i^r). \ v ' The exceptional points of Sn, that is, those which have no transforms, are the points of the $„_,. xt = 0 (i = 0,...,7--1). We call this the singular space of Sn. The locus of points in Sm which are transforms of points of Sn is given by the equations 7/,. = 0 (i>r), and is therefore a linear space of r — 1 dimensions, 2Jr_x say. This we call the vertex of the transformation. Now consider the dual transformation (6). It is also of rank r: and the primes of Sm which have no transforms are given by Vi = 0 (i = 0,...,7--1). These primes intersect in the vertex 2Jr_v Similarly, the primes oi Sn which are transforms of primes in Sm are the primes which pass through the singular space 8n_r. The equations (5) define a non-singular projective transformation T* of the space 5fr_x of Sn, given by the equations xi = 0 (i > r),
326 VIII. COLLINEATIONS on SrmmV and the equations (6), for i = 0,...,r— 1, define the dual of T*. The properties of this non-singular transformation from #,_! to Er_t follow at once from our previous discussion of the non- singular case of projective transformations. We now consider how the properties of the transformation (5) follow from a knowledge of the properties of T*. Two points (x0,..., xn) and (x'0>...,x'n) of Sn, which are not in the singular locus /Sn_r, are transformed by (5) into the same point of 8m if and only if xi = Xx'{ (i = 0,...,r—1). Geometrically this condition indicates that the points have the same projection from the singular locus Sn_r on Sr__v this projection being (x0, ...9xr_l9 0, ...,0). Now, each point P of Sn which is not in Sn_r is joined to #„_,. by a space Sn_r+1. It follows that a point Q of 8n which is not in 8n^ has the same transform as P in 8m if and only if Q lies in the #n_r+1 defined by P. Thus, to any #n_r+1 in Sn wh^ch passes through the singular space 8n_r there corresponds a unique point in 27,.^, and conversely. We may call this correspondence between the #„_,.+! through Sn_r and the points of i7r_1 a projective correspondence. It is easily seen that in this correspondence the (n — r+1) -spaces of Sn which pass through the singular locus Sn_r and lie in an (t& —r + &+ l)-space through the same 8n_r correspond to the points of a i-space in Zr_v In particular, primes of 8n through the singular locus correspond to (r — 2)-spaces in Zr_v The properties of the dual transformation (6) from 8m to 8n are obtained in a similar manner. If 77 is any prime in Sm which does not contain the vertex Sr_ly and if II' is the corresponding prime in Sn, which passes through the singular locus Sn_r, a necessary and sufficient condition that another prime nx of Sm should have the same transform IT' as 77, is that 77 and 77x meet the vertex Er_x in the same (r— 2)-space. To each Sr_2 in Er_x there corresponds a unique prime in 8n through the singular space Sn_r. This correspondence between the (r — 2)-spaces in 27r_x, and the primes of Sn through the singular space 8n^r is precisely that obtained above by the direct transformation. Now let us consider what happens when 8n and Sm, instead of being distinct spaces, are the same space, which we denote by [ri]. Then the transformation (1) takes a point P of [n] into another point P* of [ri], and, more generally, an 8k of [ri] which does not meet the singular #n_r is transformed into another Sk of [ri].
1. PROJECTIVE TRANSFORMATIONS 327 It is important not to confuse this type of transformation in [ri] with a transformation of coordinates. A transformation of coordinates is a relation between two sets of coordinates for the same point. The present correspondence is a relation between the sets of coordinates of two different points. A transformation of type y = Ax which transforms [ri] into itself is called a collineation in [ri]. Projective geometry is the study of the properties of configurations which are invariant under collinea- tions. In the study of collineations we can deduce, by what has been said above, properties of the transform of* any configuration, but we have a new problem to consider, that of the relation of a configuration to its transform. The remainder of this chapter is devoted to a study of collineations. 2. Collineations. In the space [ri] over K we choose an allowable coordinate system and consider the collineation given by the equations n yi= Ii<t>ijZj (i = 0,...,n), or, in matrix notation, y = Ax. (1) Let a transformation of coordinates in [ri] be given by x = Px*. The coordinates of the points x> y are transformed into x*, y*y where x = Px*, y = Py*y and therefore in the new coordinate system the collineation has the equations „ = p-iAPx+% The collineation dual to (1) is (cf. § 1) u = A'v, (2) u = (uQ,...yun) being the prime which contains every point x whose transform y by (1) lies in v = (vQ,..., vn). If A is non-singular, (2) can be written as r /0v v = Auy (3) where A = (A')*1 is the matrix complementary to A. The collineation (3) transforms a prime u into the prime v which contains the transforms y of all the points x of u, and no other points.
328 VIH. OOLLINBATIONS The transformation of coordinates x = Px* changes (2) into w* = (P-^Pj't;*, and (3) into v* = (f^AP)u*. Having seen the effect of a transformation of coordinates on equations (1), (2) and (3) we now define similar coUineations of [n]. Two coUineations y = Ax, and y = Bx are similar if there exists a non-singular matrix P with elements in K such that D Dnn B = P-*AP. This relation is clearly reflexive and symmetric, and if C = Q-XBQ (Q non-singular), then C = (PQ)^A(PQ), and hence the relation is also transitive. It is therefore an equivalence relation. The relation between similar coUineations is simply that there exists a non-singular projective transformation x = Px* of [n] into itself which transforms one collineation into the other. Our main problems here are (i) to determine necessary and sufficient conditions that two coUineations be similar, and (ii) to find a set of canonical forms for coUineations in [ri], that is, a set of coUineations with the properties: (a) no two coUineations of the set are equivalent; (b) every coUineation is equivalent to one member of the set. The canonical forms are chosen so that it is a simple matter to read off from them the geometrical properties of the coUineations. We now prove Theorem I. A necessary and sufficient condition that the coUineations A D y = Axy y = Bx should be similar is that the \-matrices A — \In+1 and B — \In+1 should have the same invariant factors (or, in the case in which K is algebraically closed, the same elementary divisors).
2. OOLLINBATIONS 329 By II, § 9, Th. Ill, a necessary and sufficient condition that the A-matrices A— AJn+1 and J3 —A/n+1 should be equivalent, in the sense of that section, is that they have the same invariant factors. When this is the case, B-\In+1 = M(A-AIn+1)N. But by II, § 9, Th. V, this equation implies the existence of non- singular matrices P, Q over K such that It follows that and Hence and therefore B- -Mn+1 = P(A-Mn+X)Q. B = PAQ, 4+x = PQ- p = g-1, B = Q-iAQ. Now let K be algebraically closed. We note first of all that det {A - A/n+1) and det (B - AJn+1) are each polynomials of degree n +1 containing the term (_ 1)^+1 An+1. These polynomials are therefore not zero polynomials, and rank (A — A/n+1) = rank (J3 —A/n+1) = n+ 1. Hence the form of Theorem I which is applicable when K is algebraically closed follows from II, §9, Th. IV. Now let E±{\)9 ...9En+1(\) be the invariant factors of A—\In+ly and let 2^(A) be of degree rt. Since (_l)n+innW) = det(A-Mn+l), it follows that »+* 1 and since E{(\) is a factor of Ei+1(A)9 We suppose that _ _ A and that r8+1 > 0.
330 VHI. OOLLINBATIONS Let E^X) = Ar*+aa A'*-1 + ...+airi (i > «), and consider the rt x rt matrix St- 0 0 0 ain 1 0 0 . 0 1 • • 0 ' 0 -ailt We prove that the invariant factors of Bi — XIri are 1,1,..., 1,2^(A). For consider the determinants of the submatrices of t rows and columns extracted from Bt — Mr{. If 1 < t < ri we consider the particular submatrix whose elements lie in the 1st, 2nd, ..., tth rows of Bi — \Ir.) and in the 2nd, 3rd, ..., (t + l)th columns of this matrix. It is clear that the determinant of this particular t x t submatrix is equal to unity. Hence the highest common factor of determinants of t x t submatrices extracted from Bi — XIri is of degree zero, if 1 ^ t < r^ But if t = riy det(5,-A^)-(-l)^^(A). This proves the result. Now let B be the (n + 1) x (n + 1) matrix B = /B 8+1 0 B 0 8+2 0 0 0 (4) We prove that the invariant factors of B — A/n+1 are ^(A),..., ^+1(A),..., En+X(\). We can perform elementary transformations [II, § 9] on the rows and columns of any submatrix Bt — \Irt (s+l^t^n+l)of B — XIn+1 without altering the rows and columns of any other submatrix, since these submatrices have no rows or columns in common. Hence each submatrix can be reduced to its canonical form by elementary
2. COLLINEATIONS 331 transformations on B — XIn+v By another elementary transformation this matrix can be reduced to the matrix (/,0.0 0 EM(X) . 0 • • • • 0 . . En+X(\) Since E{(A) is a factor of i^+1(A), the highest common factor property of invariant factors makes it evident that the invariant factors of this last matrix are 1, 1, ..., 1, E8+1(X), ..., i£n+i(A), that is, ^(A), ..., E§(\), EM(\), ..., En+1(\). Thus B — AJn+1 is equivalent to the matrix A—\In+1, since it has the same invariant factors [II, §9, Th. III]. We shall take the canonical form for collineations similar to (1) as y = Bxy where B is given by (4). Hence we have Theorem II. If the matrix A — A/n+1 has the invariant factors -E^A), ...,-E?n+1(A), and if the matrix B is defined as in (4) above, the collineation y = Ax (1) is similar to the collineation y = Bx, and this latter form of the equations may be taken as the canonical form for collineations similar to (1). When the ground field K is algebraically closed, and we use elementary divisors in place of invariant factors, it is usually more convenient to use another set of canonical forms for collineations. Let A — A/n+1 have the elementary divisors (A-a!)V..,(A-a&)S where, for the moment, it is not necessary to distinguish the a^ which are equal from those which are unequal. By the properties of elementary divisors obtained in II, § 9, ei>0 (i = !,...,&), and ^+...+ek = n+1.
332 vni. OOLLINEATIONS Now let Ce(a) be the e x e matrix Ce(a)= /a 1 0 . 0 a 1 . • 4 • • 0 . . a 1 ^0 . . 0 a) Then the matrix Ce(ot) — XIe has the one elementary divisor (A — a)e. For det (Ce(a) - AJe) = (a - A)e, whilst the determinant of the (e— 1) x (e— 1) submatrix obtained by omitting the last row and the first column of the matrix under consideration is equal to unity. Now let C be the (n + 1) x (n + 1) matrix C= /CeiK) 0 . 0 \ 0 G€2(a2) 0 J. (6) 0 . . Cek(aJ By methods exactly similar to those used in the proof of Theorem II we can show that C — A/n+1 has the elementary divisors (A~a1)S...,(A~a&)cA;. This gives us Theorem III. If A — A7n+1 has the elementary divisors (A~a1)%...,(A-afc)S the collineation (l)is similar to the collineation t/= Cxy where the matrix C is constructed as in (5) with the given elementary divisors. This collineation serves as a canonical form for collineations similar to (I). It should be noted, however, that geometrically the two collineations A y = Ax, and py = Ax (p4=0),
2. COLLINEATIONS 333 if given in the same allowable coordinate system, perform the same transformation on the points, lines,..., of [ri]. From this it follows that two collineations y = Ax> and y = Bx are similar in a geometrical sense if the elementary divisors of A — A/n+1 and of B — \In+1 are (A -^)^,..., (A -otk)% and (A~^)/i,...,(A~^)/* respectively, where, for a proper arrangement of these divisors, ei=U fii = P*i (i= 1,. ..,&), [i being a non-zero element of K. In the remainder of this chapter we shall assume that the ground field K is algebraically closed. 3. United points and primes of a non-singular colUneation. If we are given a colUneation in [n] over an algebraically closed field K we may, as in § 2, choose our allowable coordinate system so that the equations of the colUneation are where y = Gxy (1) c = /cClK) o . o 0 CCa(a2) . 0 0 0 . Cek(ak)j and Ce{a) = /a 1 0 . 0 a 1 . . . a 1 v0 0 0 . a> The elementary divisors of the matrix C — A Jn+1 are (A-a^, ..., (A-a^*.
334 vm. OOLLINEATIONS It is usual to call them the elementary divisors of the collineatioi (1). We note that if the original collineation is there exists a non-singular matrix P such that C = P~*AP. Hence C- A/n+1 = P~\A - A/n+1) P, and therefore det(C-A/n+1) = det(^-A4+1) = (-l)n+1Il(A-a^. 1 Hence, putting A = 0, det G = det^4 = af1... af.*. We shall assume in this section that the collineation is non-singular, that is, that , , A , - ' det A 4=0. It follows that each latent root OLi is different from zero. In the canonical form (1) the submatrices C^OLj) may be arranged in any order along the principal diagonal of C, but it will be convenient to order them so that (i) those submatrices (7^(0¾) which correspond to equal latent roots at come together, and (ii) if <xt = a,] and ei < ej then C^fa) precedes GAolj). We shall assume that the submatrices are arranged in this way in all future applications. A point (z0,..., zn) of [n] is called a united point of the collineation (1) if it is transformed into itself by the collineation. The condition for this is ^ Cz = pz> where p is a non-zero element of K. The equations (C-pIn+1)z = 0 have a solution (z0,..., zn) if and only if [II, § 8] det(C-/>4+1) = 0. Hence p must be a latent root of the collineation. Let us take p = av and suppose that at = a2 = ... = cciy fci + ty (i>t).
3. UNITED POINTS AND PRIMES 335 The equations to determine the united points arising from /) = 0¾ are axz0 = cciZ0 + zly\ 0^1 = 0^ + 32,1 (2-1) alz«j,-l = alzcx-l> J alzcx — a2zC! + ze!+l> alze1+c8-l ~ a2zC!+c8-l> J (2-2) al2«i+...+e»-i ~ a/2e1+...+«i-1 + 2ei+...+«-i+l> alzei+...+e*-l ~ a/2ei+...+«r-l> al ^+...+^ = ^+1^+...+^ +^+...+^+1 > • •••••• alzn = a*zn- From equations (2-1) we have, successively, Zi = 0, z2 = 0, ..., zei_i = 0; from (2-2) we have, successively, since ax = a2, *«i+i — 0, ..., zc .c j — 0; from (2-1) we find that % + ...+<*_!+! -o, •> ^+...+^-1 -1 = 0; (2-0 if, finally, we consider the last equation and work backwards we find that %+...+«*+! = 0, ..., zn = 0. Hence a united point (z0, ...,zn) arising from the latent root ax necessarily has all its coordinates zero except Conversely, we can verify at once that if (z0,...,zn) has all its coordinates zero except those named above, it satisfies the equations (2), and is therefore a united point.
336 VIII. COLLINBATIONS If A0,..., An are the vertices of the simplex of reference of our coordinate system we deduce from these results that the united points corresponding to the latent root ax of G form the linear space of t—l dimensions spanned by the points A0,Aei,Aei+e%,..., -4^+.,.+^-1- ^e ca^ ^k*8 sPace the fundamental space of the collinea- tion corresponding to the latent root av The united points corresponding to any other latent root of C are found in a similar way. The following rule is immediately obtained: Theorem I. Let a{ be any latent root of C, and let Ceq(*Q)> Ceg+1K+l)> •••> ^+,(««+•) be the submatrices of C which correspond to latent roots equal to 0¾. Then the united points corresponding to 0¾ form the fundamental space spanned by the points A A A and therefore lie in a space of dimension s. Since each fundamental space is determined by a set of vertices of the simplex of reference, and since also no two sets corresponding to distinct latent roots have any vertices in common we deduce Theorem II. The fundamental spaces defined by distinct latent roots have no points in common. Indeed, they are linearly independent; that is, if their dimensions are m1,...,mr their join is a space of r r — 1 + 2 mi dimensions. 1 Again, we see from Theorem I that the join of all the fundamental spaces is the space spanned by the vertices AAA A ^■O* X1c1> ^0i+ei> •••> •••> ^Lc1+c1+...+Cjt-1> and is therefore a space of k — 1 dimensions. Therefore r r-l + 2%= fc —1. 1 The join is the whole of [n] if and only if Jc = n + 1, that is, if and only if ______ 1 This case, in which all the elementary divisors are linear, will sometimes be referred to as the non-special case; the special case being that in which at least one of the elementary divisors is of degree greater than one.
3. UNITED POINTS AND PRIMES 337 The collineation (1) transforms any &-space Sk into another &-space S'k, this space consisting of all points y which are transforms of the points x of Sk. The reasoning is exactly that used in the case of a non-singular projective correspondence between two distinct n-spaces. In particular, when Jc = n— 1, if Sk is u = (u0,...yun), and S'kia v = (v0,...,vn), the correspondence is given by the equations v = On. (3) The self-corresponding primes of this collineation are obtained by finding primes w = (wQ,...,wn) such that pw = Cw (/>4=0). We are thus led to consider the elementary divisors of the matrix C-XL Since 0= /Cja,) and since 0e(a) = ar •a ar 0 0 -2 n+V 0 0*(«1> 0 0 a-1 — a- 0 0«(«*V 0 0 ar A-lf-^a-' -2//-«+l (_l)e-2a- it is easily calculated that the elementary divisors ofC — A/n+1 are (A-ar1)*, ..., (A-a»*)*. However, it will simplify our notation if we observe that a necessary and sufficient condition that Cw = pw (for some p =t= 0) is that C'w = <rw (for some <r=|= 0). Hence we need only consider the united primes of the dual transformation u = G'v (4] HP I 22
338 VIII. 00LLINBATI0NS to obtain the united primes of the collineation (3). The elementary divisors of <7' — AJn+1 are evidently the same as those of C —AJn+1, that is /\ x^ /> w (A-a^i, ..., {X-OLk)eK Hence the following equations determine the united prime (w0,..., wn) corresponding to the latent root ax: octw0 = axw09 a1w1 = wQ+ 0,^^ «i*V-i = ^-2 + ^1^-1 J a±wei = a2wCi, • ••••• alwe,+e,-l = ^+^^ +a2Wei+ei_v (5-1) (5-2) ai ^6,+...+6,-, = a/^«1+...+«_,» al ^+...+^-1 = ^+...+^-2 + ^^+...+^-1^ ai ^+... + 61 — ^+1^+...+^ ai^n = ^n-l + ^'^n- From (5-1) we find, successively, that and then, from (5-2), ..., (5-£), we find that, since ax= ... = <xt, w, ^,-2 = 0, ...,11^ = 0, a == 0, ..., w( and finally that ^+...+^-2 ^+...+^. . = °. wf = 0 (i>c1+... + e<). (5-0 Hence the united prime necessarily contains the points A A A A >4 /4 /1 A ■/:le1+...+c<-1> • • • 9 ^ei+...+e*-2> -^^+...+^ • • • > ^n> and, conversely, any prime which contains these points satisfies (4).
3. UNITED POINTS AND PRIMES 339 These points determine a linear space 8n_ly and the primes through it form a star with vertex Sn_t. This star of primes is the dual of the set of united points in the space S(^ly and the properties of the star may be determined by duality from the properties of the set of united points corresponding to the latent root olx. Corresponding to each distinct latent root of C there is a star of united primes, and we see that there is a rule, similar to Theorem I, for determining the star of united primes corresponding to a given latent root. Theorem III. If at is any latent root of C (or C) and are the submatrices of G which correspond to the latent root a^ of C, then the united primes corresponding to 0¾ form the star whose vertex is the join of the points A A A A A A A A -^+... +eq+8-\> ' -^-^+...+^+,-2^ -^+...+eg+*> "-><n-n* We have agreed to place 0^(^) before Ce (a^) in C if at = a^ and ei<ej; suppose that eq = • • • = eq+t = 1 > eq+t+l> • • • > ^q+s > 1 • Then, by Theorems I and III we see that the space S8 of united points corresponding to the latent root at and the vertex Sn_8_t of the star of united primes arising from the same latent root intersect in the space S^^ defined by the points A A -^-Ci+.-.+e^+p •••» -rie1+...+c?+,_1* Similarly we observe that the space of united points arising from any latent root a^ (a^ 4= aj lies in the vertex of the star corresponding to the latent root at. Hence we have Theorem IV. The fundamental space of united points arising from any given latent root a lies in the vertex of the fundamental star corresponding to any other latent root, and meets the vertex of the star corresponding to the same latent root a in a space of p — 1 dimensions, where p is the number of non-linear elementary divisors of the type (A-a)€(e>l). As an immediate corollary, we deduce that if a collineation is non- special the vertex of the fundamental star corresponding to any given latent root is the join of the fundamental spaces corresponding to the other latent roots. 22-2
340 VHI. OOLLINBATIONS Now let Si^1 and Sn„( denote the fundamental space and the vertex of the fundamental star, respectively, corresponding to the latent root ax of C, and suppose that 1 = ex = ... = e8<e8+1^e8+2^...^et. (6) Let x be any point of [n], let y = Cx, and z = y-axx = (C-a^+Jx. (7) Evidently z is collinear with x and y, and from (7) it is the transform of x under the collineation with matrix C —#1^+1- Since det (C — axIn+1) = 0, this collineation is a singular collineation. The point z is uniquely defined by (7) unless Cx = axxy that is, unless x lies in St_v If x is not in 8t_x then we see from (7) that the coordinates of z satisfy the equations z0 = zl = ••• == z8-l = zs+e,.n-l = ••• = z8+c,+1...+er-i = 0. That is, z lies in Sn_(. It is also easily seen that if x* is any point of the space S( which joins x to S(_x> the transform of x* by (7) is also z, and the correspondence between the ^-spaces through St_x and the points z of Sn_t is a projective one. We prove this result, which should be compared with a similar result in § 1, using the symbolic notation for points. If the points Pq>Pv --^^-1 form a basis for St_v then 0(Pi)^<t1Pi (*' = 0,...,*-l). Let P be any point not in St_v Then any point x* in the 8t joining P to #{_! is given by 0 for suitable A^, /1 in K. Now Cz* =^0^)+/10^) 0 0 and therefore z = Or* - a^* = /*[C(P) - ax P], (8)
3. UNITED POINTS AND PRIMES 341 which is independent of the particular #* chosen in the #,. If P is chosen in an S'n_( which does not meet St_v then P determines and is uniquely determined by S(. Since 8(_x is the join of the points A0,A19 ...,A8_V A„A8+es¥i, ...,4^+,,^,, we may take P to lie in the S'n_t spanned by the remaining vertices of the simplex of reference. The equations (8) determine a unique point z in 8n_^ if P is taken anywhere in S'n_t. Conversely, if we write down these equations, assuming that z is in Sn__(, we see that we can solve for the coordinates of P, and that the solution is unique if we assume that P lies in S'n_t. Hence the correspondence between f-spaces through 8(_1 and the points z ofSn_( may be described as a projective correspondence. Now let us consider the effect of the original collineation (1), y = Cx, on the 8t determined by P and the space of united points S(_v Since t~i y=Cx* = 0^2^+/iC(P), o the points of St are transformed into the points of the S't determined by the transform of P under (1) and S(_v Since z = y-^x* = [i[C(P)-axP] gives a point in Sn_t which is independent of the point x* chosen in St, it follows that the correspondence between S( and S't is a per- spectivity, vertex z. Now let us assume that in (6) we have s < t, so that the collineation (1) has non-linear elementary divisors corresponding to the latent root <xv Then St_t and Sn_( have a space of t — 5—1 dimensions, S(-8-v say> in common. If z is in St_8_v and Stis the unique f-space through St_t which corresponds to z in the projective transformation described above, the transform of St under (1) is in perspective with St, vertex z. Since this transform contains S(_v and z lies in St_v the transform must coincide with St. Algebraically, if 2 z=z Vz0» •••» Zn)> each zt except is zero, and if we solve (8) for the coordinates (#0,...,xn) of P, not restricting P this time, we find that r = (Xq, X^y..., x8_i, x8i Z8, 0,..., U, #8+e#+1> z«+e,+1> 0, ..., 0, ..., %8+e8+1+...+et-i> zs+es+1+...+et-i> ®> " *" > ®/>
342 Vni. COLLINEATIONS where each xi can be chosen arbitrarily. These points P fill the space St which corresponds to z. This expression for the coordinates of P makes it evident that the collineation (1) transforms P into a point with coordinates of the same form. Hence, again, St is transformed into itself by (1). If we anticipate Theorem VI, we can say that the collineation (1) induces a collineation in St. This induced collineation is such that /¾ contains an S(_t of united points corresponding to the latent root ax. The number of elementary divisors (A — at)fi must therefore be t. t Since also J/* = t+1, there being no united points of (1) in St l outside St_19 it follows that the elementary divisors of the induced collineation consist of £—1 elementary divisors (A — ax), and one quadratic elementary divisor (A — ax)2. We return now to the consideration of (7), x being any point of [n] which is not in a fundamental space. Its transform by (1), that is, the point y = Cx, is distinct from x. We consider the p points y-*iX {i= 1,. ..,/>), where p is the number of distinct latent roots of C. These are the points in which the line xy meets the vertices of the fundamental stars corresponding, respectively, to the latent roots ocv ...,ap. This range of points, together with the points x and y, is projectively related to any other range of p + 2 points constructed from the collineation (1) in the same way [VI, § 10]. We remark that if the line xy lies in a vertex of a fundamental star, the points x and y = Cx must he in the vertex. Conversely, if x lies in a vertex of a fundamental star, say in that corresponding to the latent root otly so that x0 = xl = ••• = ^8-1 = X8+es+1-l = ••• = #8+e,+1+...+e*-l = 0, then Vo = V\ " ••• = 2/a-i == 2/s+e,+1-i =...= ys+e5+1+...+ct_i = 0, so that y also lies in the vertex. The vertices, in fact, are all united spaces under the collineation (1). This is also evident from the fact that any prime through a vertex is a united prime, and the intersection of any number of united primes is necessarily a united space. Returning to the ranges of points discussed above, we have Theorem V. If any point x of [n] is transformed by the collineation y=Cx into a point y=M, the line xy meets the vertex of each of the
3. UNITED POINTS AND PRIMES 343 fundamental stars of the collineation. If x does not lie in a vertex, let Aly...,Apbe the unique points of intersection of xy with the vertices. Then if y* is the transform by y = Cx of any other point x* which does not lie in a vertex, and y* #= x*, and ifx*y* meets the fundamental vertices in Af9..., A*, respectively, the ranges x, y,Aly...,Apy and x*yy*,Afy..., A*9 are related. We conclude this section by noting two simple results which will frequently be used in the sequel. Theorem VI. If a space 8k is transformed into itself by a collineation, the points of Sk may be regarded as transformed by a collineation in 8k. The property of a space 8k being transformed into itself under the collineation (1) is invariant under any allowable change of coordinates. For let one of the linear equations satisfied by Sk be, in matrix notation, , _ /rkX ux = 0. (9) Then since the point y = Cx continues to lie in Sk, u'Cx = 0. (10) The allowable transformation of coordinates x = Px* transforms (9) into , * A /m, x ' v x* = 0, (9) where v = P'uy and transforms (10) into v'P-WPx* = 0, (10)' which is just the condition that the prime given by (9)' should pass not only through #*, but through y* = P-WPx*. But this is the transform of (1) under the transformation of coordinates x = Px*. Our initial result is therefore proved. We may therefore choose the coordinate system so that the equations of Sk are _ __ n xk+l ~ U> ..., £n — U. Then if the collineation in this coordinate system is n !fc = Ii<*<ijXj (» = 0,...,^), we have the equations n J^a^Xj = 0 (i = fc+1, ...,n)
344 VIII. OOLLINEATIONS for every point (x0>...,xk, 0, ...,0) of Sk\ that is, for all choices of x0,..., xk. This implies that *<j = ° (* > h j < &)• The transformation in Sk is given by the equations k Vi = liUijXj {i = 0,. ..,&)> j-o which is a collineation in Sk. The only united points possessed by this induced collineation are the united points of the given collineation which lie in Sk. Theorem VII. In any collineation there is at least one united point and one united prime. We need only prove the result for a united point, the result for a united prime following at once by duality. Since any collineation has at least one latent root, and since to each latent root there corresponds at least one united point, the theorem is evident. There are collineations with exactly one united point. As an example, consider the collineation y = ^n+i(a)«- The elementary divisors are (A —a)n+1, and the only united point is (1,0,...,0). 4. The united ^-spaces of a non-singular collineation. A col- lineation' y-Ax (1) induces a transformation, given by VII, § 2, (4), a, ioh a hik a ikh a ikik Pjo...1k (2) on the Grassmann coordinates of the ^-spaces of [n], and we could find the united ^-spaces of the collineation (1) by considering the ^-spaces which satisfy the equations s 1*,...,ik as a a, a, ikh ioJk aikik ^0-..¾ = <r^o ...*»'
4. UNITED fc-SPACES OF NON-SINGULAR COLLINEATION 345 remembering that the coordinates pio,„{k satisfy the quadratic 2>-relations[VII,§6, (4)] k Pio...ikPlo — Jk = Z*Pio:.i9-ihU+i»* ikPfo — h+iiah+i — Jk' A-0 This method of procedure is not, however, very convenient. One difficulty is that a qualitative description of the united points of the collineation (1) in [n] is not, in itself, sufficient to allow us to determine qualitatively the united points of (2), regarded as a collineation in a space of I, I — 1 dimensions. For example, if n = 3, and Vfc+1/ A= /a 0 0 0\ o y? o o 0 0 y 0 \0 0 0 8) where a, /?, y, 8 are all distinct, we have a complete qualitative description of the united points of (1) in the statement that there are four united points, at the vertices of a simplex. But if k = 1, we obtain equations (2) in the form ? = Bp9 where B is the second compound A<2) of the matrix A, and may be written, [VII, §2], = /<*/? / ° 1 ° 1 ° \ ° \o 0 ay 0 0 0 0 0 0 <x8 0 0 0 0 0 0 Py 0 0 0 0 0 0 /tt 0 0' 0 0 0 0 y8{ The collineation in [5] has six united points at the vertices of a simplex, if the elements which appear on the principal diagonal of B are all distinct. But if, say, ay? = y8, the collineation has an infinite number of united points. These do not all satisfy the quadratic ^-relations, and therefore do not all represent united lines in [3]. We therefore use a different method to find the united ^-spaces of a collineation. We begin with the facts [§ 3, Ths. VI, VII] that if an
346 Vm. OOLLINEATIONS 8k is transformed into itself a collineation is induced in Sk, and this has at least one united point, which is a united point of the original collineation. Thus we need only consider those ^-spaces which meet a fundamental space of the given collineation. Let P be any point of the fundamental space corresponding to the latent root av We seek the ^-spaces through P which are united in the collineation. Our search will be greatly simplified by the following theorem: Theorem I. If P is any united point of the collineation (1) an allowable coordinate system can be found in which the collineation is given in canonical form by the equations V = Cx, (3) where 0= /Oei(«i) o o 0 C^o.) . 0 0 0 . Gek(cck)/ and P is a vertex of the simplex of reference. As in § 3 we assume that ctx = ct2= ... = ^4=^. (i>t)y and that ei^ejifi<j^t. By the results of § 2, an allowable system of coordinates can be found in which the collineation is given in the canonical form (3). The fundamental space of united points corresponding to the latent root ax is defined by the vertices of the simplex of reference. Let P lie in this fundamental space, so that P = A0A0 + A1Aei + ... + A/_1-4ei+...+ct-1. Let us suppose that . n / • * \ *^ A$ = 0 (i<ra-l), but Am_x 4= 0. We find a transformation of coordinates which leaves (3) unchanged, and transforms P to the vertex Af1+t+em_l of the new simplex of reference.
4. UNITED ^-SPACES OF NON-SINGULAR COLLINEATION 347 The equations of the collineation (3) have been written out in detail in § 3, (2). We are now concerned only with the particular set 2/^1+...+em-i """ <xmxt1+...+em-iJt~ xex+...+em_l+l> ^1+...+^-1 ~ (Xmxe1+...+tm-l> Vei+.^+em = ^+1^61+...+6^+ X6i+...+6m+l> ^/ci+.-.+em+i-l == am+la'e1+...+em+1-l> (4*m) (4-m+l) ^+...+6^-1 — ^^61+...+^-1+^+...+6^-1+1) 2/ei+...+6^-1 ~ a*x6i+...+6^-1- (**) We perform a transformation of coordinates which affects only the coordinates involved in the above equations. We write XI ^61+...+6t-i+i = ViV-^-1+rVi^+...Hn-i+i 0 = o,...,em— l), and #*+...+e;_i+j = ^771-1^61+...+61-1+^ U = em> •••>£* — 1)> for each value of i from i = m + 1 to i = £. We perform the same transformation on the y's. The equations of the collineation in this new allowable coordinate system are obtained by multiplying the equations (4-i) by Am_x and subtracting from each of the first em of them A^ times the corresponding equation of the set (4-ra). The other equations of the collineation are unaltered. It follows at once that the new equations of the collineation are just the original equations with x*, y* written for x, y, that is, y* = Cx*. It is immediately verified that in the new coordinate system P is the vertex ^?1+...+«TO_1 of the new simplex of reference. Thus our theorem is proved. We now assume that the united point P of the collineation (3) is the vertex A^+tmt+em of the simplex of reference. Consider the line joining P to any point x of [ri]. The collineation (3) transforms this line into the line joining P to the transform y = Cx of x by (3),
348 VHI. OOLLIKBATIONS since P is a united point. The two lines we are considering meet the face of the simplex of reference in the points (x0, ...9xei^+^^l90,xei^+e^_^v ...yxn)9 and (yo> --^+...+^-1. 0, ^+...+^.,+1, ..., yn). If (x0> — »*n-i) and (j/0,..., t/n_i) denote the coordinates of these points in the coordinate system whose simplex of reference contains the vertices which lie in the face xei+_+emi = 0, namely -^■o> • • • 9 -^+...+^-^-1) -^+..>+6wi_1+1,..., -4n, the relation between x and i/ is V = 05, (5) where C is the n x n matrix obtained from C as follows: (i) if em = 1 we obtain C from C by omitting the one row and column of GJ.aJ; (ii) if em > 1 we obtain C from (7 by replacing CeJam) by Cem-i(am)> that is by omitting the (e± + ... + em-1 + 1 )th row and column of C. It is clear that a united line through P is the join of P to a united point of the collineation (5) in the prime xe1+...+em_1 = °> and conversely. More generally, any united &-space through P is the join of P to a united (k— l)-space of the collineation (5) in the face of the simplex of reference opposite P, and conversely. We thus have a means of constructing a theory of the united ^-spaces of a non-singular collineation, by induction on k. A formal statement, in all generality, of the theory of united fc-spaces is somewhat complicated, although it does not involve any serious difficulties, either of an algebraic or of a geometrical nature. It is usually simple enough to obtain the desired results in any particular case. We shall confine ourselves here to a description of the properties of the united lines of [n]. This involves a study of the united points of the collineation (5). We have seen that in the collineation (3) each submatrix C^a^) corresponds to a vertex -4^+...+^ of the simplex of reference which is a united point, and the fundamental space of united points corresponding to a latent root <Xj is the join of the vertices corresponding to submatrices (¾^) for which ^ = ^. All these corresponding
4. UNITED AJ-SPACES OF NON-SINGULAR OOLLINEATION 349 vertices except P = Atl¥mmm+em_1 lie in #^+...+^.^ = 0. It is at once seen that if Ah is the vertex associated with the submatrix GAa^) of C, where j 4= m> then Ah is also the vertex associated with the submatrix GAoLj) of C. When we consider the vertex associated with CCm(am) (remembering that ax = ... = am = ... = at) there are two cases: (i) if em = 1 the matrix Cem(am) is simply struck out of G to form C> and there is no associated vertex for the collineation (5). (ii) if em > 1 we form C from C by replacing CeJ<xm) by C^^ctJ, and the vertex associated with this latter matrix is the point A -^ei+.-.+em-i+l- In case (i) the united points of (5) are just the united points of (3) which lie in ^+...+^^ = 0, and a point associated with any latent root a of C is also associated with the latent root a of C. In case (ii) the united points associated with the latent roots of C different from olx are the same as those associated with the corresponding latent roots of C. But the united points of (5) associated with ax form the (t— l)-space defined by the vertices Hence, recalling our original reason for investigating the united points of the collineation (5), we have Theorem II. 7/ the vertex Aei+_+eml of the simplex of reference is associated with the submatrix Cr6m(am) of C, the united lines of (3) which pass through Ae+mmm+em_t either join the point to another united point of (3) or, in the case in which em > 1, they lie in the tspace which joins the fundamental space S(_x of united points of (3), which passes through ^+...+^, to the vertex ^+...+^+1- If we consider the implications of this theorem for the case em> 1, we see that the £-space joining the fundamental S^ corresponding to the latent root ax to the vertex Aei+mmm+em +1 must be united. This is immediately verifiable from the equations of the collineation. Since in connection with this new result the vertex Aei+...+em-1+i plays no special part, we evidently obtain a more general theorem thus: Let 1 = e-± = ... = e8 < e8+i ^ e8+2 ^ • • • ^ et» Then the 8( which joins S(_1 to any point in the space defined by the vertices Aei+...+es+l> ^rK..+e,-i-i+l> • • • > -^+...+^^+1
350 VTII. COLLINEATIONS is united. This result should be compared with that in § 3, p. 341, which gave the coordinates of a point P in a united 8t through S(__t as xs-fe,+1+...+e<-!> 2*+€,+1+...+c<_!> 0, ..., 0). The two results are the same. An immediate deduction from the above is Theorem III. If a collineation is non-special the only united Jc-spaces are those determined by k 4- 1 united points. Finally, we remark that if a collineation has a non-linear elementary divisor, then there is a united line which contains only one united point. With the notation used above, if s < t the line is united, and contains the one united point Ae+mmm+e§. This result will be found useful on later occasions. Our discussion of the geometry of the united points and spaces of a non-singular collineation has made use of the indices ely..., e&, and of the equalities between the latent roots; the actual values of the latent roots do not enter into our results. The geometrical properties of the collineation are therefore sufficiently indicated if we give the numbers ev ...,e& (whose sum is n+l) and indicate which of them refer to equal latent roots. We do this by writing the Segre symbol for the collineation. This consists of the ev...,ek written inside square brackets in any order, subject only to the condition that indices which refer to equal latent roots are put together, and enclosed in round brackets. Thus in [3] there are the following types of non-singular collineation: [1,1,1,1] [(1,1),1,1] [(1,1),(1,1)] [(1,1,1),1] [(1,1,1,1)], [1,1,2] [(1,2),1] [(1,1),2] [(1,1,2)], [2,2], [(2,2)], [1,3] [(1,3)], Writing (x, y, 2, t) instead of (xQ, xv x2, x3), and denoting the vertices of the simplex of reference by X, Y, Z9 T, the canonical forms of these collineations are, respectively, as follows:
4. UNITED ^-SPACES OF NON-SINGULAR COLLINEATION 351 (i) x' = ax, y' = fly, z' = yz, t' = St. The vertices X, Y, Z, T are the only united points, their joins the only united lines, and the faces of the simplex are the only united planes. (ii) x' = ax, y' = ay, z' = yz, t' = St. The united points are the points of XY, and the points Z, T. A united line joins any two united points, and a united plane is a plane containing three independent united points. In particular, there is a pencil of united planes through ZT. (iii) x1 = ax, y' = ay, z' = yz, t' = yt. The united points lie on the lines XY and ZT, and the united lines and planes are obtained as in (ii). In particular, there are pencils of united planes through XY and ZT. (iv) x1 = ax, y' = ay, z' = az, t' = St. The united points fill the plane X YZ, and the point T is also united. United lines and planes are found as in (ii). In particular, any plane through T is united. (v) x' = ax, y' = ay, z' = az, t' = at. This is the identity transformation. (vi) x' = ocx, y' = fly, z' = yz +1, t' = yt. The united points are X, Y, Z\ the united lines are YZ, ZX, X Y and also ZT. The united planes are YZT, XZT,XYZ. (vii) x' = ax, y' = ay + z, z'= az, t' = St. The fundamental spaces are XY and T. United lines either join two united points, or join a point of ZX to Y. The planes of the star with vertex YT are all united, and the plane X YZ is also united. (viii) x1 = ax, yf = ay, zf = yz + t, t' = yt. The united points are Z and the points of XY; united lines either join two united points, or coincide with ZT. The united planes are the planes through ZT, and the plane X YZ. (ix) x'= ax, y' — ay, z' = az + t, t'' = at. United points fill the plane XYZ; the united lines either lie in this plane or pass through Z. Any plane through Z is united. (x) x1 = ax + y, y' = ay, z' = flz + t, t' = flt. The united points are X, Z; the united lines are XY, XZ, ZT, and the united planes are XZT, XYZ.
352 VIH. COLLINEATIONS (xi) x' = ax + y> y' = ay, z' = az + t, t' = at. The united points form the fundamental space XZ\ the united lines join the points (a,0,c,0) and (0,a,0,c). The united planes all pass through XZ. (xii) x' = ax> yf = /?y + z, z' = /?z +J, «' = fit. The united points are X, F; the united lines XY and FZ, and the united planes YZT and XFZ. (xiii) x' = a#, y' = ay + z, z' = <xz +t, t' = at. The united points form the fundamental space XY; the united lines lie in XYZ and pass through F. The united planes form a star having YZ as vertex. (xiv) x' = ocx + y, y'^ay + z, z'= az + t, t' = at. X is the only united point, XY the only united line, and XYZ the only united plane. 5. Cyclic collineations. As in V, § 2, Th. I, we see that the non- singular collineations in [n] form a group. We now ask: what elements, if any, of this group are of finite order? That is, we seek to determine the non-singular collineations y = Ax (1) which are such that for some integral value of m, y = Amx is the identity transformation. This is equivalent to the condition A™ = kl (&*0). When this condition is satisfied, (1) is said to be cyclic, of order m. The successive transforms Ply P2,... of any point P of [n] form a finite set containing at most m points, since Pm = P. When m = 2 the collineation (1) is often called an involutory transformation of the points of [n], and in the case n = 1 the pairs of points P, Pt are said to be in involution on the line. We show that if (1) is cyclic the elementary divisors of A must be linear. We saw at the end of § 4 that if A has a non-linear elementary divisor there exists a united line which contains only one united point. The collineation induced on a united line must also be cyclic, and its order must be a factor of m. If we choose the coordinate system so that the line is given by the equations #2 = 0,...,xn = 0, we can also assume that the induced collineation is in canonical
5. CYCLIC 0OLLINBATIONS 353 form. Since there is only one united point on the line the canonical form must be y0 = otx0 + xv yx = <xxl9 corresponding to a non-linear elementary divisor. If we apply this collineation m times to (x0, xx) we obtain the collineation y0 = ccmx0 + moLm-xxXy yx = ocmXv This is the identical transformation if and only if mam-i = o. Since the ground field K is without characteristic, and a=#0, this cannot happen. Therefore, if (1) is cyclic its elementary divisors are necessarily linear. We now write the collineation (1) in canonical form, which is y = /ax 0 0 . 0 a9 0 . k0 0 0 . since the elementary divisors are all linear. The condition that (1) be cyclic of order m is now easily seen to be af = k (i= l,...,n+l) for some non-zero k. This result may be stated as Theorem I. A necessary and sufficient condition that the collineation y = Ax be cyclic of order m is that the elementary divisors of A — A/n+1 be linear, and the latent roots of A satisfy the equation xm = ky where k is a non-zero element of the ground field. The condition can also be expressed in terms of invariant factors: Theorem II. A necessary and sufficient condition that the collineation A y = Ax be cyclic of order m is that the invariant factors of A— AJn+1 have no repeated factors and that each invariant factor be a factor of the polynomial Xm — k, where k is some non-zero element of the ground field. HP! 23
354 VIH. COLLINEATIONS This last theorem also holds when the ground field K is not algebraically closed. Indeed, let A be a non-singular matrix over any field K without characteristic, and let us suppose that the collineation (1) is cyclic of order m. Then Am = fc/n+1 {k in K). Now let K' be an algebraic extension of K over which the characteristic polynomial of A is completely reducible. Each invariant factor of A — AJn+1 is then completely reducible. The reasoning used in proving § 2, Th. Ill, shows that over K' the collineation can be reduced to the canonical form used in the case of algebraically closed fields. The reasoning given above then shows that the conditions given in Theorem II are necessary and sufficient for the equality ,m _ k being in K. 6. Some particular collineations. In this section we consider some particular collineations which are of geometrical interest, confining ourselves to the case in which the ground field K is algebraically closed. Some of the results we shall obtain are particular cases of theorems proved in § 3, in particular of Theorem V of that section. We first consider the non-special collineation with only two fundamental spaces 8a and Sb of united points. Since the collineation is non-special, we have [§ 3, Th. II] a + b+1 = n. The Segre symbol for the collineation is [(1,1,...,1), (1,1,..., 1)], with a+1 units in one bracket, b +1 in the other. The collineation can be written, in canonical form, as Vo = ccx09 Vi = ccxv • • . Vn = fan.
6. SOME PARTICULAR OOLLINBATIONS 355 where a=k/3. The line joining any point x which is not in Sa or Sb to its transform y meets 8a in the point z = y — fix, and 8b in the point t = y — ocx. If #* is any other point of [n] not in a united space, and y*, z*, t* are constructed as above, we see that the ranges (x, y, z, t) and (#*, y*, z*91*) are related. We can construct this collineation geometrically, as follows. We assume that we are given the fundamental spaces 8a and Sb, that these do not meet, and that we are also given one pair P, Q of corresponding points which are not united points. Let PQ meet 8a in R and Sb in 8. We find the point B corresponding to a given point A as follows. If A is in either united space, B = A. If this is not the case the join of A to 8a is a space of a+ 1 dimensions, £a+v say. Since a+1 = n — b, Za+1 meets Sb in a point, at least. If 27a+1 met Sb in a space Se (c > 0), then Sc, lying in 27a+1, would meet #a, which also lies in 2*a+1. Hence Sa would meet Sb> contrary to hypothesis. Hence £a+1 and Sb have only one point in common. Let this be D. Then AD, being a line in 2*a+1, meets Sa> in C> say. Hence A CD is a line through A which meets both Sa and Sb. It is the only line of this kind. For if AC'D' is another, C in Sa, and 2)' in Sb> it contains two points A and 0' of £a+v and therefore lies in Ea+1. Hence D' is common to Ta+1 and £6. Therefore D' = J9, and AGD = 4C'Z>\ On ^4 CD there is a unique point B such that (J[,B,C,2))A(P,g,^,5). This is the required point B. Since the collineation is non-special, the united ^-spaces of the correspondence are the spaces determined by k+ 1 united points. There are no others [§4, Th. III]. Hence they join an Sa> in Sa to an Sb> in Sb> where a' + ft' + l = & (a'^0,6'^0), or they he in $a or in Sb. In particular, any prime through Sb (Sa) joins 8b (Sa) to an /Sa_x (S^) in #a (/¾) and is therefore united. The spaces Sb and 8a are the vertices of the stars of united primes, as is evident from the dual transformation. A particular case is that in which ot+fi = 0. The collineation is then cyclic, of order 2. If P is transformed into Q, Q is transformed into P, and, with the notation used above, (P, Q9 R9 S) is a harmonic range. This is the involutory case. Another special case of interest is that in which a = 0, 6 = 71— 1. Our construction for the point corresponding to a given point 23-3
356 VIII. COLLINEATIONS becomes very simple. The collineation in this case is called an homology, and the special case of homology, when a+fi = 0, is called involutory homology. We now turn to certain collineations in which there is only one fundamental space of united points, confining ourselves to those whose Segre symbol is [(1,1,..., 1,2)]. We first consider the case in which n = 1. The equations are y0 = ccz0 + xv yx = ocxv This is the same collineation, from the projective point of view, as and since the elementary divisor of this last collineation is (A— 1)2 it follows that any collineation on a line with Segre symbol [2] can be written in the form _ Vo -~ ^o + ^u Vi = *i> by a suitable choice of coordinate system. Hence these collineations are all projectively equivalent. Conversely, if we are given a collineation on a line with one united point, Xy say, and a pair of corresponding points P, Q, a transformation of the above type is determined. We assume that X, P, Q are distinct. There exists a transformation of coordinates such that in the new coordinate system X = (1,0), P = (1,1) and Q = (2,1). Let the collineation be given in this coordinate system by the equations y0 = ax0 + /3xv yi = yx0 + 8xv Then, since (1,0) is a united point, y = 0. Since also it is the only united point, there is only one elementary divisor, of degree two. Therefore 8 = a. Finally, since (1,1) is transformed into (2,1), we have a + P = 25, and therefore a = /3 = 8. The collineation is therefore projectively equivalent to Vx = *i- Such a collineation is sometimes called a displacement on the line, its equation in non-homogeneous coordinates being y = 3+1.
6. SOME PARTICULAR OOLLINBATIONS 357 The collineation in [n] of type [(1,1,..., 1,2)] has the equations 2/o Pi n-2 n-1 Vn = = = = = ax0, axv axn. <xxn- axn. "V > -2> -l + ^n* J The united points form the fundamental space *n = 0. If the equations of any united ^¾ are n S toifii = 0 (i = 0, ...,n — fc — 1), the equations (1) show that n a2a^+-¾^¾ = 0 (i = 0, ...,n-fc-l) for every point (x0,..., xn) of 5¾. If xn = 0 for every point of Sk then Sk lies in the fundamental space of (1). If xn is not zero for every point of 8k we must have ain-l = ° (* = 0, ...,n-fc-l), if #fc is united. Hence Sk passes through the point 4 = (0,0,...,0,1,0). Conversely, the equations show that if Sk passes through this point it is united [cf. §4]. From the equations (1) it is evident that if P is transformed by the collineation into Q> Q = aP + xnA, so that Ay P, Q are collinear. If P is not a united point, so that Ay P, Q are distinct, there is a collineation induced on the line. This has only one united point, the point A. Hence it is the displacement having A as united point which transforms P into Q. The displacement is thereby defined. Now consider any Sn_2 which lies in the fundamental space, but does not pass through A. Any prime through Sn_2 is transformed into
358 VIH. COLLINEATIONS another prime through $n_a. In particular, the join of P and Sn„% is transformed into the join of Q and Sn_2. Let any line through A meet the first prime in P' and the second in Q'. The collineation (1) transforms P' into the point on A P' which lies in the second prime, that is, into Q'. Thus, having chosen an Sn__2 in the united space, and being given one pair of corresponding points P, Q, we can find a pair of corresponding points P', Q' on each line through A> and these determine the displacement on the line. The transformation (1) then transforms any point R into the point S into which R is transformed by the displacement on AR. 7. Singular collineations. We conclude this chapter with a brief account of the singular collineations in [n]. The reasoning of § 2 shows that, if the ground field K be algebraically closed, the coordinate system can be chosen so that the coUineation may be written in the canonical form y = /ceiK) o . o \ x, (l) \ 0 0 . Gek(ock)/ where at + 0 (i ^ h), and ^ = 0 (i > A), so that the submatrices corresponding to zero latent roots are together at the end. The discussion of united points given in § 3 is valid, subject only to the modification that the fundamental space S^^i corresponding to the latent root zero is now the space of points which have no transform; but the geometrical relation of this to the other fundamental spaces is as before. Again, it is clear from (1) that the points of [n] which have a transform are all transformed into points in the space Sn„k+h whose equations are ^+...+^+1-1 = 0, #Ci+mmm+eh+2~i = 0, ..., xn = 0. (2) In the case in which e*+i = ^+2= ••• = Cfc= 1> the spaces /¾^^ and Sn_k+h have no points in common, and their join is [»]. The description of the collineation is particularly simple
7. SINGULAR COLLINEATIONS 359 in this case. Just as in § 3 we see that there is a projective correspondence between the (k - A)-spaces through 8hmmh_1 and the points of Sn_k+h. The collineation (1) may then be described as follows: A point x in Sk_h_x has no transform. If x does not lie in Sk_h_x it is projected from Sk_h_1 into the point z = \x0> • • • > xn-k+h> 0, ..., 0) -k+h- In &n-k+h we consider the non-singular collineation y* = /OeMi) ° • o \ «*, 0 Cet(a2) . 0 • • • • where y* = (2/0> — !&»-*+*)> and z* = (*<» — >aW-JH-»)- The Point y = (y*, 0,. ..,0) of #„_&+& obtained from the transform of 2* is evidently the transform of # under the collineation (1). If the indices eh+v ..., ek are not all equal to unity, the construction of (1) by a series of non-singular collineations and projections is not quite so simple, because #£_&_! and Sn_k+h have points in common. We may, however, proceed thus: Let En_k+h be the space given by This has no points in common with the space Sk_h__v which is spanned by the vertices A A A ^¢1+...+W ^¢1+...+ca+i> • • • 9 -^ei+.-.+e*-!* We project any point x not in Sk_h_1 from Sk^.h_1 into a point x^) of ZnM. Then , * —> %) == (X0> • • • > a?Ci+...+CA-l> "> XCi+...+ca+1> • • • > ^Ci+.^+CA+i-lJ 0, xe1+...+^+1+l> •••)> the zeros corresponding to the first rows of the submatrices ¢^0),...,^(0) (3) in (1). Since we wish to consider a collineation in Zn_k+h, we omit the zero coordinates in X(1)9 and write #(1) = \x0> •••J^ei+.-.+eft-l'#«!+...+ca+1> •••)•
360 VIIL OOLLINEATIONS We now perform the non-singular collineation in 2?n_&.+A, X(2) - 0 Att+1-1 x\ U)> R ek-U where 76i,...,Ieh are each unity matrices, and De is the (cyclic) exe matrix, 2), = /0 1 0 . 0\ 0 0 1 0 . . ,1 0 . The effect of the matrices De is to displace the first to the last coordinate in each batch of coordinates in xfX) which corresponds to a submatrix (3), so that x(2) - \x0> •••>xei+...+eh-l>Xei+.»+eh+2> '">xe1+...+eh+i~l>Xe1+...+eh+l> "')' Now consider the (k — h— l)-space i^-^-i whose equations are Xn — X-t — = ¾ ei+...+e/r-l X, '«Ji+...+e/k+l Xt ^+...+^+1-2 X. '«Ji+...+ca+1+1 = ... = Xt ^+...+^^+2-2 o, -2 = 0, -2 = 0, as, xn-l , + X. = o, i-i = °> ei+...+e/i ^ ^ei+...+eA+ ^+...+^+1 +^ei+...+^+1-l = °> a:. = 0. This space evidently does not meet the Sn_k+h given by (2). If we project the point xf2) in Zn_k+h from 2^.,^ on to Sn_k+h (noting that £n-k+h does not meet Ek_h_^) we obtain the point r K s X(S) = (^0» • "9Xei+... +eh-l> X6i+...+eA+l> •-^^+...+eA+i-l> ^, ...)> + If z is the point in 2?*_A_i which is collinear with x$y in £„-*+* and <r(8) in Sn_te+h9 we see that z = #<8)-#$).
7. SINGULAR COLLINEATIONS 361 where in each batch of coordinates corresponding to a submatrix (3) the order is standard. Since'%) is in Sn_k+h, the coordinate corresponding to the last row of each submatrix is absent. Hence, if we omit the zero coordinates in #(3), and write x\ '(3) — (Xq9 •••>^c1+...+CA-l>a?e1+...+CA+l> •••>a:c1+...+eA+1-l> •••)> the non-singular collineation in Sn_k+h given by TyH^ V ^= 0 ¢,(0,) 0 0 0 Geh{ah) ^+1-1 X' ** transforms x*3f into the transform of x by the collineation (1). It will be noticed that in each of the various projections and collineations we have used above, a united point of (1) remains a united point. For the coordinates of a united point are of the form (x0, xly..., #ei+###+6^_i, 0,0,.. •, 0), and are therefore unaffected. Finally, the Segre symbol can be extended to denote singular collineations. This is usually done by drawing a line above the indices et in the symbol which correspond to a zero latent root. It should be noted that there is only one line, and this covers a single index, or a single set of indices in a round bracket. Thus the singular collineations in [3] are, apart from the improper one with zero matrix, y = 0, [1,1,1,1] [(1,1),1,1] [(1,1),1,1] [(1,1)(1,1)] [(1,1,1),1] [(1,1,1),!], [1,1,2] [1,1,2] [(M), 2] [(1,1), 2], [1,(1,2)] [1,(1,2)] [(1,1,2)], [2,2] [(2¾ [3,1] [3,1] [(M)], [*]•
* CHAPTER IX CORRELATIONS Ik this chapter we assume once more that the ground field K is commutative and without characteristic. We shall indicate in the various sections whether K is also assumed to be algebraically closed. 1. Correlations. Let Sm, Sn be two projective spaces of dimensions m and n respectively over K. In each of them choose an allowable coordinate system. Then, if A = (a{j) is an (m+ 1) x (n+ 1) matrix over K, the equations n *>i = 2i<t>ijZj (i = 0,...,m), that is, in matrix notation, v = Ax, (1) determine a prime v, of equation in Sm corresponding to any point x of Sn which does not satisfy the equations ^ = Q If A is of rank r, the exceptional points in Sn> that is, those with no transform under (1), form a space of n — r dimensions. The primes in Sm corresponding to non-exceptional points of Sn form a star of dimension r — 1, and hence all pass through a space of m — r dimensions in Sm. We denote these spaces of n — r dimensions in Sn and m — r dimensions in Sm by the symbols Sn_r, Sm_r respectively. If P and Q are any non-singular matrices, of orders1 n -f 1 and m + 1 respectively, over K, the equations x = Px*, y = Qy* determine allowable transformations of coordinates in Sn, Sm. The f An (n+1) x (n+1) matrix is sometimes said to be of order n+1.
1. CORRELATIONS 363 corresponding transformations of the prime coordinates in these spaces are given by the equations w* = P'u, and v* = Q'v, and (1) becomes ^ = qAp%% in the new coordinate systems. Since A is assumed to be of rank r, we can determine P, Q so that Q'AP=tIr 0\ [o or The relation (1) now becomes v* = z* (i = 0,...,r-l), v* = 0 (i > r) in the new coordinate systems. The space #„_, is given by x0 = ... = #r_i = 0, andS^by - „* - 0 By an argument similar to that used in VIII, § 1, we can show that the points of any (n — r + l)-space through $n_r all correspond to the same prime through Sm_r, and there is a one-to-one correspondence between these (n — r+ 1)-spaces through #„_, and the primes through Sm_r. Now choose two (r — l)-spaces, /Sr_x in Sn and S^ in Sm, which have no intersection with Sn_r and Sm_r respectively. Then there is clearly a correspondence between the points of S,.^ and the primes ((r — 2)-spaces) of S^, in which a point P of S,.^ corresponds to the intersection of S^ with that prime of Sm through Sm__r which corresponds to P and to the join of P to Sn_r. The properties of the correspondence (1) can evidently be obtained at once from the properties of the point-prime correspondence between Sr_t and S'r_t. Hence we need only consider that case of (1) in which r = ra+ 1 = n+ 1. We then have a non-singular correlative correspondence between two distinct spaces of equal dimension, Sn and Sm. If the allowable coordinate systems in the two spaces are suitably chosen, the correspondence is given by the equations v* = xi (t = 0,..., m = n).
364 IX. CORRELATIONS Linearly independent (dependent) points of Sn give rise to linearly independent (dependent) primes of 8m. Hence the points of 8n which lie in a linear space Sk are transformed into the primes of Sm which pass through an 8m_k_v If (..., j?io ^,...) are the Grassmann coordinates of 8k it is evident that the coordinates of £m_fc_i are (...,gv-H,...), where Indeed we can pass from 8n to 8m by the projective transformation yi = xi (i = 0,...,m), followed by the dual transformation in Sm itself. The enumeration of the properties of this correspondence is a simple matter, and need not detain us here. In this chapter we are more concerned with the case in which the two spaces Sm and Sn coincide in a single space, which we denote by [n]. We choose an allowable coordinate system in [n], and consider the equations n u>i = IiCiijXj (* = 0,..., n), (2) that is, u = Ax, in matrix notation. These equations relate a point x of [n], which does not lie in the space whose equations are Ax = 0, to a prime u whose equation is u0x0 + ...+unxn = 0. Such a relation is called a correlation in [n]. If x = Py is an allowable transformation in [n], the induced transformation of the prime coordinates is u = Pvy where P is the complement of P. Equations (2) then become v = P'u = P'Ax = P'APy.
1. CORRELATIONS 365 One of the main objects of this chapter is to obtain canonical forms for sets of equivalent matrices A>B>..., equivalent under the relationship __ w AT> where P is non-singular. This will enable us to write the equations of any given correlation in canonical form, so that we can examine its properties. But before doing this we examine certain preliminary notions concerning correlations which, besides being of interest in themselves, will be useful in our discussion of equivalent correlations. The correlation (2) sets up a correspondence between points x and y of [n] in which there corresponds to any point x all the points y of the prime u. The point y corresponds to # if and only if y'u = 0, that is, if and only if , A 9 J y Ax = 0. (3) The equation (3) is said to determine a bilinear correspondence between x and y, and the form y'Ax> which is linear in the inde- terminates x and y, is called a bilinear form. From (3) we can always obtain (2). Since we can also write (3) as x'A'y = 0, we see that if y corresponds to # in the bilinear correspondence arising from (2), then x corresponds to y in the dual correlation v = A'y. (4) Hence the points x whose associated primes under (2) contain a given point y lie in the prime v given by (4). The correlation (2) and the dual correlation (4) suggest another correspondence, between points which give the same prime in (2) and (4), A>y _ ^ (g) This correspondence can be interpreted geometrically as follows: if z is any point corresponding to x in the bilinear correspondence (3), then z'Ax = 0. Using (5), this becomes z'A'y = 0, or y'Az = 0, which shows that y corresponds to z in (3). In the case of a non- singular correlation, that is, when A is non-singular, (5) is the collineation r A //JX y = AAx; (6) we call this the associated collineation.
366 IX. CORRELATIONS If we transform the coordinates by the equations x = Px*, writing B = P'AP, (2), (3), (4) and (5) become, respectively, u* = 2to*, 2/*'J3z* = 0, v* = J3'y*> •and B'y* = Bx*. Hence the bilinear correspondence, dual correlation and, when defined, the associated collineation are invariantly related to the given correlation. The study of the correlation (2) will be greatly simplified if, at the same time, we consider the associated correspondences given by (3), (4) and (5). We begin by considering the special case in which the correlation (2) is identical with the dual correlation (4); that is A'x = pAx for all points x in [n], p being some non-zero element of K. Then A' = pA. Forming the transpose of these equations, we obtain A = pA'. Hence A = p2A. Therefore p = ± 1. When p = 1, so that A' = A, the matrix A is a symmetric matrix. The correlation (2) is then called a polarity. When p = — 1, so that A' = —A, and A is skew-symmetric, (2) is called a null-polarity. We shall find it convenient to consider the properties of polarities and null-polarities in some detail before considering the properties of correlations in general. We conclude this introductory section by remarking that we may consider, if we wish, a correspondence between two (possibly identical) spaces Sn and #m in which primes in Sn determine points in Sm by means of the equations n Vi = Hb^ (»« 0,...,fn), or y = Bu. (7)
1. CORRELATIONS 367 It is clear that the properties of such a prime-point transformation can be obtained from the properties of the point-prime transformation we are discussing, and it is not necessary to discuss prime-point transformations separately. We note, however, that when m = n and both A and B are non-singular, the point-prime transformation (1) determines the prime-point transformation x = A -*v from 8m to Sn, and the prime-point transformation (7) determines a point-prime transformation u = B-hf from Sm to Sn. 2. Polarities. Let u = Ax (1) be the equations of a polarity in an allowable coordinate system in [ri\. The bilinear correspondence which arises from (1) is y' Ax = x' Ay = 0. Any two points £, t\ which are specialisations of x, y such that y'AZ = g'Ar, = 0 are said to be conjugate points in the polarity (1); if a point is conjugate to itself, it is self-conjugate in the polarity. If £, y\ are conjugate points, the prime A£ corresponding to £ contains 1/, and the prime Ai\ corresponding to rj contains £. We begin our study of the polarity (1) by obtaining an allowable coordinate system in [n] in which the equations of the polarity can be written in a standard form. There are several methods by which we may proceed. The one we shall follow is one which will prove useful later in this chapter. Our method involves the bilinear form n n y'Ax = ^ YA^iiVi^i (2) in the two sets of indeterminates (#0,..., xn) and (y0,..., yn), and only uses transformations of coordinates of a very restricted type. We shall, for the purpose of the present reduction, say that an allowable transformation of coordinates in [n] given by is of restricted type if the elements of the matrix P above the principal diagonal are all zero. Using the rules given in Chapter II for
368 IX. CORRELATIONS constructing products and inverses of matrices, we can prove at once that the product of two transformations of restricted type is also a transformation of restricted type, and the inverse of a (non- singular) transformation of restricted type is also a transformation of restricted type. We now prove Theorem I. By a suitable transformation x =» Px*, y = Py* of restricted type, the bilinear form y'Ax can be reduced to the form +<>2rytxt+ --+cnytxh where (iQy...,in) is a derangement of (0,..., n), r is a suitable integer such that 0 ^ r < \{n + 1), and c2r,..., cn are in K. There is nothing to prove if n = 0, and we therefore proceed by induction on n, assuming that the theorem is true for bilinear forms in the indeterminates (x0,..., xm) and (yQ,..., ym)> where m < n. Case (i). If ain = ani = 0 for all values of i9 tt—l n—1 y'Ax = S S <W*j. By hypothesis there exists a transformation n—1 n—1 xi = S <lij£p Vi = S &^; (*' = °> ...,n-l) of restricted type which reduces the bilinear form 71— 1 71—1 i =0?-0 to the form required by the theorem. Then the transformation x = Px*, y = Py*, where P = I ¢00 • 0 «10 ¢11 • • • • 9V-271-2 0 ?n-10 • • ?n-ln-l 0, 0.. 0 l/
2. POLARITIES 369 is of restricted type, and reduces y'Ax to the required form, with in = n, and cn = 0. Case (ii). ann4=0. We may write y'Ax = = Since a where n s i=0 n / n \i=0 in = = an<> ain^j Wo anixi\ n-1 i-0 the transformations i X = p W Q-H, 0 l • 0 »1» 2/ = • • a n-1 j=0 0-Y 1 n,—In 0 1 0 0 annl are of restricted type, and reduce y1 Ax to n—l n—l CnVnL+ S S M«6» i=0i-0 where cn = a"n, and bit = «„,*(«„«.<%-«<«.«»>)• BY the hypothesis of induction we can perform a restricted transformation on (lo. •••» £n-l) and (%. •••>Vn-l) s0 that n—l n—l i=Oi»o is a form of the kind required by the theorem and then, just as in Case (i), we deduce a transformation of restricted type which reduces n—l n—l 2 S buViZj + CnVnZn to the required form. Since the product of two transformations of restricted type is also of restricted type, the form y'Ax is reduced to the required form by a transformation of restricted type.
370 IX. CORRELATIONS Case (iii). ann = 0, ain = ani + 0 for some values of i. Let k be the largest value of i for which ain 4= 0. Then (k \ / k \ n~l n"x i«0 / \i-0 / i-0i-0 n—1 n-1 = Vn£k + *nVk+ 2 2 <%2/i^> i-0 i-0 where _ * From this last equation, we have fc-i #* = ~ 2 <*>nkanixi + ank£k- i-0 Replacing x& by this expression in the indeterminates #0,..., xk_v £k, and making a similar substitution for yk, we find that n— 1 n—1 i-0 i-0 w—1 n—1 = yn£k + XnVk+ 2' <%2/fc^ + E' «iA:2/^A; ?=0 i-0 w—1 n—1 +***#***+ 2' 2'^*W i-0 i-0 = ( 2' &<& + yn) £fc + 7* ( E' Mi + *n ) + afcfcanf ZkVk \i-0 / \i-0 / + 2' S'^i^ (&<* = M> i^O ;-0 where the 6^, 6^ are elements of K easily determined by direct substitution. Now, the transformation k £& = E flni^i* i-0 n-1 fc £n = 2' ^1 + ^ +\akkank S «ni*i> i-0 i-0 £, = 3¾ (i = 0,. ..,fc-l,fc+1,..., n-1), n-1 + In the summations E' the value i = A; is omitted. i-0
2. POLARITIES 371 is of restricted type, and therefore its inverse is of restricted type. The form above becomes under the inverse transformation. We now apply our hypothesis n— 1 n— 1 of induction to 2' 2' ^nVi^i- This is reduced to the form required i =0:/=0 in the theorem by a transformation of restricted type on (b0> • • • > bfc-l> bfc+l> • • • > bn-l)* If the new indeterminates are (b0> • •' 9 bfc-l> bk+V • • • y bn-l)> and we add to the equations of transformation the two equations ik = bft> bn = Sn in their correct position, the new matrix is still of restricted type, and we have reduced y'Ax by a sequence of restricted transformations to the required form. Thus our theorem is proved. The use of transformations of restricted type is of purely auxiliary importance, and when we are concerned with only one symmetric bilinear form y'Ax we can obtain a simpler standard form by using allowable transformations which are not of the restricted type. We first carry out the transformation < = ££+£& *i\ — bi0 bi!> • • • x* = £* _ £* on the standard form obtained in Theorem I. It becomes n where and c2i = 2 (» = 0,...,r-l), Cai+i = -2 (» = 0,...,r-l). 24-4
372 IX. CORRELATIONS Remembering that (iQ,...,in) is a permutation of (0,..., n), we finally use a transformation which merely amounts to a permutation of the indeterminates and we have y'Ax^ZdJtXt, where dQ,..., d8 are not zero. If the resultant of all the transformations used is the transformation then x=RX, y = RY, R'AR = i0 0 0 • 0 di • • ds • 0 . 0' 0 . 0 0 0 and therefore the rank of A is s +1. Hence we have Theorem II. (a) By a suitable choice of the allowable coordinate system any symmetric bilinear form y'Ax can be reduced to 8 where s +1 is the rank of A. (b) By a suitable choice of an allowable coordinate system the polarity u = Ax can be reduced to the equations U^d^ (i = 0,...,^), where dQ,...,d8 are not zero, but ds+1 = ... = dn = 0, the rank of A being s+l. When the ground field K is algebraically closed we may use a further transformation ZJ-djZ, (i = 0,...,5), X? = X, (i = 8+1,. ..,n). The effect of this is to replace each di (i ^ s) in the above Theorem by 1. Hence we have Theorem III. If the ground field K is algebraically closed, the only projective invariant of a polarity is its rank. By a suitable choice of the allowable coordinate system the equations of a polarity of rank 5+1 may be taken as ui^xi (i = 0,...,5), 1^ = 0 (i = 5+1,..., n).
2. POLARITIES 373 The system of allowable coordinates which reduces a polarity to the form given in Theorem III is not uniquely determined, as the following discussion will show. We assume that the matrix A in (1) is of rank s + 1, and that O 0. Let be any point which is not self-oonjugate with respect to the polarity (1). Then a-W^a^ + 0. (3) The equation ^Ax = 0 (4) determines a prime. This prime contains the #n_8_i given by Ax = 0, coinciding with it if s = 0. By virtue of (3) the point P®> does not lie in the prime (4). We may therefore choose n independent points in this prime, say, P<*> =(4,...,4) = ^) (i=l,...,n), which, together with P®\ form a simplex. We take this simplex as a simplex of reference for a new coordinate system, noting, first of all, that dWAiP = 0 (i = 1, ...9n). Let the equation of the bilinear correspondence determined by (1) in the hew coordinate system be y*'Bx* = 0 (B = (bti)). Then if x*(i) is the transform of x®, xWBx*® = 0 (i = 1,..., n). But xW=(8i0,...,8in) (i = 0,..., n). Hence boi = 0 (i = 1,...,»). Since x*«»'J3a;*«»4=0, it follows that 600 =|= 0. n n Hence y*'Bx* = b00y0x0+^l S 6^?*?, and the matrix • • •
374 IX. CORRELATIONS must be of rank 8. Repeating the process just described for the bilinear form n n we arrive at the following result: Theorem IV. If a polarity is of rank s +1 we can construct a simplex of reference, referred to which the polarity has its equation in canonical form, as follows: (i) there exists a point P0) which is not self-conjugate with regard to the polarity; (ii) there exists a point P® which is not self-conjugate with regard to the polarity and is conjugate to P0); • *.*•••• (i +1) there exists a point Pl) which is not self-conjugate with regard to the polarity, and is conjugate to P°\ ..., P^-D; • **•*•• {s+l) there exists a point P8) which is not self-conjugate with regard to the polarity, and is conjugate to P(0),..., P8-V; (s + 2) now let i*8+1),..., Pn) ben — s independent points of the space Sn-8-i whose points have no corresponding prime in the polarity. Then we can take P°\ ..., Pn) as simplex of reference, and, choosing the multipliers suitably, the equations of the polarity become u^Xt {i = 0,...,5), ^ = 0 (i = s+l,...,n). The properties of a polarity are easily read off from the canonical form found above. In particular we observe that the properties of the polarity we have been considering, ui = xi (i = 0,...,5), \ ^. = 0 {i = s+l,...,n),f K ' can be deduced immediately from those of the non-singular polarity Ui = x{ (i = 0,...,5) (6) in the space S8 whose equations are xs+l = ••• = xn == 0. The space Sn_8_v whose equations are now x0 = ... = x8 = U, is called the vertex of the polarity. No point in this space has a corresponding prime in the polarity (5). Since the prime corresponding to any point not in Sn^8^,1 passes through Sn_s_ly it follows
2. POLARITIES 375 that a point of Sn_8_x is conjugate to any point of [n] not in Sn_8„v It also follows immediately, from the equations, that two points P and Q of [n] which are not in #„_,_! are conjugate with regard to (5) if and only if their projections from Sn_8_x on to S8 are conjugate with regard to (6). The geometrical meaning of polarities will become clearer when we discuss conjugate points in the chapter on Quadrics in vol. n. We need only remark here that points which are conjugate with regard to the polarity _ . are also conjugate with regard to the quadrio x'Ax = 0, and conversely, the relation between the points being the same in each case. The non-singular polarity (6) transforms the points of an Sk whose Grassmann coordinates are (...,2^...^,...) into the primes through an S8^k^x whose dual coordinates are (...,gv-u,...), where &"'* = Pu...iv This is evident from the equations (6). Conversely, the transform of the points of Sg^^ is the set of primes through 8k. The spaces Sk and S8_k_t are said to be conjugate in the polarity. Every point in the one space is conjugate to every point in the other space. We now prove Theorem V. If Sk and #s_fc_i are conjugate in a non-singular polarity, a necessary and sufficient condition that they should meet is that 8k should contain a point P which is conjugate with regard to every point of Sk. Let lP=tol-,yi) (» = 0, ...,*) be k+1 independent points of Sk. The equations of ^-^-1 are given by the equations 8 2yfo = o (i = 0,...,¾). If Ss_k^ contains a point of Sk> say the point the equations in A0,..., Afc, a-0i-0
376 IX. CORRELATIONS must have a solution. But the condition that there exists a proper set A0,..., Afc such that the equations above are satisfied is precisely the condition that the point x SA,0» is conjugate to each of the points «/<0), ...,t/fc). The point x is then conjugate to every point of Sk. This proves the theorem, which is evident geometrically. If we follow the point-prime polarity u = Ax by the prime-point polarity y = Buf the result is the collineation y = BAx. We conclude our discussion of polarities by proving, in the case when the ground field K is algebraically closed, Theorem VI. Any collineation in [n] can be expressed as the result of a point-prime polarity followed by a non-singular prime-point polarity. We choose our coordinate system so that the collineation is given in canonical form y = Cx, where 0= /0^((¾) 0 . 0 o cjog . o 0 and Ge(a) is the exe matrix 0 CM)J OJLa) = la, /o 10 \o 1 0 a 1 . . • • • °\ ° a 11 . 0 a/
2. POLARITIES Now let £ = /Bei 0 0 0 where B, is the exe matrix 0 0 0 0 0 1 1 0 0 1 0 0 1\ 01 Of o/ Then A = BG= /B&fa) 0 0 £eaCe,(a2) But BeGe(a) is the exe matrix fo 0 0 [0C 0 0 a 1 0 a 0 0 a 1 0 0 0 0 BekCek(*kh which is symmetric. Hence A is symmetric. Now B is not symmetric, but also non-singular, and, indeed, B = B-1. The collineation y = Cx is the result of the polarity u = Ax followed by the non-singular prime-point polarity y = Bu.
378 IX. CORRELATIONS 3. Null-polarities. The properties of null-polarities may be discussed along lines similar to those followed in § 2 in the discussion of polarities. We first consider the reduction, by means of a suitable choice of allowable coordinate system, of the equations of a null- polarity to a standard form. We prove the analogue of § 2, Th. I, using the same restricted type of transformation. Theorem I. By means of a transformation x = Px*, y = Py* of restricted type, a skew-symmetric bilinear form y'Ax can be reduced to the form (*/<*** - vZK)+(y%xt - vtxt)+• • •+(yt-,xt-i - »t-i*t-.)> where (i0>--->^2r) are %r distinct numbers chosen from the set (0,..., n). y'Ax = a01(y0x1-y1x0). When a0l = 0 there is nothing to prove. If a01#= 0, we take P=/l 0\ and «01(2/0*1 - yixo) = 2/o xi - Vix*• This proves the theorem when n = 1, and we now proceed by induction. There are only two cases to consider in this theorem, since ann = 0, the matrix A being skew-symmetric. Case (i) If ain = 0 for all i, the theorem follows immediately from the hypothesis of induction (cf. § 2, Th. I). Case (ii). If not all ain are zero, let k < n be the largest value of i such that ain 4=0. Then (k \ / k \ n-1 w-1 S 0>nixi + 2 G*n2/* )xn + S S <fy*/i^- Now let k £* = ** (* = 0,...,*-l,*+l,...,n).
3. NULL-POLARITIES 379 This is a restricted transformation. Its inverse is fc-i 0 *i = ii {i = 0,...,fc-l,fc+l,...,tt)> which is also of restricted type. Since we can write (k \ I k \ n-l n-l i«0 / \i=0 / <«0 i«0 n-l n-l we have, after transformation, n—1 n-l \ i=0 / \ i-0 / i-0 i=0 where the bit bti are easily calculated elements of K. We now use the restricted transformation St = it (i = 0,...,n-l), n-l and 2/\4z = ,,*£*-,,*£* + 2' S'M?#. ' The rest of the argument follows from the hypothesis of induction, and is exactly as in § 2, Th. I. The theorem is then proved. Since the rank of PAP is equal to the rank of A, and the rank of the bilinear form described in the theorem is 2r, it follows that the rank of A is 2r. + In the summations 2', the value i = Jc is omitted. i
380 IX. CORRELATIONS By a further transformation, not of restricted type, which merely rearranges the indeterminates, the bilinear form y'Ax can finally be reduced to the form r-l S (52i-^2i+l""^2i+1^2i)* When we state this result in terms of null-polarities we have Theorem II. A null-polarity has only one projective invariant, its rank, which must be even. If the rank is 2r the coordinate system can be chosen so that the equations of the null-polarity are U2i == X2i+1> U2i+1 = ~~X2i> ^ = 0 0>2r). It is easy to show that the reduction to this form is not unique, but we do not stop to prove this explicitly. An essential difference between the null-polarity and the polarity is that every point x is conjugate to itself in a null-polarity. For, if A is skew-symmetric, x'Ax = (x'Ax)' = x'A'x = —x'Ax, and therefore x'Ax = 0. In other words the prime u = Ax always contains the point x. If two points x, y are conjugate, so that y'Ax = — x'Ay = 0, then any two points Xx+fiy and px + cry on their join are also conjugate. For (Ax' +/iy')A(px+ay) = Apx'Ax + Acrx'Ay+fipy'Ax+iicry'Ay = 0. Hence every line in [n], joining a pair of points which are conjugate in the null-polarity, contains an infinite number of pairs of conjugate points. We also have n n 0 = y'Ax = 2 2 aaViXj = 2 a^a;,-^) -SaoPw. (2) where (..., Py, •••) are the Grassmann coordinates of the line joining } (» = 0,...,r-l), (1)
3. NULL-POLARITIES 381 the two conjugate points x, y. Conversely, ifx and y are two points of a line whose coordinates (...,##,...) satisfy a relation 2^, = 0, i<j then we may write this as where bij = -bji (i>j), and bu = 0. Hence a null-polarity is associated with a system of lines (called a linear complex) whose Grassmann coordinates satisfy the relation (2), and conversely any linear complex is associated with a unique null-polarity. The geometry of a null-polarity is, essentially, the geometry of a linear complex and, as such, belongs to Line Geometry. A null-polarity can only be non-singular when n is odd, and then r = \(n +1). Let us briefly consider this case. We take the equations in canonical form. U2i — X2t+1> (» = 0,...,r-l). U2i+1 — ~~~X2i>) if j^ = (y&...,!£) (i = 0,...,&), are k +1 independent points of [n] which determine an 8ki the equations of the conjugate #„_£_! common to the transforms of all the points of Sk are S (yh+iZ2j-yijx2j+i) = 0 (i = 0,..., k). This Sn_k_x meets Sk if these equations are satisfied by 0 that is, if the equations S sWiA-^+i)A» = 0 (1 = 0,...,¾) (3) have a proper solution (A0,..., Afc). The (&+ 1) x (k+1) matrix (aa) = ( 2 (yv+i!&-!&y&+i))
382 IX. CORRELATIONS is skew-symmetric, and is therefore singular if k 4- 1 is odd. Thus there is always a proper solution of (3), and therefore 8k meets #n-ft-i> when k is even. On the other hand, if k is odd the matrix (aih) may be non-singular, and then ^¾ does not meet Sn_k_v This may be seen by taking that is, by considering an Sk determined by the first k +1 vertices of the simplex of reference. In fact the equations of Sn_k_x are then x1 = Xq = 0, 0¾ = x2 = 0, ..., 0¾. = %k-l =s 0> and this does not meet the Sk we have selected. An 8k is called self-conjugate if it coincides with its conjugate Sn-k-v This can only occur iffc = n — fe—1. We then have k = r — 1. If this condition is satisfied there are ^-spaces which are self- conjugate. For instance, the fc-space determined by the first, third, fifth, ..., nth vertices of the simplex of reference is self-conjugate. In fact, the equations of the conjugate Sk are Let us find the condition that the space Sk whose coordinates are (...,Pi0.„ik,...) is self-conjugate. Since the method does not use the canonical form of the null-polarity, we assume that the null-polarity is given by the equations u = Ax. (4) Clearly, a necessary and sufficient condition that 8k be self-conjugate is that any two points in Sk should be conjugate. Hence, any fine in 8k must belong to the linear complex 2 auPij = 0. Now any line in Sk is the intersection of Sk and an Sn_k+1. Let (...,<fo — l'*-2,...) be the dual coordinates of an Sn_k+V If this $n_fc+1 meets Sk in a space of more than one dimension we have U,...,tt-i for all values of i, j, but if the intersection is a line, its coordinates are (...,rij9...), where [VII, §5]
3. NULL-POLARITIES 383 It follows that a necessary and sufficient condition that 8k be self-conjugate is that i<i to i*-i for all choices of elements <jf*§ •••**-« which are coordinates of an Sn-k+v But we can always find an Sn_k+1 for which every Grass- mann coordinate except an assigned one is zero. We can therefore write the necessary and sufficient conditions that Sk be self-conjugate as 2jaijPiji0...ik-% — u> i<i for all choices of i0>...,tfc_2. A point-prime null-polarity followed by a prime-point null-polarity t/ = ifa leads to the collineation y = iLlz. If n is even, both A and 2? are singular, and the collineation is also singular. Hence we cannot expect to represent a non-singular collineation in this way. We may ask whether the representation is possible when n is odd. Similarly, we may enquire whether any collineation can be represented as the resultant of a point-prime null-polarity (polarity) followed by a prime-point polarity (null- polarity). In each case it is seen that certain conditions must be satisfied by the elementary divisors of the collineation, and that these are necessary and sufficient for the desired representation. Since the method of investigating the problems is the same in each case, we only consider one of them in detail, merely stating the results in the other cases. Theorem III. If n is odd and the ground field K is algebraically closed, a necessary and sufficient condition that a collineation in [»], y = Cx, can be represented as the result of a point-prime null-polarity u = Ax followed by a non-singular prime-point null-polarity y = Bu
384 IX. CORRELATIONS is that the elementary divisors of C — A/n+1 can be arranged in equal pairs: (A-0¾)% (A-(¾)^ (A-a2)c2, (A-a2)ca; .... We take C in the canonical form 0= /C^K) 0 . 0 0 Cet(oc2) . 0 . ... 0 0 . <yafc), and find the conditions that there exist skew-symmetric matrices A, B such that „ D A ,s. 9 C = BA. (5) An elementary argument shows that a necessary condition on C is that there is another elementary divisor (\ — ot)f corresponding to any elementary divisor (A —a)e of C — A/n+1. Let P = B~x. Then the elementary divisors of C — AIn+1 are also the elementary divisors of PC-\PIn+1 = A-\P, which is a skew-symmetric matrix. Hence, if A = a is a latent root of C> the rank of A — aP, and hence the rank of C — aIn+v is n+1 — 2p, where p is an integer [Th. II]. There must therefore be 2p elementary divisors of the form (A--^, ...,(A-a)c2p. This does not prove, however, that the numbers ev ...,e2p can be arranged in equal pairs, as is required by our theorem, and a more detailed argument is required to prove this. We divide A and P = J5_1 into blocks of a size suggested by the canonical form of C. We write and where Aij and P# are ei x e^ matrices. Since A and P are skew- symmetric, (AVy^-iAH), and (P«)' =-(**)• (6)
3. NULL-POLARITIES 385 Since PG = A, multiplication of the appropriate matrices shows that we must have ^^ ; , AU Using (6), we find that iA**y = (P%(o,))' = (%,(*,) (p«y = -c^p*. But (4«)' = ~ Aji = - i*Q»(«*) for all values of i, j. Hence (^.(0,)^ = ^(0,). (7) We use this result to determine the form of the submatrices P11. Write -°"=/*ll • Vle^ \Pe,l • *W and define Poh = ° (* = 1,...,^), ^o = ° (^= 1>--.,4)- Equating elements in the hth row and &th column on both sides of (7) we have <*jPhk + Ph-lk = *<Phk + Phk-l (*= l,...,c/, fc= 1,...,¾). (8) Suppose, first of all, that 0¾ 4=^-. Then, if A = k — 1 in (8), we get (^-^)^11 = 0, and therefore pn = 0. Next, if h = 1, & = 2, (8) becomes (a*-aj)Pi2 + 2>u = 0, and so ^12 = 0. Proceeding along the first row of Pji, then along the second row, and so on, we find that Pi* = 0, if 0Lt 4= 0¾. On the other hand, if aa- = aiy (8) becomes Ph-ik = Phk-i (h = l,...,e/, &= 1,...,4). 1! P I 25
386 IX. CORRELATIONS If ei < ejf all elements on lines parallel to the line joining plei to pHX are equal, and if ei > e^ all elements on lines parallel to the line joining p^ top^ are equal. Since also, if ei ^ ejy and if e€ ^eiy - ,, , 7 ^ . we obtain two forms for the submatrix PH: pn = (9) e,>e,; pn = • • • • • • • . 0 0 a 0 a b a b c (9) [0 . 0 a b c Thus the matrix P is divided into diagonal blocks of square matrices P= /P1 0 . 0 P2 . • • • 0 0 . P*J corresponding to sets of equal latent roots (av ..., akl) (afcl+1,..., ak%)... (a^1+1>..., a^), where kr = fc. It is easily seen that the matrix A is similarly divided into diagonal blocks since PC = A
3. NULL-POLARITIES 387 Writing A = /A1 0 . 0 48 . • • • • 0 0 . Ar> we may write the relation between the three matrices A, P, C as p7<V1+IK-l+i) o . o \-a*. (10) • • • • 0 0 . Ceki(ccki)j In order that P be non-singular, each submatrix P1 must be non- singular. We therefore consider whether there exists a non-singular matrix Pi made up of matrices of the form (9), such that both Pi and A1 are skew-symmetric. The latent roots which occur in the set (ockil+v . ..,¾) are all equal, and our concern is with the suffixes {ekix+1, ...,6^). To simplify the notation we omit the indices and sub-suffixes in (10), and write the relation as = A, where 0^ = ^2=...=^, and we assume that c1^e2^ ••• ^cfc- We have P = /P11 . Plk\ • • [Pki . where the elements P*1 are e^ x et matrices given by (9). Since P is to be skew, it follows from (9) that pn = 0 (»'=1,...,^)- Again, as in (6), P« + (P*)' = 0. If e4 = et, we see from (9) that pa = (pay, (H) and therefore pv = - P**. «5-3
388 IX. CORRELATIONS i We now form a new matrix Q = (g«) from P, by picking out the element q{j at the bottom left-hand corner of Pl\ Then Q is a k x k matrix, and from (9) we see that ^ = 0 (if e^e,), but, if ei = ej, &* = -?#> from (11). Now, if e± = ... = eai<eax+i — ••• = ^+08^ ...= 6ai+...+a, =5 eJfe» so that the indices el9...,eki which are assumed to be arranged in ascending order of magnitude, fall into s sets of equal indices, ¢= /Qn o . 0 / . Q22 . 0 ■ • • • • \ . • . Q88 where Qu is an at x a{ matrix, and all the submatrices above the principal diagonal are zero matrices. Since detg= ndetg**, l Q is singular unless all the matrices Qu are non-singular. Since Qu is a skew-symmetric matrix, it is only non-singular when at is even. We now show that Q cannot be singular unless P is singular. For if Q is singular, its columns are linearly dependent. But if we replace the elements of Q in their original positions in P, we see that the remaining elements of the column of P which contains q{j are all zero. Hence if the columns of Q are linearly dependent the columns of P which contain elements qtj are also linearly dependent. Therefore if Q is singular, P is singular. Since we assumed P to be non- singular it follows that the at are even, and we have proved the necessity of the conditions of the theorem, that the elementary divisors of C — A/n+1 can be arranged in equal pairs: (A-a^i, (A-a^i; (A-a2)c», (A-a2)««; ....
3. NULL-POLARITIES 380 The sufficiency of the conditions is proved more easily. Assuming that k = 2h, and that e1 = e2y e% = c4, ..., e2ft-i = e2h> we write, in equation (10), 1* = 0 Eei -Kx o 0 0 0 0 • • 0 0 x i matrix Ei = , \ 0 0 0 -Ee • • • (0 0 0 . a o 0 0 , 4, - 0 • • • . 0 . 1 • • 1 0 0 0 2h It is clear that Pi is non-singular and skew-symmetric, and by elementary calculation we see that the matrix Al defined by (10) is also skew-symmetric. We have therefore constructed the matrices A and B = P_1 required, and have proved our theorem. Other results, involving a combination of a null-polarity and a polarity, are most easily stated in terms of the properties of matrices. Theorem IV. Given a matrix C, necessary and sufficient conditions that there exist matrices Ay B (where A is skew-symmetric, and B is non-singular and symmetric), such that C = BA are that the elementary divisors of C — A/n+1 corresponding to non-zero latent roots can be arranged in pairs (A — a)e, (A + a)e, and the elementary divisors A.f corresponding to a zero latent root are even in number for any given even value off. Theorem V. Given a matrix (7, necessary and sufficient conditions that there exist matrices A, B (where A is symmetric and B is non- singular and skew-symmetric)y such that C = BA
390 IX. CORRELATIONS are that the elementary divisors of C—AJn+1 corresponding to a nonzero latent root of G can be arranged in pairs (A—a)e, (A+a)e, and the elementary divisors Xf corresponding to a zero latent root are even in number for any given odd value off. 4. Simple correlations. We now consider correlations in [n] which are neither polarities nor null-polarities. We assume that the ground field K is algebraically closed, and investigate the conditions in which the correlation _ * n. can be transformed to „ v = By by a non-singular transformation x = Py, or equivalently u = Pv> of point or prime-coordinates respectively. We may regard the transformation as a collineation in [n] instead of a transformation of coordinates, if we wish to do so. Algebraically, this is the problem of determining when a non- singular matrix P can be found such that B = P'AP. (2) If (2) holds, B, = p>A,p and therefore, if A is an indeterminate, B-XB' = P\A-\A')P9 so that the matrices A — XAf and J3 —AJ3' have the same rank and elementary divisors. This gives a necessary condition for the existence of a matrix P satisfying (2). We now show that it is also a sufficient condition when the matrices A—XA' and B — \Bf are non-singular. We can choose an element d (not ± 1) in K such that C = A-dA' and D = B-dB' are non-singular. If A—XA' and 5 — AJ5' (which are both of rank n+1) have the same elementary divisors, so have C-fiC' and D-/iD't /i being an indeterminate over K related to A by the equation (l+d/i)\ = d+/i.
4. SIMPLE CORRELATIONS 391 If there exists a non-singular matrix P such that D = P'GP, then 5-(1- d2)-i [D+dD'] = (l-d*)-1P'[C+dG']P - P'AP. Now let us suppose that C—fiC' and D—pD' have the same elementary divisors. Sinoe C and D' are non-singular, there exist [II, § 9, Th. V] non-singular matrices R and S such that D = RCS, and D' = RC'8. From the second equation we have D = S'CR', and therefore RGB = S'CR'; that is, XG = OX', where X = #i?. From XC = GX' we derive X2C = XCX' = C(X')2, and X*C = XC{X'Y = C(X'f, and finally XnC = C7(Z')re. Hence, if /(A) is any polynomial in -K"[A], /(Z) (7 = Cf(X'). But [II, § 10, Th. II] there exists a polynomial/(A) such that if 7»/(Z), and therefore YC = CY't then 72 = X Now, D = iS'Cir = S'CX'S = S'C(Y')*S ■ S'GY'Y'S = (S'Y)C(Y'S). Defining P = F'/S, we have proved that D = P'GP.
392 IX. CORRELATIONS Hence we have Theorem I. If A and B are two (n+ 1) x (h+1) matrices which wre such that A — XA' and B — XB' are non-singular, A being an indeterminate, then a necessary and sufficient condition that there should exist a non-singular matrix P such that B = P'AP is that A — XA' and B — KB' have the same elementary divisors. In the case in which A (and therefore J5) is non-singular, we deduce as an immediate corollary Theorem II. A necessary and sufficient condition that the two non-singular correlations u = Ax, u = Bx should be equivalent is that the associated collineations y = AAx, y = J5Bx should be similar collineations. The next question which suggests itself is: can we find an (n +1) x (n + 1) matrix A such that A—XA' has assigned elementary divisors? It is easily seen that this is not always possible. Let the elementary divisors of A — XA' be (X-cc1)%...,(X-akYK If a^ + O, and (A — atf* is a factor of each determinant of A — XA' of order j, then (A —a^"1)^ is a factor of each determinant of A' — XA of order j, and the converse of this statement is also true. Now, the transpose of A — XA' is A' — XA, and the transpose of a A-matrix has the same elementary divisors as the original matrix. Hence it follows that if (A — atf* is an elementary divisor of A — XA' (a( 4= 0), so is (X — oti1)6*. Again, if ax = a2 = ... = ar = 0, so that X\ ..., Ac' are the elementary divisors of A — XA' corresponding to a zero latent root, we deduce from the fact that det (A - XA') = det (A' - XA) that the characteristic polynomial of A — XA' is of degree r n+l-S^. It therefore follows that l k r 2 ef = ^+1-2¾. l l
4. SIMPLE CORRELATIONS 393 We have thus found some necessary conditions to be satisfied by the elementary divisors of A — XA\ If a€+ ± 1, the elementary divisors (other than Ac<) can be arranged in pairs, a typical pair being (A — 0¾)% (A — 0¾-1)% but when at = ± 1 the elementary divisor (A — OLifi may, possibly, be paired with itself. We now prove Theorem III. Sufficient conditions that there exists a matrix A such that A — \A' has assigned elementary divisors are (i) the elementary divisors other than those of type Xe can be arranged in pairs (A-a^, (A-a^1)^ (a^4= ± 1); (ii) if the elementary divisors are AV..,A% (A-a^A, (A-afi)A, ..., (A-a8)A, (A-O'*, then r l ^ + 2^ = fl l Let Df(a) be the 2/x 2/matrix Df(a)= /0 | 0 1 ° v \l< ThenDf(a)-AD'f{a) = /0 . . I • ■ 1 ° 1 ° I \ * \l-Aa - A , ] » • i • t • . 1 ) . 0 • • • 0 l-Aa -A • • 0 0 0 a 1 0 0 0 . 0 0 • • • a—A 0 0 • 0 -+1. • • 1 • • • • • • • 1 0 • • • . a\ a l\ 1 o\ ..01 . . 0 1 . . 0 1 . . 0/ • • . a —A 1 . • • a—A 1 0 • • 0 0 • 0 i
394 IX. CORRELATIONS Except for numerical factors, the determinant of this matrix is (A-a)'(Aa-l)', whilst the determinant of the (2/-1) x (2/-1) submatrix obtained by crossing out the first column and (/+ l)th row is (A-a)* A'-1, and the determinant of the (2/-1) x (2/-1) submatrix obtained by crossing out the first row and the (/+ l)th column is (Aa-1)A Hence the highest common factor of all determinants of submatrices of order 2/- 1 is 1, since we have assumed that a 4= + 1, and when a 4= 0 the elementary divisors of Df(<x) — XD'f(a) are therefore (A-a)' and (A-a-1)'.' If a = 0 the above argument shows that the only elementary divisor is A'. It follows that A = D. (0) DfM (3) Df.(cCsV is a matrix such that A —\A' has the assigned elementary divisors. Whenever the conditions of Theorem III hold we shall take u = Ax (where A is given by (3) above) as the canonical form of a correlation having the assigned elementary divisors. For A— \A' is non-singular, and we can apply Theorem I. The conditions of Theorem III have also been shown to be necessary in the case when A — XAf is non-singular and has no elementary divisors of the form (A + l)c. In this case A + A' and A -A1 are both non-singular. We call this the simple case, and we have Theorem IV. In the case of a simple correlation the conditions of the previous theorem are both necessary and sufficient. + If a = ± 1, we find that the elementary divisors axe (A± l)f, (A± 1)', but we must use a different argument.
4, SIMPLE CORRELATIONS 395 We observe that if n is even, A—A\ being skew, is necessarily singular, and the simple case does not arise. We could extend the notion of a simple correlation to include the case in which A — \Af has elementary divisors (A + l)e, provided that these can be arranged in pairs, but it turns out to be more convenient to proceed with the transformation of the most general type of correlation in [n] to canonical form. 5. Transformation of the general correlation. In this section we show how to obtain a canonical form for any correlation in [n] which does not fall into the categories already considered. Such &> correlation, which is not necessarily a polarity, a null-polarity, or a simple correlation will be called a general correlation. We treat the problem of transforming (or reducing) u = Ax to a canonical form as a purely algebraic one, working with the bilinear form y'Ax associated with the correlation. Let B and C be, respectively, the symmetric and skew-symmetric parts of A: that is, let B = HA+A')> C = \(A-A'). When A is transformed to P'AP, B is transformed to \(P'AP + P'A'P) = P'BP, and C is transformed to \(PAP-P'A'P) = P'CP. Hence our problem may be regarded as that of the determination of canonical forms for the simultaneous reduction of the symmetric and skew-symmetric forms y'Bx> y'Cx. If B and G are both non-singular, we are in the simple case considered in § 4, and we may use the canonical form given there for A = B + C. When, in the notation of § 4, A= De(a), y'Bx = i(cL+l)[(y0x2e_1 + y2e-1x0) + ... + (ye-iXe + yexe„1)] + £[(2/1^-1 + y*-l*l) + • • • + (ye-l*e+l + ye+l*«-l)]f and y'Cx = J(a-l)[(y0*8e-l-yto-i«b) + ... + (y«-l*e-ye*e-l)] +tf(yi**-i - y*-i*i)+• • • + (&-!*«+! - ye+i*«-i)]-
396 IX. CORRELATIONS We shall denote these expressions by y?(a, e) and y(a, e) respectively. When A is in the canonical form (Dei{ax) 0 . 0 0 Zy<x2) . 0 • • • • 0 0 . Dek(ak)j the form y'Bx is equal to a sum of forms such as /?(a, c), differing only in the a, e and in the indeterminates (y0,..., 026-i)> (#o> • • •» #2e-i) involved. Similarly, i/'Cto is equal to a sum of forms such as y(ay e). We shall write y'Bx = 2 ^(a*, et.), y'Cx = 2 y(aiy e{) i i to express these facts. When we use this notation it is to be understood that each fi{ai9 et) is paired with a y(oci9 e{) having the same <xiyeiy and involving the same indeterminates, in the same order, and that two terms fi(aiy et) and fi(ajy e^) have no indeterminates in common. Similarly, y(oc{y et) and y(ocjy e^) have no indeterminates in common. The results of § 4 can now be stated as Theorem I. If B and C are each non-singulary the symmetric and skew-symmetric forms y'Bx and y'Cx can be simultaneously reduced to the canonical forms 2 /?(<**> ^) ^nd 2 y(ociy e{). i i We must now extend this theorem to the case in which B or C is singular. We introduce five different pairs of forms (l) PM = (00*26-1 + 026-1¾) + (01*26-2 + 026-2¾) + • • • + (06-1¾+ 0e*6-l)> 7i(e) = (0^-1-026-1¾)+ (02*26-2-026-2^ + - + (0e-l*6+l"~ 0e+l*e-l)> where 7i(l) = 0; (ii) y?2(e) = {y1x2e^1 + y2e_1x1) + (y2x^2 + y2e^2x2) + ... + (06-1*6+1+ 06+l*6-l)f 72(^) = (00*26-1-026-1¾)+ (01*26-2 ~ 026-2¾) +— + (06-1¾-06«6-l). where fi%{!) = 0;
5. TRANSFORMATION OF THE GENERAL CORRELATION 397 (iii) /?8(e) = (2/oSfc-a+2/2e-2*o) + (2/1*26-3+2/2«-3*i) + — + (2/6-2*6 + y*Xe-t) + 2/6-1*6-1» y3(«) = (2/1^-2-2/26-2^1)+(2/2^-3-2/26-3^2)+... + (2/e-i*e-2/«^-i)» where #,(1) = y0*o» and y3(l) = 0; (iv) &(e) = (&«„_!+y^e-iVi) + (2/2*26-2+2/26-2*2) + — + (2/6-1*6+1 + 2/e+l*6-l) + 2/e*6> 74(e) = (t/„*26-l - 2/2e-l»o) + (2/l*26-2~ 2/26-2*l) + — + (2/6-1^-2/6^-1), where ytf4(l) = y^; (v) #>(«) = (2/0*26 + 2/2e*o) + (2/1*26-1 + 2/26-1 *i) + • • • + (2/e-i «e+i + 2/6+1*6-1). 7e(«) = (2/1^26-2/26^1) + (2/2^26-1-2/26-1^2) + ... + (2/6^6+1-2/6+1^). Our object is to obtain the following generalisation of Theorem I: Theorem II. If B, G are, respectively, symmetric and skew- symmetric (n+l)x(w+l) matrices, the bilinear forms y'Bx, y'Cx can be simultaneously transformed to S^a^e^ + SA^ + SA^ + S^W + E^O + S^K). i j k I m h and 2 y(<z*, *i) + 2 ri(^) + 2 yjfih) + 2 yM) + 2 y4(«J + 2 y5(«*) t i A; J m ft respectively, with the conventions previously introduced for the pairing of the terms and for the indeterminates involved in the two expressions. We consider the equations Bx = 0, (1) Gx = 0. (2) Let these define, respectively, linear spaces 8rirl9 S,^ of dimensions rx— 1, r2— 1, and let these spaces intersect in a space Sro-1 of r0— 1 dimensions. We transform our coordinates, choosing a new simplex of reference in the following way. In 5ffo-1 take r0 independent points <4n_fp+1,..., An. In Srx_x there exist rx - r0 further independent
398 IX. CORRELATIONS points -4n_fl+1, ...,^4n-fo, and in S,^ there are r2 — r0 independent points which are independent of An_ro+ly..., An, say A A The points ^4n_ri_r8+ro+1,..., An are linearly independent, and define the join of the spaces Sr v Sr^v We complete the new simplex of reference by choosing n —^-r2 + r0+l points in [n] which are independent of those already chosen, denoting them by A A In the new coordinate system Sro__x has the equations xQ = ... = %n-r0 = 0, ^-1 kas *ke equations x0 = ... = xn_ri = 0, and Sr x has the equations xQ = ... = #n_ri_ri+ro = xn_ri+1 =...= xn_ro = 0. (3) In this coordinate system y'Bx is a form in the indeterminates (x0,...,#n_fi), (y0, ...,2/n-rx) only> and y'Ck is a f°rm in tlie ^determinates , V (#0,..., a?n_ri_f8+ro, #n_ri+1,..., xn_ro)y only. In proving our theorem we shall use a series of transformations of coordinates, but we shall arrange matters so that at every stage the equations of Sy^, Sri_x and S,^ are given in the new coordinate system by (3). The reader will readily verify that any transformation of coordinates which ensures this result is of the form £ = PXy ri = Py9 where P = D E G L 0 F 0 M 0 0 H N 0 0 0 K the matrices in the first, second, third and fourth rows (columns) having n —rx —ra + r0+l, r2 — rQy rx — r0 and r0 rows (columns) respectively.
5. TRANSFORMATION OF THE GENERAL CORRELATION 399 Our proof of Theorem II will be by induction on the number of indeterminates xiy y^ involved, and for the purpose of the induction it is found convenient to impose a further restriction on P by requiring that the elements of F and H above the principal diagonal are zero, and that K = I . The transformation is then called a permissible transformation. We shall prove that y'Bxy y'Gx can be reduced to the forms given in the enunciation of Theorem II by a permissible transformation, after the preliminary transformation has been made. We commence our induction proof by considering the case n = 1. y' Bx = ay0x0 + b{y0x1 + ytx0) + cyxxx, y'Cx^kiyoXy^-y^o), before any preliminary transformation is carried out. We have three cases to consider: (i) If B and C are non-singular, r0 = rx = r2 = 0. No preliminary transformation is required, and any non-singular transformation is a permissible one. The result follows from the consideration of the simple case treated in § 4. (ii) If B = 0, C 4= 0, rx = 2, r0 = r2 = 0, and y'Cx is not essentially altered by the preliminary transformation. The permissible transformation £=/1 0\x [0 Jc) transforms y'Cx to 72(1), an(i y'Bx to /?2(1). If C = 0, J5 + 0, r2 == 2, r0 = rx = 1 or r0 = fi = 0. If rx = 1, a preliminary transformation reduces y'Bx to the form ay0x0, and the permissible transformation completes the reduction to fiB(l). We take y'Cx in its transformed form as y3(l). If rx= 0 we need no preliminary transformation, and the restricted transformation of § 2, Th. I, which is permissible in this case, transforms y'Bx to the form x0yt + xty0 (which is /^(1), and can be paired with y^l)), or to the form ay0x0 + cyxxv This last form can be reduced by a permissible transformation to the form Voxo + Vl xi - Thus we have the reduction y'Bx = /ff3( 1)+^3( 1), ^ = 73(1) + 7^(1).
400 IX. CORRELATIONS (iii) The only remaining case is that in which G is not singular, and therefore k is not zero, and B is singular, but not zero, so that ac = 62. Then rQ = ra = 0, rx = 1, and the preliminary transformation reduces the forms y' Bx> y'Gx to forms of the type %o*o P(y 0*1-yi*o)> where d and p are not zero. The transformation £= l-d* 0 \x /-d* 0 \; \ 0 pd-i) is permissible, and reduces the two bilinear forms to £o?7o> ?i£o-?7o£i> which are yff4(l), y4(l) respectively. We may therefore assume as hypothesis of induction that Theorem II holds for a pair of forms y'Bx, y'Gx which can be expressed in terms of the indeterminates (y0,..., ym), (%...,«J, where m < n. Now, our preliminary transformation expresses y'Bx, y'Gx in terms of the indeterminates (y& .-.,yn-r^, (%0> ~->xn-ro)> anc* n°t in terms of fewer indeterminates. By the hypothesis of induction we can reduce these forms to canonical forms, as required by Theorem II, by a permissible transformation if (i) r0 > 0, and (ii) the indeterminates (y0, ...,2/n-r0)> (xo>"->xn-r0) are subjected to a preliminary transformation. But no preliminary transformation is necessary since, in the new coordinate system r'0 = 0, r[ = r1 — rQi r'% = r2-r0, and, if n' = n — rQi the equations of S^^ are #0 = ••• = Xn' = ^> those of ^, are _ _ and those of /Sy^ are #0 = ... = a?n'-r'i-r'i = xn'-r\+l = ••• = #n' = ^> in accordance with (3). Hence the permissible transformation transforms the bilinear forms in the indeterminates (y0, ...,2/n-r0)>
5. TRANSFORMATION OF THE GENERAL CORRELATION 401 (x0, ...,#n-r0) into *he required forms. Since we obtain from this the permissible transformation £ = ID 0 0 0 \ x D E G ,0 0 F 0 0 0 0 H 0 0 0 0 ** on the indeterminates (y0i ...yyn)y (x0i..., xn), which also reduces y'Bx, y'Cx to canonical form, our theorem is proved if r0>0. We may therefore now confine ourselves to the case r0 = 0. Now suppose that rx > 0. A preliminary transformation gives us n—rx y'Bx = 2 byytz,, o n y'Cx = YloaVi^ o n where 2' denotes summation over the values o 0, ...,^-^-^5 71-/^+1,...,71, (4) of the suffixes i, j. In order to avoid the inconvenience of having to introduce new symbols for the indeterminates with each successive transformation, we shall start each step of the argument afresh with the indeterminates (x0,..., xn), (y0,..., j/n), except in cases where this would cause misunderstanding. Using the transformation of §3 applied to y'Cx, we have to consider only Case (ii), in which not all cin = 0, since, if cin = 0 for all values of i, r0>0, contrary to hypothesis. The restricted transformations of that section are permissible, because of the absence of terms c{i in y'Cx in which either i or j lies between n — rt — r2 + 1 and 7i — rv both values inclusive. Hence, using the same symbols as above, a permissible transformation reduces our bilinear forms to n—rx y'Bx = 2 byytx,, o n-l y'Cz= 2' 0^^ +(ynxk-ykxn)y 0 • n-l where k is one of the suffixes given by (4), and 2' omits the suffixes o n and fc, besides those already mentioned. We now have two cases: HP I 26
402 IX. CORRELATIONS Case (i). k > n — rv The forms n—rx n— 1 0 0 are forms in the indeterminates (XQ, ..., Xk_v #&+i> ..., #n-l)> \2/o> • • • y Vk-1> Vk+V • • • > 2/n-l) for which r£ = 0, r^ = r2 — 2, rj = r2, n' = w—2, and we see that these forms require no further preliminary transformation. By the hypothesis of induction, they can be reduced to canonical forms by a permissible transformation, and it is easily seen that a permissible transformation on the indeterminates (xQy...,xk_v xk+1>...,#n_!) is still permissible when we add to the equations of transformation the further equations bfc = xk> bn = xn> in their proper place. Since the theorem is proved in the case k > n — rv Case (ii). k^n — rv Since cknykxn was a term in the prepared form y'Cx, we must have k^n — rx — r2. Since a transformation of coordinates which permutes x0, ...,#n_ri_ra, but does not affect the remaining xi9 is permissible we may, after a suitable permissible transformation, take k = 0. We now write n-rx y'Bx = y0£+vxo + S hVixP n-l y'Cx = ynx^y^xn+ 2' c^y^ n-l where the summation £' is over the values l 1,...,71,-^-7-2,71-^+1,...,71-1 of the suffixes i> j, and n-n n-n 0 0 where 260 = 600, and 6^ = 6i. (6)
5. TRANSFORMATION OF THE GENERAL CORRELATION 403 Now, by hypothesis, the matrix (6) is of rank n — rx+l. We first prove that this implies that the submatrix (7) 260 W • Pn-n bi &u • ^n-rx 1 h &12 • • • • • • ®n—ri ^ln-n ^i—f^n-iY ) \®n-ri 1 • On—rx n—rv is of rank n — rx or n—rx — 1. For, suppose that the rank of (7) were less than n — rx—l. Then there would exist at least two distinct relations n_ri 2^6^ = 0 (j= 1,...,71-^), i = l n—rx 2 /^ = 0 (j= 1,...,^-^), connecting the rows of (7). If p, or are chosen so that n-rx n—rx p S A^ + cr S /*<&< = <>, l l then the multiples 0, pXx + <r/il9..., pAn_ri + cr/in„ri of the first, second, ..., (n —rx+l)th rows respectively of (6) are linearly dependent, and (6) is singular, contrary to hypothesis. Case (iia). Suppose that the matrix (7) is of rank n — rl9 and consider the forms n— rt n—1 l l For these forms in the indeterminates (xv ..., #n_i), {yl9 .-,2/11-1)we have n' = n-2, ^ = ^-1, r'2 = r2. Hence no preliminary transformation is necessary, and by hypothesis there is a permissible transformation which reduces these to canonical form. The equations of transformation, together with the equations r _ » __ 26-2
404 IX. CORRELATIONS are seen to define a permissible transformation of the indeter- minates (#0, ...,¾). Before making this transformation, however, we make the permissible transformation Xq = Xq, xi ==^.-¾¾ (* = 1,...,^-7^), Xi^Xi (i = 71-7^+1,...,71) on the original forms (5), our object being to simplify the term y0£, + rfx0 before carrying out the final permissible transformation described above. We find that n—rx n — i\ y'Bx = y0{b„x0+ 2 Mi ~ »o S ^cf) l l + *o(&o£o + ZbiVi-yo 1,'biCA n—rx + 2 M%-^)(^-¾¾) = 2b0xQyQ + yQ£ + i)xQ+ 2 b^y^ (n-r, n-r, \ 2 ftyCi^-2 2 Vd^O, and _ n-n-r, _ _ _ _ y'Ck = yn*o - 0o*n +- 2 fy(y< - c<y0) (¾ - Ci*o) + 2 ^¾¾ n-ri+l n—1 n — r^ — ri n—ri — rt + 2' ^¾¾ - y0 2 ^¾¾ - £0 2 tyCjy<, n-fi-r, since x0i/0 2 cijcicj = 0. In the first expression we have written n—n n—rx 1=2 &<»<. ^=2 &<y<. i i
5. TRANSFORMATION OF THE GENERAL CORRELATION 405 Now, since the matrix (7) is non-singular, we can choose the constants cv ...,cn_fl so that This simplifies the expression for y'Bx to _ __ n-rt /n-rt \ %boxoy0+ S b^y^-l £ 6<c,ls0y0 say. Now make the further permissible transformation xf = ¾ (i = 0, ...,n-l), xn=zXn+ £ cijcixj* We then have y'Bx = py$x$ + 2 b^tjfxf, .71-1 y'Cfc = ^4-*/*** + £' c«y*«*. Now p cannot be zero, otherwise J? would be of rank n — rv instead of n — rA + 1. Hence, by another permissible transformation which consists of replacing p*x$ by a?}, we may take p = 1. iVow apply the hypothesis of induction to the forms n—n n—1 and, as explained above, we deduce a permissible transformation of the indeterminates (xj,..., x*), (y*,...,yt) which gives us y'Bx = ^^ + SP(ai,ei) + ZZpi(ei)1 y'Gx = ^atf-^xS + iTyfo, ^) + 2727^(¾). Since these forms may finally be written y'Bx = faW + Zftat^ + ZZfifa), y'Cx = 7,(1)+2:7(^, «<) + 2727y,(e,), this completes the proof of Case (iia).
406 IX. CORRELATIONS Case (ii&). Suppose that the matrix (7) is of rank n — rx— 1. Then the equations n_ri 2 6^ = 0 (i = 1,...,71-^) have just one solution. Apply the methods of § 2 to reduce 2 bijViZj to canonical form. The restricted transformation of that section can evidently be amplified to a permissible transformation of the indeterminates (x0, ...,xn), (y0,..., yn). Now consider the canonical (n — rx) x {n — r^ matrix corresponding to the reduced n—ri form 2 bijijiXj. Since this matrix is of rank n — rx — 1, it is clear that one row consists entirely of zeros. That is, preserving the same notation, , - ,. , v t>ih = 0 (»= l,...,n-f!), for some value of & (1 ^h^n — rx)9 so that the one solution of the equations above is now / .A *Av We have two cases to consider\h^n — r1 — r2,h>n — r1 — r2. Case (iiftx). h > n — rt — r2. We have, with the previous notation, n—rx y'JSa = y0£+?Fo+ 2* &<iy<*/f y'c* = ynxo-yoxn+ 2'w*v> where the summation Z* omits the value fe. We now proceed as in Case (iia), making the substitution xo = xo> xi = Xi-CiXo (i = 1,...^-1,^+1,...,71-^), Xi = ¾ (i = n-rx+1,...,^). This is a permissible transformation, and gives us y'Bx = ^0[-2S* 6,0,+ S* W>]+2* &ii^^ +M^o^+^^o)+yo 2* Mi+^o 2* M* 0 0 n—n n-n
5. TRANSFORMATION OF THE GENERAL CORRELATION 407 We can choose c^,...,ch_l9 ch+1>...,cn-- so that n—rx the matrix of coefficients having rank n — rx — 1, and then y'Bx = y0(i>*o+2%) + (m + ^A)*o + 2*Mi*i- Since I? is of rank n—rx+l, j cannot be zero. At this stage we have y'Cx = ynx0-y0xn since n — rx — ra < & ^ w — rv The permissible transformation x* = #* (* = 0, ...,?&—1), _ n-n-r, _ ^==^71 + S cijcixj leaves y'JSx unaltered, and transforms y'Cx into n-l y'Cx = yM-^o^ + S'^***- As a final transformation on we write & = «? (i = 0,...,*-l,A+l,...,n), S* = ^4 + 83¾. This is a permissible transformation, and gives us n—rx y'Bx = Vo£h+Vh£o+ S* 6<i^^, n-l y'Cx = yn£0-VoL+ S' %7i^- We now make use of the hypothesis of induction to reduce Z*Mi& and ^'ct)Vd)
408 IX. CORRELATIONS to canonical forms by a permissible transformation on the indeterminates^, ...,^-i,^+i,..., L-i), (Vv •••> Vh-v Vh+v ...> Vn-i)>noting that no preliminary transformation is necessary. For this set of indeterminates, n' = n-3, rjssfj-l, r'2 = r2-l, and the transformation induces a permissible one of (£0, ...,£„), (Vo> •••>?»)• We have, finally, that is, y'Jfcc = y?6(l) + 27y?(ai,e,) + 2Ty?,(ei), y'C* = 76(l) + rr(ai, eJ + ZZyfa). Case (iife2). /^^-7^-^. By means of a permissible transformation, which interchanges the suffixes 1 and h, we have, preserving the notation, y'Bx = y0£ + ^o+ 2 b^y^, 2 n-l y'Cfc = ynxo-yoXn+ 2' ^¾. 1 In considering Cases (iia) and (ii6x) we left the forms unaltered until the end, and then reduced them to canonical form. In this case it is necessary to reduce the forms n—ri n— 1 2 b^y^j and £' c^a, 2 1 to canonical form at this stage. In applying the hypothesis of induction to the two forms, we take the indeterminates in the order #2, • • •, #n_ri_ra, #n_ri_r2+i, .... ^n-rjj #i> #n_ri+1> • • • > ^n-1* We see that for these indeterminates, ri = n-2, r[ = rv *a = r2, and no preliminary transformation is necessary. Furthermore, a permissible transformation in these indeterminates induces a permissible transformation in (x0,...,xn), (y0,...,yn). For the sub- matrix H which appears in the matrix of a permissible trans-
5. TRANSFORMATION OF THE GENERAL CORRELATION 409 formation has only zeros above its principal diagonal, and therefore the equations of such a transformation are, as far as the displaced indeterminate xx is concerned, £1= S aixi + bxv 2 Hence, when xx is replaced, and the additional equations £0 = xo> are added, we have a permissible transformation of {xQ,..., xn), (2/o> '">Vn)' At this stage we have y'Cx = ynx0-y0xn + Zy(aiiei)^Ui:yi{ej)i where £, y involve the same set of indeterminates as before the transformation, with different coefficients. For simplicity we preserve our previous notation, and write n—ri y'Bx = y0£+i}x0 + 2 btjytxit 2 n-1 y'Cx = ynx0-y0xn + 2' cyy^, n—r! n—1 w/&ere 2 ^iiVi^j and 2' cijVixj are ^n canonical form. We now 2 1 proceed as in Case (ii6x), this time taking h = 1, and obtain n—ri 2 n-1 y'Cx = y„x0-t/0xn+ 2' CuV&y l w^ere tAe coefficients bti, cti are unchanged by these final transformations. Hence we have y'Bx = y0x1 + y1x0 + Zfi{at, et) + EZfifo), y'Cx = yn*o - y0xn+Zy(*i> «<)+zzyfa)- The indeterminates xlt yx occur in <£y(««.«i)+££y«(«j). but not in Zft^tt + ZEMt,).
410 IX. CORRELATIONS It follows that xv yx appear in some term y^). On the other hand, #o> y& xn> Vn do not appear in either of these two expressions. Suppose that xly yx appear in The corresponding ^) = (^^0,+2/0,^) + -^ where d\ j c^oy • • • are different from 0,1, n. Now, we may write (2/0*1 + 2/1*0) + (Va^ar + Va^a) + • • • = fifa + *)» {yn^o-y^n) + {yiXai-yaix1) +... = y<(e,+1), the terms in /?<(fy +1), y^+l) being taken in the order •^n* x0> xl> ^oi* xo,> •••• This completes the proof of Theorem II when rt > 0. We suppose now that r0 = rx = 0, r2 > 0. The analysis in this case is similar to that followed when r1>0, and it will only be necessary to indicate the changes. After our preliminary transformation we have n n—rt y'Bx = E&a&fy jfOx = 2 c^y^. 0 0 We may assume that not all bin = 0, for in that case the hypothesis of induction would apply immediately. Hence the permissible transformations of § 2 give us either n-l y'Bx = 2 byytXj + kynxn (k * 0), o and ,n n~r« o n-l or y'Bx = 2' b^y^ + (ynxk + ykxn)9 o and ,n n"r* y Cx = 2 %*/<*;• o In the former case the permissible transformation xi^xi (i = 0,..., n-l), xn = k*xn
5. TRANSFORMATION OF THE GENERAL CORRELATION 411 enables us to take k = 1, and by applying the hypothesis of induc- n—1 n—rt tion to 2 bijVixiy 2 cijl/ixj> we finally obtain 0 0 y'Bx = Z£(0,,6,) + ZZfifaHMl), y'Cx = 2:7(0,,6,) + 2:2:7,(6^) + 73(1). We now consider the second alternative: Case (i). k>n — r2. This is dealt with in exactly the same way as Case (i) above, interchanging the roles of B and C. We find that y'Bx = 27/?(o„ 6,) + 2^,(6,)+^(1), y'ttr = 2:7(0,,6,) + 2:2:7,(6^) + 7^1). Case (ii). If k^n — r2, the interchange of suffixes 0 and k is a permissible transformation, and we may write n-l y'Bx = 2 &<#<**+ (yn*o + yo*n)> n—rt 0 The methods followed above, for rx > 0, still apply, interchanging the roles of B and C, but since G is skew-symmetric and therefore always of even rank, only Case (iife) appears. Case (ii6x) gives us y'Bx = 27y?(o,,e,) + r2;/?,(6i)+y?6(l), y'Cx = 2:7(0,, 6,) + 2:2:7,(6,) + 75(1)- Case (ii&2) leads to a result exactly similar to Case (ii&2) above. Finally, if r0 = rt = r2 = 0, we are in the simple case, and the reduction of § 4 applies. 6. Canonical forms for correlations. Having completed the reduction of the forms y'Bx, y'Cx to the canonical forms y'#r=£/?(o,,6,)+£ SA(eJ), 2,'C* = 27(o,, 6,)+s Sy«(«J), 1 i=i j-i we arrange the indeterminates involved more conveniently so that the 26, indeterminates involved in y#(a,, 6,) and 7(0,, 6,) are a:2(c1+...+ci-1)> •••> ^2^+...+e,)-l>
412 IX. CORRELATIONS We then order those in ^(e^), y*(fy) so that their suffixes exceed the suffixes of the indeterminates in fia(eb), ya{eb) if a < i, or if 6 < j vhen a = i. We now consider the matrix of A = B+G. We have at once A = ID^cc,) DM \ M *iK) *i(ej) (1) F,{<) J the diagonal submatrices being defined as follows: De(a) is defined in the proof of Th. Ill, § 4. It is a 2e x 2e matrix, and De(a) — AZ)g(a) has the elementary divisors (A — ot)e9 (A — a-1)6 if a 4= 0, and Ae if a = 0. J^(e) (i = 1,..., 5) is the matrix of the bilinear form fii(e) + 7iie). It is a simple matter to verify that these matrices are: (i) FM = 1 1 -1 I ll -1 1 1 The methods already used for determining elementary divisors show that J?i(e) - XF[{e) has (A-l)«, (A-l)6 as its elementary divisors.
6. CANONICAL FORMS FOE CORRELATIONS 413 (ii) *,(«)- I 1 1 \ i l l L -:: -1 \ -1 1 ! / and Ft(e)—&F'2(e) has the elementary divisors (A+ l)e, (A+ l)c. (iii) F,(e) = 1 1 1 1 1 -1 1 -1 1 1 1 1 -1 1 1 1 1 1 1 • • 1 1 1 and F3(e) — AF'3(e) has the one elementary divisor (A— l)2*-1. (iv) FA(e) = -1 -1 1 • • -1 1 -1 1 1 1 I 1 1 1 / and Ft(e) — \Fi(e) has the elementary divisor (A+ 1)26. Finally,
414 IX. CORRELATIONS o '• 1 11 -l l ~i I and the (2c + 1) x (2c + 1) matrix Fb{e) - \F'b(e) (A indeterminate) is of rank 2c. Now, if A — \A' is non-singular for indeterminate A, there are no zero matrices on the diagonal of (1), nor are there any F5(e) matrices. We can now give Th. Ill, § 4 in its complete form as Theorem I. Necessary and sufficient conditions that there exists a matrix A such that A — \A' is non-singular and has assigned elementary divisors are: (i) the elementary divisors, other than those of type Ae, (A— 1)^, (A+ I)*7 can be arranged in pairs (A — a^, (A —a^1)^; (ii) the elementary divisors (A— 1)^, for a given even integer /, are even in number; (iii) the elementary divisors (A +1)^, for a given odd integer g> are even in number; (iv) if the elementary divisors are Aci,..., A6*, (A— 1)A,..., (A— 1)A, (A+ 1)*,...,(A+ 1)^, (A-a1)\(A«ar1)\...,(A-ad)^, (A-aj1)^, then 2Se< + S/* + ifc + 2 2A< = n+l. iii l The necessity of these conditions is an immediate consequence of the reduction to a canonical form and of the results given above, which give the elementary divisors of the various submatrices in the canonical matrix (1). The sufficiency of the conditions is evident, because we can construct a matrix A as given by (1), after grouping the elementary divisors into suitable pairs.
6. CANONICAL FORMS FOR CORRELATIONS 415 The canonical form (1) may be simplified to some extent. We observe that Fx(e) — XF[(e) has the same elementary divisors (A- l)c, (A- l)e as De(l). Hence, by§ 4, Th. I, there exists a non- singular matrix P such that P^(e)P = De(l). For similar reasons we can find non-singular matrices Q, R, 8 SUch that r\>vii \n t\ i i\ B'W.e) 0 \R = D2e_1(l), \ 0 Fz(e)) and S'(FA(e) 0 \ S = 2)to(-l). Hence, by a further transformation of coordinates, we can assume that (1) contains no submatrices Fx(e) or F%(e), and that for given fif gi there is at most one ^(e), and one -F4(e). In future we shall suppose that this has been done. We now consider the case in which A—XA' is singular, for indeterminate A. In the canonical form of A the number of zero rows (and columns) is r0, where n — rQ 4- 1 is the rank of the matrix a and therefore n — r0 4-1 is also the rank of the matrix (})■ This rank is invariant under transformations of coordinates, since I P'AP \ = (P'AP\ = P'IA\P. \(P'AP)') \P'A'P] \A'J If the canonical matrix A contains k matrices F5(e), we see that the rank of A — \A' is n — r0 +1 — k. Hence k is also an invariant of A. We now prove that the ef which appear in the submatrices -F5(e$), and which determine the number of terms in y'Ax corresponding to each submatrix -F6(ef), are also invariants of A. We assume that k > 0.
416 IX. CORRELATIONS We consider the linearly dependent rows of the matrix (A-AA')x, writing _- nz,r» . v ,. %i = 2 (^-Aa^)^ (% = 0,...,n-r0). These linear forms have k independent relations connecting them, say n_r ¢= S^(A)Ii = 0, (2) where we may assume that the d^(A) are polynomials in A. Amongst all possible relations of the form (2) we consider those in which the maximum of the degrees in A of the polynomials d0,..., dn_ro is least. Let this smallest maximum be mv and suppose that there are qx relations (2) of this degree, say We now consider all relations (2) which are not of the form l and amongst these we again consider those of lowest degree, which degree we denote by m2. If is a basis for this set of polynomials, we then consider those relations (2) of lowest degree which are not expressible in the form and so on. We obtain a set of polynomials which form a basis for the relations (2). We prove that the integers #i> •••>?,» mv -~>mP are invariants under allowable transformations of coordinates. If ^- x = Px is such a transformation, the relation becomes Zd^q^Xj = 0, where Q = (q^) = P.
6. CANONICAL FORMS FOR CORRELATIONS 417 Since q{j is in K, the degree of any relation is unaltered by an allowable transformation. This proves the invariance of qv ...,qp, mv ...,mp. We now associate these integers with the e\. In the canonical form (1) the only submatrices 0 such that G — XG' is singular are the k matrices Fb(e\). Since all the sub- matrices are on the principal diagonal of A, we obtain relations (2) only from the Fb(e\) matrices, and there is one relation corresponding to each matrix. Let Fb(e)) have rows and columns corresponding to the indeterminates xm> • • • 9 xm+2e (e = ej)* Then Xm = (1 - A) xm+2e, Xm+1 = (1 - A) aW2*-l +(!+*) Xm+2e> -^■m+2 = (1 ~~ *) xm+2e-2 + (1 + A) #m+2e-l> • •••••• -^-m+e-1 = (1 "" A) #m+e+l + (1 + A) #m+e+2> Xm+e = (l+A)^m+e+i, Xm+C+1 = (1 ~ A) #m+e-l - (! + ^) xm+e, • •••••» Xm+2e= (1 -^)^-(1 + ^)^+1- The relation of least degree connecting Xm,..., Xm+2e is ^s(l + A)«Zm-(l + A)-Ml-A)Xm+1 _(l+A)e-2(i_A)2Zm+a...±(l-A)«Xm+e = 0, and ^lf..., ^fc form a basis for the relations (2). We deduce that if ef<e|<...<ejj, then and therefore the integers ef,...,e\ are invariants of .4 under allowable transformations of coordinates. We sum up our results in Theorem II. Let A be any (n+ 1) x (n+ 1) matrix, let (A, A') be ofrankn — r+l, let A — AA' be of rank n — r -f 1 — k, and let HPI *7
418 IX. CORRELATIONS be a basis of polynomials of minimum degree for the relations connecting the rows of (A-XA')x. Let the degree of fa be mit and suppose that mj = ... = mQl - %<mft+1 = ... = mq% = n2< ... Let the elementary divisors of A—XA' be A%...,AS (A-l)Vi-i,...,(A-l)»»-i, (A +1)^,..., (A + 1)«*, (A —0¾)¾. (A —of1)*!,.... (A — ad)h*, (A — aj 1)/><*, where a{ may be ±1, andf1<f3<... <fb, gt<g2<... <ge. Then 2Ze + E(2f-l) + 2Zg + 2Zh = n-r+\-ki and there exists a non-singular matrix P such that the correlation « = Ax is transformed into the correlation v = By by the transformation x = Py, where B = P'AP = K(o) Ft(gc) FMi) K(mqp) 0/ q =» k, and the zero matrix has r rows and columns.
7. GEOMETRICAL PROPERTIES OF CORRELATIONS 419 7. Some geometrical properties of correlations. We begin with a simple correlation [§4], u as Ax, (1) this having been defined as one for which A +A' and A—A1 are both non-singular. The canonical form is given by (1), where A= /Dei(0) JUO) A,(0) A point x has no transform if it lies in the space given by Ax = 0. Remembering that De(oc) is the 2e x 2e matrix DJL*) = a 1 a 1 (2) a 1 1 0 1 0 I1 ° it is easily seen that the space given by (2) is of r — 1 dimensions, being spanned by the vertices -^•Ci> ^tex+Ct* •••» -^2(^+...+^.J+er> of the simplex of reference. This space may be called the fundamental space of the correlation corresponding to the zero latent root of A — XA\ Similar working shows us that the transform of any point which has a transform is a prime which always passes through the space spanned by -^0» -^2^ X^+ej, ..., -^2(^+...+^-^ «7-2
420 IX. CORRELATIONS This is also a space of r— 1 dimensions which, we see at once, is the space defined by the equations A'x = 0. (3) We may regard this as the fundamental space corresponding to the infinite latent root of A — XA\ We note that the spaces defined by (2) and (3) do not meet. The space given by (A-atA^x** 0, where cct is a latent root of A — XA\ is called the fundamental space of the correlation corresponding to olv Recalling that the dual correlation is v = A% (4) we see that a necessary and sufficient condition that a point x should have the same transform under (1) and (4) is that it should lie in a fundamental space corresponding to a latent root of A — /L4\ Let us consider the space defined by (A-^A^x^ °> and assume that 04=^ = ^= ... =a,, and a*4=ai (i>t). We see that this space is an 8(__v defined by the vertices The correlation (1) transforms the points of this St_x into the primes which pass through the Sn_t given by the equations #22^+^-1 = 0> X2£e+2fx+f%-l = 0> •••» ^22^+2(^+...+/^)+/^-1 = 0. This Sn_t contains the St_v and also contains every fundamental space corresponding to a latent root of A — XAf distinct from af1. The fundamental space corresponding to af * is seen to be an Sf^ spanned by the vertices 22Te> A2re+2/!> -*2re+2(/!+/2)> •"> -A2Te+2(/1+...+/i-|)'
7. GEOMETRICAL PROPERTIES OF CORRELATIONS 421 Hence 8n^ and #jLx meet in a space which is spanned by the vertices ^2re+2(A+...+/i)> ^2^+2(^+...+/<+1)> •••» -^2^6+^+...+/1-^ where we have assumed that/1}/2, ...,/^ are so arranged that and fi+vfi+2> ...,/<>L We now turn to the general correlation discussed in § 6. In the simple case discussed above, the possibility of latent roots equal to ± 1 was excluded. Such latent roots arise here. Before finding the fundamental space corresponding to A = 1 we remark that in discussing the geometrical properties of any correlation we are led to consider certain loci: (i) the locus of points which lie in their transforms under (1). These are the points whose coordinates x = (x0, ...9xn) satisfy the equation ^ _ %lA% = Q This equation may also be written (x'Ax)'=x'A'x = 0, or x'Bx = 0, (5) where B = \(A + A') is the symmetric part of A. (ii) the locus of points which have the same transform under (1) and (4). These points satisfy the equations (A-aA')x = 0, (6) where a is any latent root of A — AA', and were considered in the simple case. If x is any solution of (6), xrAx = aux'A'x = olx'Ax. Hence, if a 4=1, x<Ax = 0, and therefore every point which lies in a fundamental space corresponding to a latent root a 4= 1 satisfies equation (5). We assume that A is in the canonical form described in § 6, and suppose that a = 1 in equations (6). These become Cx = 0, (7) where C is the skew-symmetric part of A. In solving these equations
422 IX. CORRELATIONS we observe that the only submatrices of A which yield any non-zero coordinates are those submatrices 0, say, such that Q—G' is a singular matrix. These are the submatrices 2),(1), Fz(e), FB(e). It is therefore sufficient to consider the case in which ■4- IDJI) r By direct calculation we find that a basis for the solutions of (7) is given by the set of vertices Xo> Xex> ^2e^ -^2ex+ei» •••> ^2(^+...+^-^+¾ (2a points), ^r2e> -^r2e+2/x-l> •••> ^26+2(^+...+/^)-(6-1) (& points), -^2Te+2r/-6> ^-2re+2r/-6+2(7i+l> •••> ^2^+22:/-6+2(^+...+0^)+(0-1) (c points), Xm,Xm+1, ...,1, (m = 2re + 2r/+2rgr~6 + c). The point P = S^X^ where the summation is over the suffixes contained in the above table, lies on the locus x* Ax = EdyXiXj = 0 if and only if d'-l 2(A0A!+ ... + A2d_2^2d-l) + 2 ^1^+...+^)+^ = 0, where i""° «i = ••• =«*= If Cd+i^f and/!= ... =/<*'= l,/d'+i>l. A correlation may be singular for several reasons. If we consider the canonical form, we see that A is singular if (i) there are zero matrices on the principal diagonal; (ii) there are submatrices De(0) on the principal diagonal; (iii) there are submatrices F5(e) present; and (iv) if there is a combination of (i), (ii) and (iii).
7. GEOMETRICAL PROPERTIES OF CORRELATIONS 423 If a correlation is singular because of (i) only, we shall say that it is of the first kind. If it is singular because of (ii) and (ill) only, we shall say that it is of the second kind. If a correlation is singular on account of (i), by itself, or on account of (iv), the canonical form is (t !)• where Ax is an (m +1) x (m +1) matrix such that the correlation u = Axx (8) in the 8m whose equations are ^m+l = ••• — xn = 0 is either non-singular, or singular only because of (ii) and (iii) above. If Sn_m_x is the space Xq = ... = xm = 0, no point of this space has a transform under the given correlation in [ri], but the transform of any other point, if it exists, passes through Sn_m_v We thus obtain the following construction for the given correlation in [n]; if P is any point not in Sn_m^ly project P from Sn_m_1 into the point P' of 8m. Let /7' be the transform of P' by the correlation (8) in Sm, and let 77 be the prime joining 77' to Sn_m_v Then 77 is the transform of P by the given correlation in [n]. From this result it follows that the geometry of any correlation in [n] can be deduced from the geometry of a correlation of the second kind. The properties.of a singular correlation of the second kind are most easily discussed by showing that such a correlation can be represented as the product of a non-singular collineation (of a kind which is, geometrically, particularly simple) and a correlation (in the same space) of the first kind. We do this below. Ultimately, therefore, the properties of singular correlations can be made to depend on those of non-singular correlations in a subspace. Let u — Ax be a correlation of the second kind, where we suppose that A is in canonical form and, to simplify our notation, we omit the submatrices 2)e(a)(a*0), P3(e), P4(e). Let 2)^(0),...,2)^(0) be the
4Z4 IX. CORRELATIONS diagonal submatrices corresponding to the latent root A = 0, and Iftt Be the diagonal submatrices of type F5(e). The space of points having no transform in the correlation is defined by the equations Ax = 0. By direct calculation this space is seen to be the Si+C_^1 spanned by the t points and the c points whose coordinates are of the form (0,0,...,0, 1,1,...,1,0,...,0). 01+1 Each one of these c points arises from a submatrix -F^(^), and the non-zero coordinates are gt+l in number, their position corresponding to that of the first ^+1 rows of FB(g€) in the matrix A. Again, the primes which correspond to points in [ri] with a transform all pass through the space $/+<._! spanned by the t points and the c points (0,0,...,0, 1,-1,1,-1,..., ±1, 0,...,0). v „ ' 0i+l Here also, each one of these c points arises from a submatrix F^g^, and the non-zero coordinates, which alternate in sign, are gi + 1 in number, their position corresponding to that of the first gi + 1 rows of F^) in the matrix A. This space ^+c_x is evidently defined by the equations A, A It is immediately verified that #/+<._! and #/+<._! do not meet. This may also be seen from the fact that if the equations Ax = 0 and A'x = 0 have a common solution, the integer r0 introduced into our discussion, p. 415, of the canonical form is not zero, and therefore zero matrices appear on the principal diagonal of A, contradicting our hypothesis that the correlation is of the second kind.
7. GEOMETRICAL PROPERTIES OF CORRELATIONS 425 When we consider the points which have the same transform in the dual correlation as in the given one, we have only to concern ourselves with the submatrices ^5(^1),...,^5(^), since De(0) - AZ>e(0) is non-singular if A 4= 0. Now has a solution for each value of A. We see that all such solutions lie in the space _ __ _ _ #0 ~ xl — ••• — x2(ex+... +ei)-l — 0, X2Ze+gi+l =...= x2Ee+2g{ = 0, (10) • •••••• x2Ze+2(gi+...+gc-i)+(c-l)+gc+l = ••• = x2 Ze+2 Zg+c-1 = 0, where 2Ze + 2Zg + c- 1 = n+1. Now consider the collineation y«!+i %"M ¢ = 0,...,^-1), (1M) +i = ^ J — ^9.0. JL-e^JL-i, I ^2e1+i " *2e1+e,+< (< = 0, ...,e,-1), (11-2) 2/2(^+...+^)+^ = Z^+.-.+e^+eH-tj ^ = Q> _^_ ^ (n.^ 2/2^+...+^-1)+^ ~ ^(cx+.-.+ci-^+i i 2/2^+...+e*)+2i = X2(c1+...+^)+2i> 1 (lht') 2/2^+...+^)+21+1 = "" ^2(^+...+^)+21+^) where in (1W) the suffix i varies so that the equations range over all the remaining coordinates. This collineation is evidently an involutory collineation. If we write it in matrix form as y = Cx9 (ii) then C2 = In+V It transforms 8e+i^t into #c+/_i, and leaves the space defined in (10) invariant. Now consider the correlation u = ACy. (12) Since u = (AC) (Cx) = ACy = Ax, our given correlation of the second kind is now expressed as the
426 IX. CORRELATIONS product of the non-singular collineation (11) and the correlation (12). We show that the correlation (12) is of the first kind. In fact, actual multiplication of the matrices involved shows that the matrix AG is symmetric, and therefore the correlation (12) is necessarily of the first kind. The vertex of the primes u in the correlation (12) is $c+j-i> since the solutions of (AC)'y = 0 are the solutions of A, - A 'y = 0, the matrix G being non-singular. We conclude with some remarks on non-singular correlations. These are most easily discussed by means of the associated collineation r a ,m\ y = AAx. (13) The points, lines, etc., which have the same transform under (1) and the dual correlation are precisely the united points, lines, etc., of (13), and we have discussed the methods of obtaining these in Chapter VIII. We recall the theorem established in § 4, that the elementary divisors of AA—\In+ly which are those of A — XA\ cannot be chosen arbitrarily, but if A = a (a #= ± 1) is a latent root, so is A = a-1. Applying VIII, § 3, Th. V, gives us an interesting geometrical result. Let x be any point of [n] which is not a united point, and y its transform under (13). The line xy meets all the fundamental spaces of the collineation (13), the point of intersection with the fundamental space corresponding to the latent root a being Aa = y-ax. We see that the points Aa> ^«-1 are in involution, and if A = ± 1 are latent roots of the collineation, the points Av A_x are the united points of the involution. As we have already seen, the correlation u = Ax, when A is non-singular, gives a prime-point transformation y = Av. (14) For if x lies in the prime v, u passes through the point y, where u'y = x'Afy = x'v. We can develop a similar theory for this prime-point correlation,
7. GEOMETRICAL PROPERTIES OF CORRELATIONS 427 which is, indeed, the same correlation as u = Ax. In particular, the correlation associated with (14) is w = AAv, and since {AA) = AA, this is the dual form of (13). Finally, some aspects of the elementary theory of quadrics are of interest in connection with the theory of correlations. If the prime u which is the transform of the point x contains the point x> this point satisfies the equation ' x'u = x'Ax = 0, or, equivalent^, ^ = Q> (16) where B = \(A + A'). We call the locus of points given by (16) the incidence quadric associated with the correlation u = Ax. Similarly, the primes v which contain their transforms under (14) satisfy the equation v'Dv = 0, (16) where D = \{A + A~l). We see that B = A'DA. Let us assume that the correlation u = Ax, besides being non- singular, is such that B and D are non-singular. The point-equation of the locus (16) is x'lrhi = 0. For this to coincide with the incidence quadric (15) we must have kD-1 = B for some k in K. We determine the simple non-singular correlations for which the 'envelope' incidence quadric (16) is the dual of the incidence quadric (15). The above condition is 4Mn+1 = (A-l+l)(A+A') = 2ln+1 + lA+A-iAf, or AA + (lA)~i = 2(2¾ -1) /„+!. Let Y = AA.
428 IX. CORRELATIONS Our last condition is 7*-2(2fc-l)r + 4+1 = 0. Hence, since Y 4= A/n+1, the minimum function of the matrix Y is [cf.II,§10] ^(A) = A2-2(2fc-l)A+l. By II, § 10, Th. Ill, the last invariant factor of 7- \In+l is ^(A), and therefore all invariant factors divide ^(A). We consider the following possibilities: (i) 2& — 1 #= ± 1. If a, a"1 are the roots of ^(A) = 0, neither of these roots is equal to ± 1. The elementary divisors of 7 —AJTn+1, as we remarked in §4, occur in the pairs (A —a)ei, (A —a_1)ei. Since the product of all the elementary divisors is of degree n + 1, it follows that n+ 1 is even. Finally, since the last invariant factor is simply (A — a) (A— a""1), the elementary divisors of 7 —A/n+1 consist of |(n+l) elementary divisors (A— a), and an equal number of elementary divisors (A— a-1). Hence A can be reduced to the form A = /Dx{a) DM) where, in the usual notation, D^oc) = /0 a] (-)■ Conversely, it is easily verified that if A is taken in this form, with any a which is neither zero nor ± 1, then JcD-i^B (fc*0)f where , _ (a+1)2 4a (ii) 2Jfc—1 = 1. The minimum function is now (A—l)2, and the elementary divisors of Y — \In+1 are all of the form (A—l) or (A—l)2. Since, by §6, Th. I, there must be an even number of
7. GEOMETRICAL PROPERTIES OP CORRELATIONS 429 elementary divisors of the latter type, A can be reduced to the form A= /Fx(2) Ji(2) *1(2) 1 where [§6, (i)] >i(2) = 0 0 0 1 0 0 1 -1 0 1 0 0 1 1 0 o, We may again verify that if A is given in this form it satisfies all our requirements. (iii) 2k — 1 = — 1. The minimum function is (A +1)2, and A = — 1 is a latent root of the matrix F. It follows that B = \{A + A') is singular, contrary to our hypothesis. *
* BIBLIOGRAPHICAL NOTES Book I A considerable part of the material of Chapters i-iv will be found in works on Modern Algebra, such as Albert (1), Dickson (5), Macaulay (9) and van der Waerden (10). As far as we know, the development of the theory of linear dependence over a non-commutative field, given in Chapter n, is new, and the algebraic theory of Jacobians given in § 7 of Chapter in has not previously been published. In Chapter rv we have preferred to use the term resultant-forme instead of the literal translation inertia-forms of the Tragheitsformen introduced by Hurwitz. Book II Chapter v. The method given here for defining a projective space has been used by Lefschetz (8) for the case of a commutative field, and is derived from Veblen and Whitehead (11). The development for a non- commutative field given here has not previously been published. Chapter vi. Much has been written on the axiomatic foundations of projective geometry. For a full account of these writings, the reader is referred to the article by Enriques (6) and to Baker (2). The method followed here bears most resemblance to that adopted by Veblen and Young (12), but differs from their treatment in the choice of the initial axioms, and in the introduction of coordinate systems before the imposition of limitations which effectively restrict the field from which the coordinates are taken. The development in Chapter n of the theory of linear dependence over a non-commutative field was designed to make this treatment possible. Chapter vn. The use of coordinates to distinguish linear spaces originated with Grassmann and was developed into line-geometry by Plucker. References will be found in the article by Segre (7), and an account is given in Bertini (3). Properties of determinants which are in effect Grassmann coordinates play an important part in the algebraic theory of invariants. The basis theorem of § 7 was first proved by Mertens. Our proof is similar to that given by Weitzenbock [Proc. K. Akad. Wetensch. Amsterdam, vol. xxxix, p. 503 (1936)]. Chapter viii. The geometrical theory of collineations has been fully discussed by Segre. For references see Segre (7). An account is given in Bertini (3). The treatment given here is more algebraic than either of these. Chapter ix. The account of correlations given here is based on the simultaneous consideration of symmetric and skew-symmetric bilinear forms, which derives from the treatment of two quadratic forms given in Bromwich (4), who bases his work on that of Kronecker. For the geometrical aspects reference should be made to Segre (7). *
431 BIBLIOGRAPHY (1) Albert, A. A. Modern Higher Algebra (Cambridge, 1938). (2) Baker, H. F. Principles of Geometry, vol. i (Cambridge, 1922). (3) Bertini, E. Geometria Proiettiva degli Iperspazi (Messina, 1923). (4) Bromwich, T. J. FA. Quadratic Forms and their Classification by means of Invariant Factors (Cambridge, 1906). (5) Dickson, L. E. Modern Algebraic Theories (Chicago, 1926). (6) Ency. der Math. Wissenschaften, in, A, B 1, F. Enriques. (7) Ency. der Math. Wissenschaften, in, C 7, C. Segre. (8) Lefschetz, S. Lectures on Algebraic Geometry (Princeton, 1937). (9) Macaulay, F. S. Algebraic Theory of Modular Systems (Cambridge, 1916). (10) van der Waerden, B. L. Moderne Algebra (Berlin, 1930). (11) Veblen, O. and Whitehead, J. H. C. The Foundations of Differential Geometry (Cambridge, 1932). (12) Veblen, O. and Young, J. W. Projective Geometry, vol. i (Boston, 1910). *
* INDEX (The numbers refer to the page where the term occurs for the first time in the book or is defined) Abelian or commutative group, 3 of operations on a matrix, 73 Addition of points on a line, 253 Additive group, 3 Adjugate matrix, 81 Adjunction of algebraic element, 103 of indeterminate, 19 Algebra of points on a line, 253 Algebraic closure of a field, 96, 112 dependence, 108 element of extension field, 102 Algebraic extension over a field, 106 primitive element of, 126 simple, 101 Algebraic functions, 107 differentiation of, 128 Algebraic function field, 107 independence, 108 point of projective space, 179 Algebraically closed field, 96, 112 Allowable coordinate systems, 178 Allowable transformations, 59 of restricted type, 367 Alternating group, 5 Apolar vectors, 69 Associated collineation, 365 Associated coordinate systems, 186 Associative law of composition, 2 for multiplication of matrices, 52 Basis, finite algebraic, 108 • for extensions of finite degree, 105 for ideal, 142 for linear set, 45 Basis for linear space, 183 derived from Grassmann coordinates, 300 in incidence space, 209 Basis, minimal algebraic, 108 Basis theorem for Grassmann coordinates, 315 theorem, Hilbert's, 142 Bilinear correspondence, 365 form, 365 Binary form, 148 forms, resultant of two, 150 Block matrices, 53 Canonical form for collineations, 331 correlations, 411 A-matrices, 88, 92 matrices, 60, 62 null-polarity, 380 polarity, 372 projective transformation, 323 skew-symmetric bilinear form, 378, 380 symmetric bilinear form, 368, 372 Chain of perspectivities or projectivity, 217 Characteristic equation of matrix, 95 Characteristic function of matrix, 95 Characteristic of ring, 14 Characteristic matrix, 95 polynomial of algebraic element, 102 Coefficients of polynomial, 20 Cofactor of element in determinant, 77 Collinear points, 191 Collineations, 327 as product of null-polarities, 383 as product of null-polarity and polarity, 389 as product of polarities, 376 associated with correlation, 365 canonical form, 331, 332 cyclic, 352 fundamental space corresponding to latent root, 336 in [3], 350 induced in linear space, 342, 343 non-special, 336 particular, with only one fundamental space, 356 Segre symbol for, 350 similar, 328 singular, 358 special case, 336 united points of, 334 with only one united point, 344 Column of matrix as right-hand vector, 66 matrix printed as row matrix, 50 rank of matrix, 66 HPI 28
434 INDEX Commutative or Abelian group, 3 Complement of matrix, 72 Complete factorisation of element in integral domain, 29 ^Completely reducible polynomial, 112 Compound matrix, 291 Concurrent lines, 191 Condition for commutative field, 202 Conjugate points in polarity, 367 spaces in polarity, 375 Consequences of assuming Pappus* Theorem, 274 Construction for 0+ U(a+/5), 197 for O+Uafi, 199 Coordinate of point in a scale, 259 Coordinate systems, allowable, 178 associated, 186 related, 283 scale, 283 Coordinates, Grassmann, of linear space, 289 of a projected space, 308 Coordinates of point in a reference system, 262 set of, for point in projective space, 178 Correlation between distinct spaces, 362 Correlation, canonical forms for general correlation, 418 null-polarity, 380 polarity, 372 simple correlation, 394 Correlation, exceptional points, 362 geometrical loci connected with, 421 in [n], 364 non-singular, 426 permissible transformations for reduction of, 399 simple, geometrical properties of, 419 singular, 422 singular of first and second kind, 423 Cyclio collineations, 352 Decomposition of element into factors, 28 Degree of extension of finite degree, 105 of a polynomial, 20 Delta symbols used in determinant theory, 74 Dependence, algebraic, 108 linear, 43 Derivative of a polynomial, 119 of a rational function, 120 Desargues* Theorem, 191 assuming Pappus' Theorem, 272 converse, assuming the theorem, 196 general case, 194 in an incidence space, 213 Determinant, cofactor of element in, 77 effect of interchange of rows, 80 Laplace's expansion of, 75 of matrix product, 78 of skew matrix, 86 Determinants, 73 value of certain, 79 Diagonal points of a quadrangle, 244 Diagonals of a quadrangle, 244 Differentiation of algebraic functions, 128 of polynomials, 119 Dimension of extension over a field, 110 of a linear set, 45 Displacement on a line, 356 Distributive law for elements in a ring, 6 for multiplication of matrices, 52 Division algorithm, 24 Divisor of a polynomial, 25 of zero, right-hand, left-hand, 13 Dual correlation, 365 Grassmann coordinates, 292 projective transformation, 323 relationship between solutions of linear equations, 70 spaces, 188 theorems, 190 Duality, Principle of, 190 Elementary divisors of A-matrix, 92 Elementary transformations, 63 on A-matrices, 88 Envelope incidence-quadric, 427 Epsilon symbols used in determinant theory, 73 Equation of projectivity on a line, 281 Equations, homogeneous linear, 68 solution by determinants, 84 Equations in several unknowns, 159 non-homogeneous linear, 85 of a linear space, 188 Equivalence relation, 10 Equivalent extensions of a commutative field, 101 factorisations, 29 A-matrices, 88 (n+l)-tuples, 176 primes in a u.f.d., 30 rings, 10 solutions of homogeneous equations, 140
INDEX 435 Even permutation, 5 Exceptional points of a oorrelation, 362 of a singular projective transformation, 322 Exchange theorem for linear sets, 45 generalised form, 108 in projective space, 182 Extensions, equivalent, of a commutative field, 101 Extensions of a commutative field, 101 of finite degree, 105 of projective number space, 178 of projective space, 179 Factor of element in integral domain, 28 Factorisation, complete, 29 in an integral domain, 27 in polynomial rings, 33 of polynomials, 111 Factorisations, equivalent, 29 Field, 13 algebraic closure, 96, 112 algebraically closed, 96, 112 commutative, 13 example of non-commutative, 39 formed by points on a line, 258 ground, 41 Field obtained by adjunction of algebraic element, 103 of transcendental element, 104 Field of characteristic p, 37 of coefficients of set of equations, 146 Field of geometry algebraically closed, 274 commutative, 272 in an incidence space, 259 without characteristic, 273 Field of rational functions of #, 38 of*! xry 39 Field, root, of a polynomial, 114 Fields, examples of, 37 Finite algebraic basis, 108 Finite geometries, 205 Form, 140 binary, 148 Forms, canonical, see Canonical form Fundamental projectivities on a line, 278 Fundamental space corresponding to latent root of collineation, 336 to infinite latent root of correlation, 420 to zero latent root of correlation, 419 Fundamental space, rule for its determination, 336 spaces, their join, 336 theorem of algebra, 112 General correlation, 395 Generic linear space, 315 Geometrical loci connected with correlations, 421 properties of a simple correlation, 419 Grassmann coordinates, 288 basis theorem, 315 determine basis for linear space, 300 dual, 293 elementary properties, 297 intersections and joins, 304 linear space with given, 312 of linear space, 289 of projected space, 308 quadratic ^-relations, 309 relation between, and dual, 294 three-term p-relations, 311 Ground field, 41 Group, 2 additive, 3 alternating, 5 commutative or Abelian, 3 inverse of element in, 4 non-commutative, 3 of operations on a matrix, 73 of projective transformations, 177 subgroup, 5 symmetric, 5 unity of, 4 zero of, 4 Harmonic conjugates, 224 special case, 229 Harmonically conjugate pairs, 230 Highest common divisor of two polynomials, 26 factor of elements of integral domain, 31 Hilbert's basis theorem, 142 zero-theorem, 165 Homogeneous equations, 140 linear equations, 68 Homographic ranges, 217 Homology, 356 involutory, 356 Ideal, 11 basis for, 142 defines equivalence relation in rings, 12 improper, 12 prime, 168
436 INDEX Ideal (cont.) resultant, 166 right-hand, left-hand, 11 unit, 12 Ideals associated with sets of equations, 140 Improper ideal, 12 Incidence, Propositions of, 185, 208 Incidence quadric of correlation, 427 Incidence space of n dimensions, 211 as a Pln(K), 272 as a Prn(K), 259 Independence, algebraic, 108 linear, 43 Independent indeterminates over a field, 105 Indeterminate, adjunction of, 19 Indeterminate coefficients, system of equations with, 146 Indeterminates, independent over a field, 105 Infinity, properties of, 280 Integral domain, 13 indeterminate over, 21 quotient field of, 15 Intermediary line in projectivity, 218 Intersection of linear spaces, 189 in an incidence space, 211 in Grassmann coordinates, 304 Invariant factors of A-matrix, 92 Invariant of null-polarity, 380 of polarity, 372 Inverse matrix, 55, 81 of element in a group, 4 Involution on a line, 352 Involutory homology, 356 particular collineation, 355 transformation, 352 Irreducible polynomial, 27 Isobaric, 155 Isomorphism between rings, 10 Isomorphism, skew, 175 Jacob ian, 133 matrix, 133 Jaoobi's theorem on determinant of submatrix of ad jugate matrix, 82 Join of fundamental spaces in a collineation, 336 Join of linear spaces, 189 in an incidence space, 211 in Grassmann coordinates, 306 Kronecker delta, 54 A-matrices, canonical forms for, 88, 92 elementary transformations on, 88 equivalence of, 88 A-matrix, 86 elementary divisors of, 92 invariant factors of, 92 Laplace's expansion of determinant, 75 Latent roots of a matrix, 95 Law of composition, 2 Leading term of polynomial, 142 Left-hand projective space or Pln(K), 185 Left-hand vector, 48 row of matrix as, 56 Line, united, containing only one united point, 350 meeting vertices of fundamental stars, 342 Linear complex, 381 dependence, 43 independence, 43 Linear dependence, in Prn(K)y 180 of one point on others in projective space, 182 of points in an incidence space, 208, 268 Linear homogeneous equations, 68 manifold of solutions, 69 right-hand, 69 solution by determinants, 84 Linear non-homogeneous equations, 85 Linear set, 41 basis for, 45 dimension of, 45 left-hand, 41 minimal basis for, 45 right-hand, 42 Linear space, 183, 208 basis for, 183 equations of, 188 generic, 315 with given Grassmann coordinates, 312 Linear spaces, intersection of, 189, 211 join of, 189, 211 Lying in, relation between linear spaces, 184 Manifold of solutions of linear homogeneous equations, 69 Matrices, 49 over a commutative field, 71 postmultiplication of, 51 premultiplication of, 51 product of, 51 ring of square, 7, 54
INDEX 437 Matrices (corU.) square, 54 sum of, 50 Matrix, adjugate, 81 characteristic, 95 characteristic function of, 95 column printed as row, 50 complement of, 72 compound, 291 determinant of, 74 determinant of product, 78 division into blocks, 53 Jacobian, 133 product, 51 rank, 65 regular, 55 ring, 54 rows as vectors, 60 singular, 72 skew-symmetric, 86 square root of, 96 submatrix, 52 symmetric, 86 transformations of, 58 transpose of, 66 transpose of product, 72 zero, 51 Minimal algebraic basis, 108 basis of linear set, 45 Minimum function of matrix, 96 Minus sign, multiplicative property, 10 Multiplication of matrices, associative law, 52 distributive law, 52 Multiplication of points on a line, 256 (n+ l)-tuple, 176 (n-f 1)-tuples, equivalent, 176 Negative of element in additive group, 4 Non-commutative field, 39 group, 3 Non-Desarguesian geometry, 213 Non-homogeneous equations, 162 linear equations, 85 Non-singular correlations, 426 projective transformations, 322 Non-special collineation, 336 Null-polarity, 366, 378 canonical form for, 380 invariant of, 380 Odd permutation, 5 One-to-one correspondence, 10 denned by perspectivity, 216 Order of square matrix, 362 Pn(K) or left-hand projective space of dimension n over K, 185 Prn(K) or right-hand projective space of dimension n over K, 178 Pappus' theorem, 203 consequences of, 274 Partial derivatives, 121 Particular collineations, 354 with only one fundamental space, 356 Particular involutory collineation, 355 Pencil of lines, 216 Permissible transformations for reduction of correlation, 399 Permutation, even, odd, 5 Perspective ranges, 216 Perspective, triangles in, 192 PNln(K) or left-hand projective number space of dimension n over K, 176 PN^iK) or right-hand projective number space of dimension n over K, 176 Point in projective space, 178 algebraic, 179 rational, 179 symbolic notation for, 181 transcendental, 179 Point triad, 245 Polarities, 367 Polarity, 366 canonical form, 368, 372 conjugate points in, 367 conjugate spaces in, 375 invariant of, 372 vertex of, 374 Polynomial, 20 characteristic, of algebraic element, 102 coefficients of, 20 completely reducible, 112 degree of, 20 derivative of, 119 divisor of, 25 irreducible, 27 leading term of, 142 primitive, 33 Polynomial rings, 19 Polynomial, root of, 114 root field of, 114 symmetric functions of roots of, 115 total degree of, 115 zero of, 114 Polynomials, in a non-commutative ring, 23
438 INDEX Polynomials (cont.) relatively prime, 26 specialisation of, 22 Taylor's theorem for, 122 Postmultiplication of matrices, 51 Premultiplication of matrices, 51 Prime or Sn_x, 211 Prime ideal, 168 element of integral domain, 28 Primes, equivalent, in integral domain, 30 star of, in projective space, 339 Primitive element of algebraic extension, 126 polynomial, 33 Principle of duality, 190 Product of matrices, 51 of projective transformations, 177 Projective correspondence between points and spaces, 326 Projective number space, 176 extension of, 178 Projective ranges, 217 Projective space of n dimensions, 178 point in, 178 Projective transformation, 322 canonical form, 323 dual, 323 exceptional points of, 322 non-singular, 322 rank of, 323 singular, 322 singular space of, 325 vertex of, 325 Projective transformations, 177, 322 group of, 177 Projectively identical linear spaces, 186 invariant constructions, 231, 238 Projectivities, fundamental, on a line, 278 Projectivity or chain of perspectivities, 217 equation of, on a line, 281 Proper divisor of a polynomial, 25 Properties of infinity, 280 Propositions of incidence, 185, 208 Quadrangle, 244 Quadrangular set of points, 245 Quadratic ^-relations for Grassmann coordinates, 309 alternative form, 312 Quadric, 375 envelope incidence, 427 incidence, of correlation, 427 Quaternions, 40 Quotient field of integral domain, 15 r-way spaces, 206 Range of points, 216 Ranges, homographically related, 217 in perspective, 216 projectively related, 217 Rank of a matrix, 65 column-rank, 66 connection with determinants in commutative case, 84 row-rank, 60 Rank of projective transformation, 323 Rational point of projective space, 179 Reference system in incidence space, 249 Regular element, or unit, of ring, 13 matrix, 55 Related coordinate systems, 283 ranges, 216 Relation of one linear space lying in another, 184 Relatively prime polynomials, 26 Remainder theorem, 25 Representation of incidence space as a Prn(K), 259 Restricted type of allowable transformation, 367 Restrictions on the geometry in an incidence space, 272 Resultant-forms, 166 Resultant ideal, 166 for set of binary equations, 156 of (n-f 1) homogeneous equations in (n+1) unknowns, 170 of two binary forms, 150 system for equations in several unknowns, 159 Resultant, properties of, 152 Right-hand linear homogeneous equations, 69 Right-hand vectors, 48, 65 columns of matrix as, 66 Ring, 6 characteristic of, 14 condition for field, 13 condition for integral domain, 13 not a u.f.d., 29 of coefficients of set of equations, 146 of square matrices, 54 polynomial, 19 regular element, or unit, of, 13 unit, or regular element of, 13 unity of, 12 without characteristic, 14
INDEX 439 Rings, 6 classification of, 12 equivalence between, 10 extension of, 11 isomorphic, 10 ' subrings, 11 Root of polynomial, 114 field of a polynomial, 114 Roots, latent, of a matrix, 95 Row-rank of a matrix, 60 Rows of matrix as vectors, 66 Rule for determining fundamental space of collineation, 336 vertex of united primes of dual collineation, 339 Scale coordinate systems, 283 Scale on a line, 253 Segre symbol for collineations, 350 for singular collineations, 361 Self-conjugate points in a polarity, 367 Sk in a null-polarity, 382 Set, linear, 41 Set of coordinates of point in projective space, 178 Signed minor or cofactor, 77 Similar collineations, 328 Simple algebraic extensions, 99 Simple correlations, 390 geometrical properties, 419 Simplex in an incidence space, 211 of reference in projective space, 182 Singular collineations, 358 Segre symbol, 361 Singular collineations in [3], 361 Segre symbol, 361 Singular correlations, 422 of first and second kind, 423 Singular matrix, 72 projective transformation, 322 space of a projective transformation, 325 Skew-isomorphism, 175 Skew-symmetric bilinear form, 366 canonical form, 378, 380 Skew-symmetric matrix, 86 Solution of linear homogeneous equations, 68 by determinants in the commutative case, 84 "Solutions of a set of equations, 139 Space, incidence, 211 linear, 183, 208 of points with no transform, 358 ' projective, 178 . projective number, 176 Special case of collineation, 336 of harmonic conjugacy, 229 Specialisation of polynomials, 22 Square matrices, 54 root of a matrix, 96 Star of primes, 339 vertex of, 339 Subgroup, 5 Submatrix, 52 Subring, 11 Symbolic notation for points in projective space, 181 Symmetric bilinear form, 366 canonical form, 368, 372 Symmetric functions of roots of a polynomial, 115 group, 5 integral function, 115 Symmetric matrix, 86 Symmetric matrix, lemma on rank, 403 System of equations with indeterminate coefficients, 146 System of reference in incidence space, 249 Taylor's theorem for polynomials, 122 Three-term relation for Grassmann coordinates, 311 Total degree of a polynomial, 115 Transcendental element of extension field, 102 extension of a field, 104 point of projective space, 179 Transform of set of vectors, 49 Transformations, allowable, 59 elementary, 63 of general correlation, 395 of matrix, 58 projective, 177, 322 Transpose of matrix, 66 determinant of, 74 Transpose of matrix product, 72 Transposition permutation, 6 Triangle, 191 Triangle triad, 245 Triangles in perspective, 192 Trivial decomposition into factors, 28 U.f.d. or unique factorisation domain, 29 Unit or regular element of ring, 13 Unit ideal, 12 United k-spaces of a collineation, 344 lines of [n], 348 points of a collineation, 334
440 INDEX Unity of a group, 4 of a matrix ring, 8, 64 of a ring, 12 Unknowns, 139 pf linear homogeneous equations, 68 u-resultant of system of equations, 171 Vectors, apolar, 69 given by columns of matrix, 66 given by rows of matrix, 66 right-hand, left-hand, 48 Vertex of a polarity, 374 of a projective transformation, 326 of a star of primes, 339 Weight of isobaric polynomial, 155 Zero, divisor of, 13 matrix, 51 of a group, 4 of a matrix ring, 7 of a polynomial, 114 -theorem, Hilbert's, 165 *