Skip to main content

Full text of "Projective Geometry Volume I"

See other formats


THE TEXT IS 

LIGHT IN 

THE BOOK 



CARNEGIE INSTITUTE 
OF TECHNOLOGY 

LIBRARY 




PRESENTED BY 

Dr, Lloyd L* Dines 



PEOJECTIVE GEOMETRY 



BY 

OSWALD VEBLEN 

PROFESSOR OP MATHEMATICS, PRINCETON UNIVERSITY 
AND 

JOHN WESLEY YOUNG 

PROFESSOR OF MATHEMATICS, UNIVERSITY OF KANSAS 



VOLUME I 



GINN AND COMPANY 

BOSTON * NEW YORK CHICAGO LONDON 



ENTERED AT STATIONERS' HALL 



COPYRIGHT, 1910, BY 
OSWALD VEBLEN AND JOHN WESLEY YOUNG 



ALL RIGHTS RESERVED 
810.9 



fltfrenteum 



GINN AND COMPANY PRO- 
PRIETORS BOSTON U.S.A. 



PREFACE 



Geometry, which had been for centuries the most perfect example 
of a deductive science, during the creative period of the nineteenth 
century outgrew its old logical forms. The most recent period has 
however brought a clearer understanding of the logical foundations 
of mathematics and thus has made it possible for the exposition of 
geometry to resume the purely deductive form. But the treatment 
in the books which have hitherto appeared makes the work of lay- 
ing the foundations seem so formidable as either to require for itself 
a separate treatise, or to be passed over without attention to more 
than the outlines. This is partly due to the fact that in giving the 
complete foundation for ordinary real or complex geometry, it is 
necessary to make a study of linear order and continuity, a study 
which is not only extremely delicate, but whose methods are those 
of the theory of functions of a real variable rather than of elemen- 
tary geometry. 

The present work, which is to consist of two volumes and is in- 
tended to be available as a text in courses offered in American uni- 
versities to upper-class and graduate students, seeks to avoid this 
difficulty by deferring the study of order and continuity to the sec- 
ond volume. The more elementary part of the subject rests on a 
very simple set of assumptions which characterize what may be 
called "general protective geometry." It will be found that the 
theorems selected on this basis of logical simplicity are also elemen- 
tary in the sense of being easily comprehended and often used. 

Even the limited space devoted in this volume to the foundations 
may seem a drawback from the pedagogical point of view of some 
mathematicians. To this we can only reply that, in our opinion, 
an adequate knowledge of geometry cannot be obtained without 
attention to the foundations. "We believe, moreover, that the 
abstract treatment is peculiarly desirable in protective geometry, 
because it is through the latter that the other geometric disciplines 
are most readily coordinated. Since it is more natural to derive 

iii 



iv PllEFAOE 

the geometrical disciplines associated with the names of Euclid, 
Descartes, Lobatchewsky, etc., from protective geometry than it 
is to derive projective geometry from one of them, it is natural to 
take the foundations of projective geometry as the foundations of 
all geometry. 

The deferring of linear order and continuity to the second vol- 
ume has necessitated the deferring of the discussion of the metric 
geometries characterized by certain subgroups of the general pro- 
jective group. Such elementary applications as the metric proper- 
ties of conies will therefore be found in the second volume. This 
will be a disadvantage if the present volume is to be used for a 
short course in which it is desired to include metric applications. 
But the arrangement of the material will make it possible, when 
the second volume is ready, to pass directly from Chapter VIII of 
the first volume to the study of order relations (which may them- 
selves be passed over without detailed discussion, if this is thought 
desirable), and thence to the development of Euclidean metric 
geometry. We think that much is to be gained pedagogioally as 
well as scientifically by maintaining the sharp distinction between 
the projective and the metric. 

1 The introduction of analytic methods on a purely synthetic basis 
in Chapter VI brings clearly to light the generality of the set of 
assumptions used in this volume. What we call " general projective 
geometry " is, analytically, the geometry associated with a general 
number field. All the theorems of this volume are valid, not alone 
in the ordinary real and the ordinary complex projective spaces, but 
also in the ordinary rational space and in the finite spaces. The 
bearing of this general theory once fully comprehended by the 
student, it is hoped that he will gain a vivid conception of the 
organic unity of mathematics, which recent developments of postu- 
lational methods have so greatly emphasized. 

The form of exposition throughout the book has been condi- 
tioned by the purpose of keeping to the fore such general ideas as 
group, configuration, linear dependence, the correspondence be- 
tween and the logical interchangeability of analytic and synthetic 
methods, etc. Between two methods of treatment we have chosen 
the more conventional in all cases where a new method did nol 
to have unquestionable advantages. We have tried also to 



Jb'JbC.U.FAU.U V 

avoid in general the introduction of new terminology. The use 
of the word on in connection with duality was suggested by Pro- 
fessor Frank Morley. 

We have included among the exercises many theorems which in 
a larger treatise would naturally have formed part of the text. 
The more important and difficult of these have been accompanied 
by references to other textbooks and to journals, which it is hoped 
will introduce the student to the literature in a natural way. There 
has been no systematic effort, however, to trace theorems to their 
original sources, so that the book may be justly criticized for not 
always giving due credit to geometers whose results have been 
used. 

Our cordial thanks are due to several of our colleagues and stu- 
dents who have given us help and suggestions. Dr. H. H. Mitchell 
has made all the drawings. The proof sheets have been read in whole 
or in part by Professors Birkhoff, Eisenhart, and Wedderburn, of 
Princeton University, and by Dr. R. L. Borger of the University 
of Illinois. Finally, we desire to express to Ginn and Company our 
sincere appreciation of the courtesies extended to us. 

It is expected that the second volume will appear during the 

coming year. 

O. VEBLEN 
J. W. YOUNG- 

August, 1910 



CONTENTS 



INTRODUCTION 

SECTION PAGE 

1. Undefined elements and unproved propositions 1 

2. Consistency, categoricalness, independence. Example of a mathematical 

science 2 

3. Ideal elements in geometry 7 

4. Consistency of the notion of points, lines, and plane at infinity 9 

5. Protective and metric geometry 12 

CHAPTER I 
THEOREMS OF ALIGNMENT AND THE PRINCIPLE OF DUALITY 

6. The assumptions of alignment 15 

7. The plane 17 

8. The first assumption of extension 18 

9. The three-space 20 

10. The remaining assumptions of extension for a space of three dimensions . 24 

11. The principle of duality 26 

12. The theorems of alignment for a space of n dimensions 29 

CHAPTER II 
PROJECTION, SECTION, PERSPECTIVITY. ELEMENTARY CONFIGURATIONS 

13. Projection, section, perspectivity 34" 

14. The complete n-point, etc 36 

15. Configurations 38 

16. The Desargues configuration 39 

17. Perspective tetrahedra 43 

18. The quadrangle-quadrilateral configuration 44 

19. The fundamental theorem on quadrangular sets 47 

20. Additional remarks concerning the Desargues configuration 51 

CHAPTER III 

PROJEOTIVITIES OF THE PRIMITIVE GEOMETRIC FORMS OF ONE, TWO, 
AND THREE DIMENSIONS 

21. The nine primitive geometric forms 55 

22. Perspectivity and projectivity 56 

23. The projectivity of one-dimensional primitive forms 59 

vii 



viii CONTENTS 

SECTION PAGE 

24. General theory of correspondence. Symbolic treatment ....... 64 

25. The notion of a group ..................... 06 

26. Groups of correspondences. Invariant elements and figures ..... 07 

27. Group properties of pro jectivi ties ................ 08 

28. Protective transformations of two-dimensional forms ........ 71 

29. Protective collineations of three-dimensional forms ......... 75 



CHAPTER IV 

HARMONIC CONSTRUCTIONS AND THE FUNDAMENTAL THEOREM OF 
PROJECTIVE GEOMETRY 

30. The projectivity of quadrangular sets .............. 70 

31. Harmonic sets ........................ 80 

32. Nets of rationality on a line .................. 84 

33. Nets of rationality in the plane ................. 86 

34. Nets of rationality in space ................... 80 

35. The fundamental theorem of projectivity ............. 03 

36. The configuration of Pappus. Mutually inscribed and circumscribed tri- 

angles .......................... 08 

37. Construction of projectivities on one-dimensional forms ....... 100 

38. Involutions ......................... 102 

39. Axis and center of homology .................. 103 

40. Types of collineations in the plane ............... 106 

CHAPTER V 
CONIC SECTIONS 

41. Definitions. Pascal's and Brianchon's theorems .......... 100 

42. Tangents. Points of contact .................. 112 

43. The tangents to a point conic form a line conic ........... 116 

44. The polar system of a conic ................. . 120 

45. Degenerate conies ...................... 120 

46. Desargues's theorem on conies ................. 127 

47. Pencils and ranges of conies. Order of contact .......... 128 

CHAPTER VI 
ALGEBRA OF POINTS AND ONJG-DIMENSIONAL COORDINATE SYSTEMS 

48. Addition of points ...................... 141 

49. Multiplication of points .................... 144 

50. The commutative law for multiplication ............. 148 

51. The inverse operations .................... 148 

52. The abstract concept of a number system. Isomorphism ....... 149 

53. Nonhomogeneous coordinates ................. 150 

54. The analytic expression for a projectivity in a one-dimensional primitive 

form ........................... 152 

55. Von Staudt's algebra of throws ................. 157 



CONTENTS ix 

SECTION PAGE 

56. The cross ratio 159 

57. Coordinates in a net of rationality on a line , 1C2 

58. Homogeneous coordinates on a line 163 

59. Protective correspondence "between the points of two different lines . . 166 



CHAPTEE VII 
COORDINATE SYSTEMS IN TWO- AND THEEE- DIMENSIONAL FORMS 

60. Nonhomogeneous coordinates in a plane 169 

61. Simultaneous point and line coordinates 171 

62. Condition that a point be on a line 172 

63. Homogeneous coordinates in the plane 174 

64. The line on two points. The point on two lines 180 

65. Pencils of points and lines. Projectivity 181 

66. The equation of a conic 185 

67. Linear transformations in a plane . * 187 

68. Collineations between two different planes 190 

69. Nonhomogeneous coordinates in space 190 

70. Homogeneous coordinates in space 194 

71. Linear transformations in space 199 

72. Finite spaces 201 

CHAPTER VIII 
PROJECTIVITIES IN ONJE- DIMENSIONAL FORMS 

73. Characteristic throw and cross ratio 205 

74. Projective projectivities 208 

75. Groups of projectivities on a line 209 

76. Projective transformations between conies 212 

77. Projectivities on a conic 217 

78. Involutions 221 

79. Involutions associated with a given projectivity 225 

80. Harmonic transformations 230 

81. Scale on a conic 231 

82. Parametric representation of a conic 234 

CHAPTER IX 
GEOMETRIC CONSTRUCTIONS. INVARIANTS 

83. The degree of a geometric problem 236 

84. The intersection of a given line with a given conic 240 

85. Improper elements. Proposition K 2 241 

86. Problems of the second degree 245 

87. Invariants of linear and quadratic binary forms 251 

88. Proposition K n 254 

89. Taylor's theorem. Polar forms 255 



X. CONTENTS 



90. Invariants and covariants of binary forms 257 

91. Ternary and quaternary forms and their invariants 258 

92. Proof of Proposition K n 200 

CHAPTER X 
PROJECTIVE TRANSFORMATIONS OF TWO-DIMENSIONAL FORMS 

93. Correlations between two-dimensional forms 202 

94. Analytic representation of a correlation between two planes 200 

95. General projective group. Representation by matrices 208 

96. Double points and double lines of a collineation in a piano 271 

97. Double pairs of a correlation 278 

98. Fundamental conic of a polarity in a plane 282 

99. Poles and polars with respect to a conic. Tangents 284 

100. Various definitions of conies 285 

101. Pairs of conies 287 

102. Problems of the third and fourth degrees 294 

CHAPTER XI 
FAMILIES OF LINES 

103. The regains 21)8 

104. The polar system of a regulus 300 

105. Projective conies 304 

106. Linear dependence of lines 1)11 

107. The linear congruence 312 

108. The linear complex 319 

109. The Pliicker line coordinates 327 

110. Linear families of lines 329 

111. Interpretation of line coordinates as point coordinates in S 6 331 

INDEX 335 



PROJECTIVE GEOMETRY 



INTRODUCTION 

1. Undefined elements and unproved propositions. Geometry deals 
with the properties of figures in space. Every such figure is made up 
of various elements (points, lines, curves, planes, surfaces, etc.), and 
these elements bear certain relations to each other (a point lies on a 
line, a line passes through a point, two planes intersect, etc.). The 
propositions stating these properties are logically interdependent, and 
it is the object of geometry to discover such propositions and to 
exhibit their logical interdependence. 

Some of the elements and relations, by virtue of their greater 
simplicity, are chosen as fundamental, and all other elements and 
relations are defined.in terms of them. Since any defined element or 
relation must be defined in terms of other elements and relations, 
it is necessary that one or more of the elements and one or more of 
the relations between them remain entirely undefined; otherwise a 
vicious circle is unavoidable. Likewise certain of the propositions 
are regarded as fundamental, in the sense that all other propositions 
are derivable, as logical consequences, from these fundamental ones. 
But here again it is a logical necessity that one or more of the prop- 
ositions remain entirely improved ; otherwise a vicious circle is again 
inevitable. 

The starting point of any strictly logical treatment of geometry 
(and indeed of any "branch of mathematics) must then le a set of un- 
defined elements and relations, and a set of unproved propositions 
involving them ; and from these all other propositions (theorems) are 
to le derived ly the methods of formal logic. Moreover, since we 
assumed the point of view of formal (Le. symbolic) logic, the unde- 
fined elements are to be regarded as mere symbols devoid of content, 
except as implied by the fundamental propositions. Since it is mani- 
festly absurd to speak of a proposition involving these symbols as 

1 



2 INTEODUCTION [INTROD. 

self-evident, the unproved propositions referred to above must be re- 
garded as mere assumptions. It is customary to refer to these funda- 
mental propositions as axioms or postulates, but we prefer to retain the 
term assiwiption as more expressive 'of their real logical character. 

We understand the term a mathematical science to mean any set 
of propositions arranged according to a sequence of logical deduction. 
From the point of view developed above such a science is purely 
abstract. If any concrete system of things may be regarded as sat- 
isfying the fundamental assumptions, this system is a concrete ap- 
plication or representation of the abstract science. The practical 
importance or triviality of such a science depends simply on the 
importance or triviality of its possible applications. These ideas will 
be illustrated and further discussed in the next section, where it will 
appear that an abstract treatment has many advantages quite apart 
from that of logical rigor. 

2. Consistency; categoricalness, independence. Example of a math- 
ematical science. The notion of a class * of objects is fundamental 
in logic and therefore in any mathematical science. The objects 
which make up the class are called the elements of the class. The 
notion of a class, moreover, and the relation of belonging to a class 
(being included in a class, being an element of a class, etc.) are primi- 
tive notions of logic, the meaning of which is not here called in 
question.f 

The developments of the preceding section may now be illustrated 
and other important conceptions introduced by considering a simple 
example of a mathematical science. To this end let S be a class, the 
elements of which we will denote by A 9 B,C,... Further, let there 
be certain undefined subclasses $ of S, any one of which we will call 
an m-class. Concerning the elements of S and the m-classes we now 
make the following 

ASSUMPTIONS : 

I. If A and J3 are distinct elements of S, there is at least one 
m-class containing both A and B. 

* Synonyms for class are set, aggregate, assemblage, totality; in German, Menge; 
inPrench, ensemble. 

1 Of. B. Russell, The Principles of Mathematics, Cambridge, 1003 ; and L, Cou- 
turat, Les Principes des mathe'matiques, Paris, 1905. 

$ A class S' is said to be a subclass of another class S, if every element of S' is 
an element of S. 



2] A MATHEMATICAL SCIENCE 3 

II. If A and B are distinct elements of S, there is not more than 
one m-class containing loth A and B. 

III. Any two m-classes have at least one element of S in common. 

IV. There exists at least one m-class. ' 

V. JEJvery m-class contains at least three elements of S. 
VI. All the elements of S do not belong to the same m-class, 

VII. No m-class contains more than three elements of S. 

The reader will observe that in this set of assumptions we have 
just two undefined terms, viz., element of S and m-class, and one 
undefined relation, belonging to a class. The undefined terms, more- 
over, are entirely devoid of content except such as is implied in the 
assumptions. 

Now the first question to ask regarding a set of assumptions is : 
Are they logically consistent? In the example above, of a set of 
assumptions, the reader will find that the assumptions are all true 
statements, if the class S is interpreted to mean the digits 0, 1, 2, 3, 
4, 5, 6 and the m-classes to mean the columns in the following table : 

0123456 

(1) 1234560 

3456012 

This interpretation is a concrete representation of our assumptions. 
Every proposition derived from the assumptions must be true of this 
system of triples. Hence none of the assumptions can be logically 
inconsistent with the rest ; otherwise contradictory statements would 
be true of this system of triples. 

Thus, in general, a set of assumptions is said to le consistent if a 
single concrete representation of the assumptions can le given* 

Knowing our assumptions to be consistent, we may proceed to de- 
rive some of the theorems of the mathematical science of which they 
are the basis : 

Any two distinct elements of S determine one and only one m-class 
containing "both these elements (Assumptions I, II). 

* It will be noted that this test for the consistency of a set of assumptions 
merely shifts the difficulty from one <do main to another. It is, however, at present 
the only test known. On the question as to the possibility of an absolute test of 
consistency, cf. Hilbert, Grundlagen der Geometrie, 2d ed., Leipzig (1903), p. 18, and 
Verhandlungen d. III. intern, math* Kongresses zu Heidelberg, Leipzig (1904), 
p. 174; Padoa, L'Enseignement niathe'matique, Vol. V (1903), p. 85. 



4 INTEODUCTJOK [INTBOD. 

The m-class- containing the elements A and B may conveniently 
be denoted by the symbol AB. 

Any two m-classes have one and only one dement of S in common 
(Assumptions II, III). 

Tliere exist three elements of S which are not all in the same 
m-class (Assumptions IV, V, VI). 

In accordance with the last theorem, let A, B, C be three elements 
of S not in the same m-class. By Assumption V there must be a 
third element in each of the m-classes AB, BC, CA, and by Assump- 
tion II these elements must be distinct from each other and from 
A, By and C. Let the new elements be D, E, G, so that each of 
the triples ABD, BCE, CAG belongs to the same m-class. By 
Assumption III the m-classes AE and BG, which are distinct from 
aH the m-classes thus far obtained, have an element of S ia common, 
which, by Assumption II, is distinct from those hitherto mentioned ; 
let it be denoted by F, so that each of the triples AEF and BIG 
belong to the same m-class. No use has as yet been made of As- 
sumption VII. We have, then, the theorem : 

Any class S subject to Assumptions / VI contains at least seven 
elements. 

Now, making use of Assumption VII, we find that the m-classes 
thus far obtained contain only the elements mentioned. The m-classes 
CD and AEF have an element in common (by Assumption III) 
which cannot be A or E, and must therefore (by Assumption VII) 
be F. Similarly, AOG and the m-class DE have the element G in 
common. The seven elements A, B, C, I), E, F, G have now been 
arranged into m-classes according to the table 

A B C D E F G 

(!') B C D E F G A 

D E F G A B G 

in which the columns denote m-classes. The reader may note at once 
that this table is, except for the substitution of letters for digits, 
entirely equivalent to Table (1); indeed (I/) is obtained from (1) by 
replacing by A, 1 by B, 2 by (7, etc. We can show, furthermore, 
that S can contain no other elements than A, B, <7, D, E, F, G. For 
suppose there were another element, T. Then, by Assumption III, 



2] CATEGORICALKESS 5 

the m-classes TA and BFG would have an element in common. This 
element cannot be B> for then ABTD would belong to the same 
w-class ; it cannot be F, for then AFTJE would all belong to the same 
w-class; and it cannot be G, for then AGTC would all belong to the 
same m-class. These three possibilities all contradict Assumption VII. 
Hence the existence of T would imply the existence of four elements 
in the wi-class BFG, which is likewise contrary to Assumption VII. 

The properties of the class S and its m-classes may also be repre- 
sented vividly by the accompanying figure (fig. 1). Here we have 
represented the elements of S *by 
points (or spots) in a plane, and 
have joined by a line every triple 
of these points which form an m- 
class. It is seen that the points 
may be so chosen that all but one 
of these lines is a straight line. 
This suggests at once a similarity 
to ordinary plane geometry. Sup- 
pose we interpret the elements of 
S to be the points of a plane, and interpret the m-classes to be the 
straight lines of the plane, and let us reread our assumptions with this 
interpretation. Assumption VII is false, but all the others are true 
with the exception of Assumption III, which is also true except when 
the lines are parallel. How this exception can be removed we will 
discuss in the next section, so that we may also regard the ordinary 
plane geometry as a representation of Assumptions I- VI. 

Returning to our miniature mathematical science of triples, we are 
now in a position to answer another important question : To what eoo- 
tent do Assumptions I VII characterize the class S and the m-classes ? 
We have just seen that any class S satisfying these assumptions may 
be represented by Table (I/) merely by properly labeling the ele- 
ments of S. In other words, if S x and S 2 are two classes S subject 
to these assumptions, every element of S x may be made to correspond * 
to a unique element of S 2 , in such a way that every element of S 2 
is the correspondent of a unique element of S y and that to every 
w-class of Sj there corresponds an m-class of S r The two classes are 

* The notion of correspondence is another primitive notion which we take over 
without discussion from the general logic of classes. 




6 INTBODUCTION" [INTROB. 

then said to be in one-to-one reciprocal correspondence, or to be simply 
isomorphic* Two classes S are then abstractly equivalent ; i.e. there 
exists essentially only one class S satisfying Assumptions I- VII. 
This leads to the following fundamental notion : 

A set of assumptions is said to "be categorical, if there is essentially 
only one system for which the assumptions are 'valid ; i.e. if any, two 
such systems may le made simply isomorphic. 

We have just seen that the set of Assumptions I-VII is categor- 
ical. If, however, Assumption VII be omitted, the remaining set of 
six assumptions is not categorical. We have already observed the 
possibility of satisfying Assumptions I- VI by ordinary plane geom- 
try. Since Assumption III, however, occupies as yet a doubtful posi- 
tion in this interpretation, we give another, which, by virtue of its 
simplicity, is peculiarly adapted to make clear the distinction between 
categorical and noncategorical. The reader will find, namely, that 
each of the first six assumptions is satisfied by interpreting the class S 
to consist of the digits 0, 1, 2, - *, 12, arranged according to the fol- 
lowing table of m-classes, every column constituting one m-class : 

0123456 789 10 11 12 

9 1234567891011120 

^ 3 4 5 6 7 8 9 10 11 12 1 2 

9 10 11 12 012 345 6 78 

Henee Assumptions I-VI are not sufficient to characterize completely 
the class S, for it is evident that Systems (1) and (2) cannot be made 
isomorphic. On the other hand, it should be noted that all theorems 
derivable from Assumptions I-VI are valid for both (1) and (2). 
These two systems are two essentially different concrete representa- 
tions of the same mathematical science. 

This brings us to a third question regarding our assumptions : Are 
they independent ? That is, can any one of them be derived as a log- 
ical consequence of the others ? Table (2) is an example which shows 
that Assumption VII is independent of the others, because it shows 
that they can all be true of a system in which Assumption VII is 
false. Again, if the class S is taken-to mean the three letters A, B, (7, 

* The isomorphism of Systems (1) and (1") is clearly exhibited in fig, 1, where 
each point is labeled both with a digit and with a letter. This isomorphism may, 
moreover, be established in 7-64 different ways. 



2,3] - IDEAL ELEMENTS 7 

and the m-classes to consist of the pairs AB, JBC, CA, then it is 
clear that Assumptions I, II, III, IV, VI, VII are true of this class 
S, and therefore that any logical consequence of them is true with 
this interpretation. Assumption V, however, is false for this class, 
and cannot, therefore, be a logical consequence of the other assump- 
tions. In like manner, other examples can be constructed to show 
that each of the Assumptions I- VII is independent of the remain- 
ing ones. 

3. Ideal elements in geometry. The miniature mathematical science 
which we have just been studying suggests what we must do on a 
larger scale in a geometry which describes our ordinary space. We 
must first choose a set of undefined elements and a set of funda- 
mental assumptions. This choice is in no way prescribed a priori, 
but, on the contrary, is very arbitrary. It is necessary only that the 
undefined symbols be such that all other elements and relations that 
occur are definable in terms of them ; and the fundamental assump- 
tions must satisfy the prime requirement of logical consistency, and 
be such that all other propositions are derivable from them by formal 
logic. It is desirable, further, that the assumptions be independent* 
and that certain sets of assumptions be categorical. There is, further, 
the desideratum of utmost symmetry and generality in the whole 
body of theorems. The latter means that the applicability of a theo- 
rem shall be as wide as possible. This has relation to the arrange- 
ment of the assumptions, and can be attained by using in the proof 
of each theorem a minimum of assumptions-! 

Symmetry can frequently be obtained by a judicious choice of 
terminology. This is well illustrated by the concept of "points at 
infinity" which is fundamental in any treatment of projective geome- 
try. Let us note first the reciprocal character of the relation expressed 
by the two statements : 

A point lies on a line. A line passes through a point. 

To exhibit clearly this reciprocal character, we agree to use the phrases 
A point is on a line ; A line is on a point 

'* This is obviously necessary for the precise distinction between an assumption 
and a theorem. 

t If the set of assumptions used in the proof of a theorem is not categorical, the 
applicability of the theorem is evidently -wider than in the contrary case. Cf . exam- 
ple of preceding section* 



g INTRODUCTION [INTROD. 

to express this relation. Let us now consider the following two 
propositions : 

1. Any two distinct points of 1'. Any two distinct lines of a 

a plane are on one and only one plane are on one and only one 

Urn* P int - 

Either of these propositions is obtained from the other by simply 
interchanging the words point and line. The first of these propositions 
we recognize as true without exception in the ordinary Euclidean 
geometry. The second, however, has an exception when the two 
lines are parallel. In view of the symmetry of these two propositions 
it would clearly add much to the symmetry and generality of all 
propositions derivable from these two, if we could regard them both 
as true without exception. This can be accomplished by attributing 
to two parallel lines apoint of intersection. Such a point is not, 
of course, a point in the ordinary sense ; it is to be regarded as an 
ideal point, which we suppose two parallel lines to have in common. 
Its introduction amounts merely to a change in the ordinary termi- 
nology. Such an ideal point we call a point at infinity ; and we 
suppose one such point to exist on every line.f 

The use of this new term leads to a change in the statement, 
though not in the meaning, of many familiar propositions, and makes 
us modify the way in which we think of points, lines, etc. Two non- 
parallel lines cannot have in common a point at infinity without 
doing violence to propositions 1 and 1'; and since each of them has a 
point at infinity, there must be at least two such points. Proposition 
1, then, requires that we attach a meaning to the notion of a line on 
two points at infinity. Such a line we call a line at infinity, and 
think of it as consisting of all the points at infinity in a plane. 
In like manner, if we do not confine ourselves to the points of a 
single plane, it is found desirable to introduce the notion of a plane 
through three points at infinity which are not all on the same line 
at infinity. Such a plane we call a plane at infinity, and we think 

* By line throughout we mean straight line. 

f It should be noted that (since we are taking the point of view of Euclid) we do 
not think of a line as containing more than one point at infinity ; for the supposi- 
tion that a line contains two such points would imply either that two parallels can 
be drawn through a given point to a given line, or that two distinct lines can have 
more than one point in common. 



3,4] CONSISTENCY OF IDEAL ELEMENTS 9 

of it as consisting of all the points at infinity in space. Every ordi- 
nary plane is supposed to contain just one line at infinity ; every sys- 
tem of parallel planes in space is supposed to have a line at infinity 
in common with the plane at infinity, etc. 

The fact that we have difficulty in presenting to our imagination 
the notions of a point at infinity on a line, the line at infinity in a 
plane, and the plane at infinity in space, need not disturb us in this 
connection, provided we can satisfy ourselves that the new terminol- 
ogy is self-consistent and cannot lead to contradictions. The latter 
condition amounts, in the treatment that follows, simply to the con- 
dition that the assumptions on which we build the subsequent theory 
be consistent. That they are consistent will be shown at the time 
they are introduced. The use of the new terminology may, however, 
be justified on the basis of ordinary analytic geometry. This we 
do in the next section, the developments of which will, moreover, 
be used frequently in the sequel for proving the consistency of the 
assumptions there made. 

4. Consistency of the notion of points, lines, and plane at infinity. 
We will now reduce the question of the consistency of our new ter- 
minology to that of the consistency of an algebraic system. For this 
purpose we presuppose a knowledge of the elements of analytic geom- 
etry of three dimensions.* In this geometry a point is equivalent 
to a set of three numbers (x, y, z). The totality of all such sets of 
numbers constitute the analytic space of three dimensions. If the 
'numbers are all real numbers, we are dealing with the ordinary "real" 
space ; if they are any complex numbers, we are dealing with the ordi- 
nary " complex " space of three dimensions. The following discussion 
applies to either case. 

A plane is the set of all points (number triads) which satisfy a 
single linear equation 

ax + ly + cz + d = 0. 

A Ime is the set of all points which satisfy two linear equations, 

a^x + \y + CjZ + d^ 0, 
0, 



* Such knowledge is not presupposed elsewhere in this hook, except in the case 
of consistency proofs. The elements of analytic geometry are indeed developed 
from the beginning (cf. Chaps. VI, VH). 



10 INTRODUCTION [INTEOD. 

provided the relations 



do not hold.* 

Now the points (x, y, z), with the exception of (0, 0, 0), may also be 
denoted by the direction cosines of the line joining the point to the 
origin of coordinates and the distance of the point from the origin ; 

say by f 

I, m,n, -)> 
dj 

where d = V^+^+r, and I = -> m = --> n = ^ The origin itself 

d d cL 

may be denoted by (0, 0, 0, k), where i is arbitrary. Moreover, any 
four numbers (x v aj a , 8 , a 4 ) (x = 0), proportional respectively to 

I, m, n, - )> will serve equally well to represent the point (x, y, ), 



provided we agree that (x v x 2 , x s , x 4 ) and (cx v cx^ c% s , cx 4 ) represent 
the same point for all values of c different from 0. For a point 
(x 9 y, z) determines 

ex 7 c?/ 

J cmi, 




where e is arbitrary (c = 0), and (x v x v x v a; 4 ) determines 

/-t \ ^1 ^o ^o 

(1) # = -^y=:->;s:=-3> 

aj 4 a? 4 as, 

provided a? 4 = 0. 

We have not assigned a meaning to (x v # 2 , x s , x 4 ) when ^ 4 = 0, but 

it is evident that if the point ( d, cm, en, - ) moves away from the 

\ *v 

origin an unlimited distance on the line whose direction cosines are 
/, m, n, its coordinates approach (cl, cm, en, 0). A little consideration 
will show that as a point moves on any other line with direction 

* It should be noted that we are not yet, in this section, supposing anything 
known regarding points, lines, etc., at infinity, but are placing ourselves on the 
"basis of elementary geometry. 



4] CONSISTENCY OF IDEAL ELEMENTS 11 

cosines Z, m, n, so that its distance from the origin increases indefi- 
nitely, its coordinates also approach (cl, cm, en, 0). Furthermore, these 
values are approached, no matter in which of the two opposite direc- 
tions the point moves away from the origin. We now define (x v # 3 , 
# 8 , 0) as a point at infinity or an ideal point. We have thus associ- 
ated with every set of four numbers (x v x z , x s , x^ a point, ordinary 
or ideal, with the exception of the set (0, 0, 0, 0), which we exclude 
entirely from the discussion. The ordinary points are those for which 
x is not zero, and its ordinary Cartesian coordinates are given by the 
equations (1). The ideal points are those for which # 4 = 0. The num- 
bers (x v # 2 , # 3 , x) we call the homogeneous coordinates of the point. 
We now define a plane to be the set of all points (x v # 2 , # 8 , # 4 ) 
which satisfy a linear homogeneous equation : 

ax^+ bx% -f cx 9 + dx^ = 0. 

It is at once clear from the preceding discussion that as far as all 
ordinary points are concerned, this definition is equivalent to the one 
given at the beginning of this section. However, according to this 
definition all the ideal points constitute a plane # 4 = 0. This plane 
we call the plane at infinity. In like manner, we define a line to 
consist of all points (x v x 2 , X B> 4 ) which satisfy two distinct linear 
homogeneous equations : 

a^+ \x 2 + c&+ d&= 0, 
cx + d = 0. 



Since these expressions are to be distinct, the corresponding coefficients 
throughout must not be proportional. According to this definition 
the points common to any plane (not the plane at infinity) and the 
plane # 4 = constitute a line. Such a line we call a line at infinity, 
and there is one such in every ordinary plane. Finally, the line de- 
fined above by two equations contains one and only one point with 
coordinates (x v x v x s , 0) ; that is, an ordinary line contains one and 
only one point at infinity. It is readily seen, moreover, that with 
the above definitions two parallel lines have their points at infinity 
in common. 

Our discussion has now led us to an analytic definition of what 
may be called, for the present, an analytic projectile space of three 
dimensions. It consists of : 



12 INTRODUCTION [INTROD. 

Points : All sets of four numbers (x v x# X B) # 4 ), except the set 
(0, 0, 0, 0), where (cx v cx^ cx z) cxj is regarded as identical with 
(x# # 2 , s g , # 4 ), provided c is not zero. 

Planes: All sets of points satisfying one linear homogeneous 
equation. 

lines: All sets of points satisfying two distinct linear homoge- 
neous equations. 

Such a protective space cannot involve contradictions unless our 
ordinary system of real or complex algebra is inconsistent. The defi- 
nitions here made of points, lines, and the plane at infinity are, 
however, precisely equivalent to the corresponding notions of the 
preceding section. We may therefore use these notions precisely in 
the same way that we consider ordinary points, lines, and planes. 
Indeed, the fact that no exceptional properties attach to our ideal 
elements follows at once from the symmetry of the analytic formu- 
lation; the coordinate # 4 , whose vanishing gives rise to the ideal 
points, occupies no exceptional position in the algebra of the homo- 
geneous equations. The ideal points, then, are not to be regarded 
as different from the ordinary points. 

All the assumptions we shall make in our treatment of projective 
geometry will be found to be satisfied by the above analytic creation, 
which therefore constitutes a proof of the consistency of the assump- 
tions in question. This the reader will verify later. 

5. Projective and metric geometry. In projective geometry no 
distinction is made between ordinary points and points at infinity, 
and it is evident by a reference forward that our assumptions pro- 
vide for no such distinction. We proceed to explain this a little 
more fully, and will at the same time indicate in a general way 
the difference between projective and the ordinary Euclidean metric 
geometry. 

Confining ourselves first to the plane, let m and m r be two distinct 
lines, and P a point not on either of the two lines. Then the points 
of m may be made to correspond to the points of m/ as follows : To 
every point A on m let correspond that point A r on m f in which m f 
meets the line joining A to P (fig. 2). In this way every point on 
either line is assigned a unique corresponding point on the other 
line. This type of correspondence is called perspective, and the points 
on one line are said to be transformed into the points of the other by 



4, 5] 



PEOJECTIYE AND METRIC GEOMETBY 



13 



a perspective transformation with center P. If the points of a line m 
be transformed into the points of a line m 1 by a perspective transfor- 
mation with center P, and then the points of m 1 be transformed into the 
points of a third line m ff by a perspective transformation with a new 
center Q ; and if this be continued any finite number of times, ulti- 
mately the points of the line m will have been brought into corre- 
spondence with the points of a line m (n) , say, in such a way that every 
point of m corresponds to a unique point of m (n \ A correspondence 
obtained in this way is called projectile, and the points of m are said 




to have been transformed into the points of m (n) by a protective 
transformation. 

Similarly, in three-dimensional space, if lines are drawn joining 
every point of a plane figure to a fixed point P not in the plane TT 
of the figure, then the points in which this totality of lines meets 
another plane 7r f will form a new figure, such that to every point of 
TT will correspond a unique point of TT', and to every line of TT will 
correspond a unique line of TT'. We say that the figure in TT has been 
transformed into the figure in TT' by & perspective transformation with 
center P. If a plane figure be subjected to a succession of such per- 
spective transformations with different centers, the final figure will 
still be such that its points and lines correspond uniquely to the 
points and lines of the original figure. Such a transformation is again 
called a projectile transformation. In projective geometry two figures 
that may be made to correspond to each other by means of a projec- 
tive transformation are not regarded as different. In other words, 



14 INTRODUCTION [INTROD. 

protective geometry is concerned with those properties of figures that 
are left unchanged when the figures are subjected to a projective 
transformation. 

It is evident that no properties that involve essentially the notion 
of measurement can have any place in projective geometry as such ;* 
hence the term projective, to distinguish it from the ordinary geom- 
etry, which is almost exclusively concerned with properties involving 
the idea of measurement. In case of a plane figure, a perspective 
transformation is clearly equivalent to the change brought about in 
the aspect of a figure by looking at it from a different angle, the 
observer's eye being the center of the perspective transformation. 
The properties of the aspect of a figure that remain unaltered when 
the observer changes his position will then be properties with which 
projective geometry concerns itself. For this reason von Staudt called 
this science Geometric der Lage. 

In regard to the points and lines at infinity, we can now see why 
they cannot be treated as in any way different from the ordinary 
points and lines of a figure. For, in the example given of a per- 
spective transformation between lines, it is clear that to the point at 
infinity on m corresponds in general an ordinary point on m ! 9 and 
conversely. And in the example given of a perspective transforma- 
tion between planes we see that to the line at infinity in one plane 
corresponds in general an ordinary line in the other. In projective 
geometry, then, there can be no distinction between the ordinary 
and the ideal elements of space. 

* The theorems of metric geometry may however be regarded as special cases 
of projective theorems. 



CHAPTER I 

THEOREMS OF ALIGNMENT AND THE PRINCIPLE OF DUALITY 

6. The assumptions of alignment. In the following treatment of 
protective geometry we have chosen the point and the line as unde- 
fined elements. We consider a class (cf. 2, p. 2) the elements of 
which we call points, and certain undefined classes of points which 
we call lines. Here the words point and line are to be regarded 
as mere symbols devoid of all content except as implied in the as- 
sumptions (presently to be made) concerning them, and which may 
represent any elements for which the latter may le valid propositions. 
In other words, these elements are not to be considered as having 
properties in common with the points and lines of ordinary Euclidean 
geometry, except in so far as such properties are formal logical conse- 
quences of explicitly stated assumptions. 

We shall in the future generally use the capital letters of the 
alphabet, as A, B, C, P, etc., as names for points, and the small let- 
ters, as a, &, c, Z, etc., as names for lines. If A and B denote the same 
point, this will be expressed by the relation A = B ; if they repre- 
sent distinct points, by the relation A^B. If A = B, it is sometimes 
said that A coincides with B, or that A is coincident with B. The 
same remarks apply to two lines, or indeed to any two elements of 
the same kind. 

All the relations used are defined in general logical terms, mainly 
by means of the relation of 'belonging to a class and the notion of one- 
to-one correspondence. In case a point is an element of one of the 
classes of points which we call lines, we shall express this relation 
by any one of the phrases : the point is on or lies on or is a point of 
the line, or is united with the line ; the line passes through or con- 
tains or is united with the point. We shall often find it convenient 
to use also the phrase the line is on the point to express this relation. 
Indeed, all the assumptions and theorems in this chapter will be 
stated consistently in this way. The reader will quickly become ac- 
customed to this " on " language, which is introduced with the purpose 

15 



16 THEOREMS OF ALIGNMENT AND DUALITY [CHAJP.I 

of exhibiting in its most elegant form one of the most far-reaching 
theorems of projective geometry (Theorem 11). Two lines which have 
a point in common are said to intersect in or to meet in that point, or 
to le on a common point. Also, if two distinct points lie on the same 
line, the line is said to join the points. Points which are on the 
same line are said to be collinear ; points which are not on the same 
line are said to be noncollinear. Lines which are on the same point 
(i.e. contain the same point) are said to be copunctal, or concurrent* 
Concerning points and lines we now make the following assump- 
tions : 

THE ASSUMPTIONS OF ALIGNMENT, A : 

A 1. If A and B are distinct points, there is at least one line on 
both A and B. 

A 2. If A and B are distinct points, there is not more than one 
line on both A and B. 

A3. If A, B, are points not all on the same line, and D and 
E (D = E] are points such that B, C, D are on a line and C, A, E 

are on a line, there is a point F 
such that A, B, F are on a line 
and also D, E, F arc on a line 
(fig. 3).f 

It should be noted that this set 
of assumptions is satisfied by the 
triple system (1), p. 3, and also 
by the system of quadruples (2), 

p. 6, as well as by the points and lines of ordinary Euclidean geom- 
etry with the notion of " points at infinity " (cf. 3, p. 8), and by 

* The object of this paragraph is simply to define the terms in. common use in 
terms of the general logical notion of belonging to a class. In later portions of 
this book we may omit the explicit definition of such common terms when such 
definition is obvious. 

f The figures are to be regarded as a concrete representation of our science, in 
which the undefined " points" and "lines" of the science are represented by 
points and lines of ordinary Euclidean geometry (this requires the notion of ideal 
points; cf. 3, p. 8). Their function is not merely to exhibit one of the many 
possible concrete representations, but also to help keep in mind the various rela- 
tions in question. In using them, however, great care must be exercised not 
to use any properties of such figures that are not formal logical consequences 
of the assumptions ; in other words, care must be taken that all deductions are 
made formally from the assumptions and theorems previously derived from the 
assumptions* * 




7] THE PLANE 17 

the " analytic projective space " described in 4. Any one of these 
representations shows that our set of Assumptions A is consistent* 

The following three theorems are immediate consequences of the 
first two assumptions. 

THEOREM 1. Two distinct points are on one and only one line. 
(Al,A2)f 

The line determined by the points A, B (A = B) will often be 
denoted by the symbol or name AB. 

THEOREM 2. If C and D (C = D) are points on the line AB, A and 
B are points on the line CD. (Al, A 2) 

THEOREM 3. Two distinct lines cannot be on more than one common 
point. (A 1, A 2) 

Assumption A3 will be used in the derivation of the next theo- 
rem. It may be noted that under Assumptions A 1, A 2 it may be 
stated more conveniently as follows : If A 9 B, C^are ^points not jtll^on 
the samej-ine, the line joining any point -Z> on the. line BO to _any 
point JEJ (D = E) on the line CA meets the line AB in a point F. 
This is the form in which this assumption" is~ generally "used in the 
sequel. 

7. The plane. DEFINITION. If P t Q, E are three points not on 
the same line, and I is a line joining Q and R, the class S 2 of all 
points such that every point of S 2 is collinear with P and some 
point of I is called the plane determined by P and I. 

We shall use the small letters of the Greek alphabet, a, /3, y, TT, etc., 
as names for planes. It follows at once from the definition that P and 
every point of I are points of the plane determined by P and I. 

THEOREM 4. If A and B are points on a plane TT, then every point 
on the line AB is on TT. (A) 

Proof. Let the plane TT under consideration be determined by the 
point P and the line I 

* In the multiplicity of the possible concrete representations is seen one of the 
great advantages of the formal treatment quite aside from that of logical rigor. It 
is clear that there is a great gain in generality as long as the fundamental assump- 
tions are not categorical (cf. p. 6). In the present treatment our assumptions are 
not made categorical until yery late. 

t The symbols placed in parentheses after a theorem indicate the assumptions 
needed in its proof. The symbol A will be used to denote the whole set of Assump- 
tions A 1, A 2, A3. 




FIG. 



18 THEOREMS OF ALIGNMENT AND DUALITY [CHAP. I 

1. If both A and B are on Z, or if the line AB contains P, the 
theorem is immediate. 

2. Suppose A is on I, B not on I, and AB does not contain P (fig. 4). 
Since B is a point of w, there is a point 2?' on I collinear with B and P. 

If (7 be any point on AB 9 the line 
joining C on -4J5 to P on .B1?' 
will have a point T in common 
with AB'=l (A 3). Hence is a 
point of TT. 

3. Suppose neither A nor .2? is 
on I and that -4J? does not con- 
tain P (fig. 5). Since A and .# are 
points of TT, there exist two points 
J 7 and ff on Z collinear with A, P and 5, P respectively. The line join- 
ing A on A'P to .# on PB 1 has a point Q in common with J?'.4' (A 3), 
Hence every point of the line AB AQ 
is a point of TT, by the preceding case. 
This completes the proof. 
If all the points of a line are points 
of a plane, the line is said to be a line of 
the plane, or to lie in or to le in or to 
le on the plane; the plane is said to 
pass through, or to contain the line, 
or we may also say the plane is on the 
line. Further, a point of a plane is said 
to le in or to lie in the plane, and the 
plane is on the point 

8. The first assumption of extension. The theorems of the pre- 
ceding section were stated and proved on the assumption (explicitly 
stated in each case) that the necessary points and lines exist. The 
assumptions of extension, M, insuring the existence of all the points 
which we consider, will be given presently. The first of these, how- 
ever, it is desirable to introduce at this point. 
AN ASSUMPTION OF EXTENSION : 
E 0. There are at least three points on every line. 
This assumption is needed in the proof of the following 
THEOREM 5. Any two lines on the same plane TT are on a common 
point. (A, E 0) 



I 




FIG. 5 



8] 



ASSUMPTION OF EXTENSION 



19 



Proof. Let the plane TT be determined by the point P and the line I, 
and let a and 6 be two distinct lines of TT. 

1. Suppose a coincides with I (fig. 6). If I contains P, any point 

B of 6 (E 0) is colHnear with P and 
some point of Z=a, which proves the 
theorem when I contains P. If I does 
not contain P } there exist on 5 two 
points A and I? not on I (E 0), and 
since they are points of TT, they are 
collinear with P and two points A 1 
and B ! of / respectively. The line 
joining A on A'P to .# on PB 1 has a 
point JS in common with A'B 1 (A 3) 
ine^ theplane 




PIG-. 6 



i.e. I = a and & have a point in common. Hence 
TT has a point in common with L 
2. Let a and 5 both be distinct 
from L (i) Let a contain P (fig. 7). 
The line joining P to any point 
B of 5 (E 0) has a point B r in com- 
mon with I (Case 1 of this proof). 
Also the lines a and 5 have points 
,4' and R respectively in common 
with I (Case 1). Now the line 
A!P = a contains the points A J of 
EB } and P of ^^ and hence has a point A in common with 




JFia. 7 




Hence 

in common 



= Z>. 

h as a point 
it]^ any line om 



gA P. (ii) Let neither a nor 
contain P (fig. 8). As before, 
& and 6 meet Z in two points Q 
and 2 respectively. Let B 1 be a 
point of I distinct from Q and E 
(E 0). The line PB r then meets 
a and 6 in two points A and .Z? 
respectively (Case 2, (i)). If 
J[ = B y the theorem is proved. If A = .#, the line 6 has the point 
E in common with Q^ r and the point B in common with B f A, and 
hence has a point in common with AQ = a (A3). 




20 THEOREMS OF ALIGNMENT AND DUALITY [CHAP, i 

THEOREM 6. Tlie plane a determined ly a line I and a point P is 
identical with the plane /3 determined ly a line m and a point Q, 
provided m and Q are on a. (A, E 0) 

Proof. Any point B of /3 is collinear with Q and a point A of m 
(fig. 9). A and Q are both points of a, and hence every point of the 

line AQ is a point of a (Theorem 4). 
Hence every point of /3 is a point 
of a. Conversely, let B be any point 
of a. The line BQ meets w in a 
point (Theorem 5). Hence every 
point of a is also a point of /J. 

COROLLARY. There is one and only 
one plane determined "by three non- 
collinear points, or ly a line and a 
point not on the line, or ly two inter- 
secting lines. (A, E 0) 
The data of the corollary are all equivalent by virtue of EO. We 
will denote by ABC the plane determined by the points A, B, C\ 
by aA the plane determined by the line a and the point A 9 etc. 

THEOREM 7. Two distinct planes which are on two common points 
A y B (A = 3) are on all the points of the line AB, and on no other com- 
, mon points. (A, EO) 

Proof. By Theorem 4 the line AS lies in each of the two planes, 
which proves the first part of the proposition. Suppose C 9 not on AB, 
were a point common to the two planes. Then the plane determined by 
,A, B, C would be identical with each of the given planes (Theorem 6), 
which contradicts the hypothesis that the planes are distinct. 

COROLLARY. Two distinct planes cannot "be on more than one com- 
mon line. (A, E 0) 

9. jnbiejhjee-space. DEFINITION. If P, Q, E, T are four points 
not in the same plane, and if TT is a plane containing Q, R, and T, 
the class S g of all points such that every point of S 3 is- collinear with P 
and some point of TT is called the space of three dimensions , or the 
three-space determined by P and TT. 

If a point belongs to a three-space or is a point of a three-space, it 
is said to le in or to lie in or to le on the three-space. If all the points 
of a line or plane are points of a three-space S 3 , the line or plane is said 



9] 



THE THREE-SPACE 



21 




FHK 10 



to lie in or to le in or to le on the S 3 . Also the three-space is said to 
le on the point, line, or plane. It is clear from the definition that P and 
every point of TT are points of the three-space determined by P and TT. 

THEOREM 8. If A and B are distinct points on a three-space S s , 
every point on the line AB is on S 3 . (A) 

Proof. Let S 3 be determined by 
a plane TT and a point P. 

1. If A and B are both in TT, the 
theorem is an immediate conse- 
quence of Theorem 4. 

2. If the line AB contains P, 
the theorem is obvious. 

3. Suppose A is in TT, B not in 
TT ; and AB does not contain P 
(fig. 10). There then exists a point 
S f (= A) of TT collinear with B 

and P (def.). The line joining any point M on AB to P on BB f has 
a point M f in common with 2^-4 (A 3). But M f is a point of w, since 
it is a point of AB f . Hence jYis a point of S 3 (del). 

4. Let neither A nor J5 lie in TT, and let AB not contain P (fig. 11). 
The lines PA and P# meet TT in 
two points A r and B r respectively. 
But the line joining A on A r P to 
B on P-B' has a point C in common 
with B r A r . C is a point of TT, which 
reduces the proof to Case 3. 

It may be noted that in this 
proof no use has been made of E 0. 

In discussing Case 4 we have 
proved incidentally, in connection 
with EO and 'Theorem 4, the fol- 
lowing corollary: 

COROLLARY 1. If S 3 is a three-space determined ly a point P and a 
plane TT, then TT and any line on S 3 but not on TT are on one and only 
one common point (A, E 0) 

COROLLARY 2. JEvery point on any plane determined by three non- 
eollinear points on a three-space S 8 is on S s . (A) 




FIG. 11 



22 THEOREMS OF ALIGNMENT AND DUALITY [CHAP. I 

Proof. As before, let the three-space be determined by TT and P, 
and let the three noncollinear points 'be A, B, 0. Every point of the 
line BC is a point of S s (Theorem 8), and every point of the plane 
ABC* is collinear with A and some point of BC. 

COROLLARY 3. If a three-space S 8 is determined ly a point P and 
a plane TT, then TT and any plane on S 3 distinct from TT are on one 
and only one common line. (A, E 0) 

Proof. Any plane contains at least three lines not passing through 
the same point (del, A 1). Two of these lines must meet TT in two 
distinct points, which are also 
points of the plane of the lines 
(Cor. 1). The result then follows 
from Theorem 7. 

THEOREM 9. If a plane a and 
a line a not on a are on the same 
threes-space S gJ then a and a are 
on one and only one common point. 
(A,EO) 

Proof. Let S 3 be determined by 
the plane TT and the point P. 

1. If a coincides wither, the theo- 
rem reduces to Cor. 1 of Theorems. 

2. If # is distinct from TT, it has 

a line I in common with TT (Theorem 8, Cor. 3). Let A be any point 
on a not on a (EO) (fig. 12), The plane aA, determined by A and a, 
meets TT in a line m & I (Theorem 8, Cor. 3). The lines I, m have 
a point B in common (Theorem 5). The line AB in aA meets a in 
a point Q (Theorem 5), which is on a, since AB is on a. That a 
and a have no other point in common follows from Theorem 4. 

COROLLARY 1. Any two distinct planes on a three-space are on one 
and only one common line. (A, E 0) 

The proof is similar to that of Theorem 8, Cor. 3, and is left as an 
exercise. 

COROLLARY 2. Conversely, if two planes are on a common line, there 
exists a three-space on loth. (A, E 0) 

* The proof can evidently be so worded as not to imply Theorem 6. 




FIG. 12 



9] THE THREE-SPACE 23 

Proof. If the planes a and ft are distinct and have a line I in 
common, any point P of /3 not on I "will determine with a a three- 
space containing I and P and hence containing ft (Theorem 8, Cor. 2). 

COROLLARY 3. Three planes on a three-space ivhich are not on a 
common line are on one and only one common point. (A, E 0) 

Proof. This follows without difficult} 7 * from the theorem and Cor. 1. 

Two planes are said to determine the line which they have in com- 
mon, and to intersect or meet in that line. Likewise if three planes 
have a point in common, they are said to intersect or meet in the point. 

COROLLARY 4. If a, y3, 7 are three distinct planes on the same S 3 
but not on the same line, and if a line I is on each of t'wo planes p, v 
which are on the lines /3<y and ya respectively, then it is on a plane \ 

which is on the line aft. (A, E 0) p , 

r Q 

Proof. By Cor. 3 the planes a, I ,r 




j3, 7 have a point P in common, /\| 7 ~~7 7T7 

so that the lines /3y, ya, a/3 all XT ^-9 -f- -i 

contain P. The line I, being com- / X v / s x / / 

mon to planes through @y and ya, 
must pass through P, and the 
lines I and a/3 therefore intersect 
in P and hence determine a plane 
\ (Theorem 6, Cor.). 

THEOREM 10. The three-space 
S 3 determined by a plane TT, and 

a point P is identical with the three-space S determined 1y a plane 
7r f and a point P', provided TT' and P f are on S 3 . (A, EO) 

Proof. Any point A of S^ (fig. 13) is collinear with P 1 and some 
point A f of TT'; but P f and A f are both points of S 3 and hence A is a 
point of S 3 (Theorem 8). Hence every point of S^ is a point of S s . 
Conversely, if A is any point of S 3 , the line AP f meets TT' in a point 
(Theorem 9). Hence every point of S s is also a point of S. 

COROLLARY. TJiere is one and only one three-space on four given 
points not on the same plane, or a plane and a line not on the plane, 
or two nonintersecting lines. (A, E 0) 

The last part of the corollary follows from the fact that two 
nonintersecting lines are equivalent to four points not in the same 
plane (EO). 



24 THEOREMS OF ALIGNMENT AND DUALITY [CHAP. I 

It is convenient to use tlie term coplanar to describe points in the 
same plane. And we shall use the term skew lines for lines that have 
no point in common. Four noncoplanar points or two skew lines 
are said to determine the three-space in which they lie. 

10. The remaining assumptions of extension for a space of three 
dimensions. In 8 we gave a first assumption of extension. We will 
now add the assumptions which insure the existence of a space of 
three dimensions, and will exclude from our consideration spaces of 
higher dimensionality. 

ASSUMPTIONS OF EXTENSION, E : 

E.l. There exists at least one line. 

E 2. All points are not on the same line. 

E 3. All points are not on the same plane. 

E 3'. If S 3 is a three-space, every point is on S 3 . 

The last may be called an assumption of closure.* 

The last assumption might be replaced by any one of several equiv- 
alent propositions, such as for example : 

Every set of five points lie on the same three-space ; or 

Any two distinct planes have a line in common. (Of. Cor. 2, Theo- 
rem 9) 

There is no logical difficulty, moreover, in replacing the assumption 
(E3 ; ) of closure given above by an assumption that all the points 
are not on the same three-space, and then to define a " four-space " 
in a manner entirely analogous to the definitions of the plane and 
to the three-space already given. And indeed a meaning can be given 
to the words point and line such that this last assumption is satisfied 
as well as those that precede it (excepting E3 ; of course). We 
could thus proceed step by step to define the notion of a linear 
space of any number of dimensions and derive the fundamental 
properties of alignment for such a space. But that is aside from our 
present purpose. The derivation of these properties for a four-space 
will furnish an excellent exercise, however, in the formal reasoning 
here emphasized (cf. Ex. 4, p. 25). The treatment for the ^dimensional 
case will be found in 12, p. 29. 

* The terms extension and closure in this connection were suggested by N. J. Lennes. 
It will be observed that the notation has been so chosen that Ei insures the exist- 
ence of a space of i dimensions, the line and the plane, being regarded as spaces of 
one and two dimensions respectively. 



10] ASSUMPTIONS OF EXTENSION 25 

The following corollaries of extension are readily derived from the 
assumptions just made. The proofs are left as exercises. 

COROLLARY 1. At least three coplanar lines are on every point. 
COROLLARY 2. At least three distinct planes are on every line. 
COROLLARY 3. All planes are not on the same line. 
COROLLARY 4. All planes are not on the same point. 
COROLLARY 5. If $ 3 is a three-space, every plane is on $ 8 . 

EXERCISES 

1. Prove that through a given point P not on either of two skew lines I 
and I' there is one and only one line meeting both the lines Z, V. 

2. Prove that any two lines, each of which meets three given skew lines, 
are skew to each other. 

_^ 3. Our assumptions do not as yet determine whether the number of points 
on a line is finite or infinite. Assuming that the number of points on one line 
is finite and equal to n + 1, prove that 

i. the number of points on every line is n + 1; 
ii. the number of points on every plane is n 2 + n + 1; 
iii. the number of points on every three-space is n 8 + n 2 + n + 1; 
*& iv. the number of lines on a three-space is (n 2 + 1) (n z + n + 1); 

v. the number of lines meeting any two skew lines on a three-space is 



*< vi. the number of lines on a point or on a plane is n 2 + n + 1. 
- -* 4. Using the definition below, prove the following theorems of alignment for 
a four-space on the basis of Assumptions A and E : 

DEFINITION. If P, Q, R, S, T are five points not on the same three-space, 
and S 3 is a three-space on Q, R, S, T, the class S 4 of a.11 points such that every 
point of S 4 is collinear with P and some point of S 3 is called the four-space 
determined by P and S 8 . 

i. If A and B are distinct points on a four-space, every point on the line A B 
is on the four-space. \ 

ii. Every line on a four-space which is not on a given three-space of the 
four-space has one and only one point in common with the three-space. 

iii. Every point on any plane determined by three noncollinear points on 
a four-space is on the four-space. 

iv. Every point on a three-space determined by four noncoplanar points 
of a four-space is on the four-space. 

v. Every plane of a four-space determined by a point P and a three-space 
S 3 has one and only one Jjne in common with S 3 , provided the plane is not on S 3 . 

vi. Every three-space on a four-space determined by a point P and a three- 
space S 3 has one and only one plane in common with S 3 , provided it does 
not coincide with S a . 



26 THEOREMS OF ALIGNMENT AND DUALITY [CHAP. I 

viL If a three-space S 3 and a plane a not on S 3 are on the same four-space, 
S 3 and a have one and only one line in common. 

viii. If a three-space S 3 and a line I not on S g are on the same four-space, 
S 3 and I have one and only one point in common. 

ix. Two planes on the same four-space hut not on the same three-space 
have one and only one point in common. 

x. Any two distinct three-spaces on the same four-space have one and only 
one plane in common. 

xi. If two three-spaces have a plane in common, they lie in the same four-space. 

xii. The four-space S 4 determined l>y a three-space S 3 and a point P is 
identical with the four-space determined by a three-space 83 and a point P', 
provided 83 and P' are on S 4 . 

5. On the assumption that a line contains n + 1 points, extend the results 
of Ex. 3 to a four-space. 

11. The principle of duality. It is in order to exhibit the theorem 
of duality as clearly as possible that we have introduced the syrn- 
metrical, if not always elegant, terminology : 

A point is on a line. A line is on a point. 

A point is on a plane. A plane is on a point. 

A line is on a plane. A plane is on a line. 

A point is on a three-space. A three-space is on a point. 

A line is on a three-space. A three-space is on a line. 

A plane is on a three-space. A three-space is on a plane. 

The theorem in question rests on the f olio wing observation : If any 
one of the preceding assumptions, theorems, or corollaries is expressed 
by means of this "on" terminology and then a new proposition is 
formed by simply interchanging the words point and plane, then 
this new proposition will be valid, i.e. will be a logical consequence 
of the Assumptions A and E. We give below, on the left, a complete 
list of the assumptions thus far made, expressed in the " on " termi- 
nology, and have placed on the right, opposite each, the corresponding 
proposition obtained by interchanging the words point and plane 
together with the reference to the place where the latter proposition 
occurs in the preceding sections : 

ASSUMPTIONS A 1, A 2. If u4and THEOREM 9, COR. 1. If a and /9 
are distinct points, there is one are distinct planes, there is one and 
and only one line on A and B. only one line on a and /3.* 

* By virtue of Assumption E 8' it is not necessary to impose the condition that the 
elements to be considered are in the same three-space. This observation should empha- 
size, however, that the assumption of closure is essential in the theorem to he nrovad 



ill] 



THE PKINCIPLE OF DUALITY 



27 



ASSUMPTION A 3. If A, B, Q are 
points not all on the same line, and 
D and JS (D 3* E] are points such 
that B, C 9 D are on a line and 9 
A, E are 011 a line, then there is a 
point F such that A, B, F are on a 
line and also D, fi, F are on a line. 

ASSUMPTION EO. There are at 
least three points on every line. 

ASSUMPTION El. There exists 
at least one line. 

ASSUMPTION E 2. All points are 
not on the same line. 

ASSUMPTION E 3. All points are 
not on the same plane. 

ASSUMPTION E3'. If S, is a 

o 

three-space, every point is on S 3 . 



THEOBEM 9, COR. 4. If a, ft, <y 
are planes not all on the same line, 
and ^ and v(p*pv) are planes such 
that , % //, are on a line and % a, v 
are on a line, then there is a plane X 
such that a, /3, X are on a line and 
also /, v, X are on a line. 

COR. 2, p. 25. There are at 
least three planes on every line. 

ASSUMPTION E 1. , There exists 
at least one line. 

COR. 3, p. 25. All planes are 
not on the same line. 

COB. 4, P- 25. All planes are 
not on the same point. 

COE. 5, p. 25. If S s is a three- 



space, every plane is on S s . 

In all these propositions it is to be noted that a line is a class 
of points whose properties are determined by the assumptions, while 
a plane is a class of points specified by a definition. This definition 
in the "on" language is given below on the left, together with a* 
definition obtained from it by the interchange of point and plane. 
Two statements in this relation to one another are referred to as 
(space) duals of one another. 



If P, Q, JR are points not on 
the same line, and I is a line on 
Q and It, the class S 2 of all 
points such that every point of 
S 2 is on a line with P and some 



If X, JJL, v are planes not on the 
same line, and I is a line on" ^ 
and v > the class B 2 of all planes 
such that every plane of B 2 is on 
a line with X and some plane on 
I is called the "bundle determined 
by X and I. 



point on I is called the plane 
determined by P and I 

Now it is evident that, since X, /A, v and I all pass through a point 0> 
the bundle determined by X and I is simply the class of all planes on- 
the point 0. In like manner, it is evident that the dual of the defini- 
tion of a three-space is simply a definition of the class of all planes on 
a three-space. Moreover, dual to the class of all planes on a line we 
have the class of all points on a line, ie. the line itself, and conversely. 



28 THEOBEMS OF ALIGNMENT AND DUALITY [CHAP. I 

With the aid of these observations we are now ready to establish 
the so-called principle of duality : 

THEOEEM 11. THE J^J^E^OF^DUALITY^FOR A SPACE OF THREE 
DIMENSIONS. Any proposition deducible from Assumptions A and E 
concerning points, lines, and planes of a three-space remains valid, if 
stated in the "ow-^Jcraj^fo^j^A^ the words "point" and "plane^ 
are intercJianqed. (A, E) 

Proof. Any proposition deducible from Assumptions A and E is 
obtained from the assumptions given above on the left by a certain 
sequence of formal logical inferences. Clearly the same sequence of 
logical inferences may be applied to the corresponding propositions 
given above on the right. They will, of course, refer to the class of 
all planes on a line when the original argument refers to the class of 
all points on a line, i.e. to a line, and to a bundle of planes when the 
original argument refers to a plane. The steps of the original argu- 
ment lead to a conclusion necessarily stated in terms of some or all 
of the twelve types of " on " statements enumerated at the beginning 
of this section. The derived argument leads in the same way to a 
conclusion which, whenever the original states that a point P is on a 
line l y says that a plane TT' is one of the class of planes on a line V, 
i.e. that TT' is on V\ or which, whenever the original argument states 
that a plane TT is on a point P, says that a bundle of planes on a 
point P f contains a plane -TT', Le. that P r is on TT'. Applying similar 
considerations to each of the twelve types of " on " statements in 
succession, we see that to each statement in the conclusion arrived 
at by the original argument corresponds a statement arrived at by 
the derived argument in which the words point and plane in the 
original statement have been simply interchanged. 

Any proposition obtained in accordance with the principle of dual- 
ity just proved is called the space dual of the original proposition. 
The point and plane are said to be dual elements ; the line is self- 
duaL We may derive from the above similar theorems on duality ^in 
a plane and at a point. For, consider a plane TT and a point P not on 
TT, together with all the lines joining P with every point of TT. Then 
to every point of TT will correspond a line through P, and to every 
line of TT will correspond a plane through P. Hence every proposi- 
tion concerning the points and lines of TT is also valid for the corre- 
sponding lines and planes through P. The space dual of the latter 



ii, 12] SPACE OF N DIMENSIONS 29 

proposition is a new proposition concerning lines and points on a 
plane, which could have been obtained directly by interchanging 
the words point and line in the original proposition, supposing the 
latter to be expressed in the "on" language. This gives 

THEOREM 12. THE THEOREM OF DUALITY IN A PLANE. Any prop- 
osition deducible from Assumptions A and E concerning the points 
and lines of a plan&remains valid, if stated in the "on" terminology, 
when the words "point" and "line" are interchanged. (A, E) 

The space dual of this theorem then gives 

THEOREM 13. THE THEOREM OF DUALITY AT A POINT. Any prop- 
osition deducible from Assumptions A and E concerning the planes^ 

^alid y if stated in the "on" termi- 



nology, when the words "plane" and " line" are interchanged. (A, E) 

The principle of duality was first stated explicitly by Gergonne (1826), but 
was led up to by the writings of Poncelet and others during the first quarter 
of the nineteenth century. It should be noted that this principle was for 
several years after its publication the subject of much discussion and often 
acrimonious dispute, and the treatment of this principle in many standard 
texts is far from convincing. The method of formal inference from explicitly 
stated assumptions makes the theorems appear almost self-evident. This may 
well be regarded as one of the important advantages of this method. 

It is highly desirable that the reader gain proficiency in forming the duals 
of given propositions. It is therefore suggested as an exercise that he state 
the duals of each of the theorems and corollaries in this chapter. He should 
in this case state both the original and the dual proposition in the ordinary 
terminology in order to gain facility in dualizing propositions without first 
stating them in the often cumbersome "on" language. It is also desirable 
tnat he dualize several of the proofs by writing out in order the duals of each 
proposition used in the proofs in question. 

EXERCISE 

Prove the theorem of duality for a space of four dimensions : Any propo- 
sition derivable from the assumptions of alignment and extension and closure 
for a space of four dimensions concerning points, lines, planes, and three- 
spaces remains valid when stated in the " on " terminology, if the words 
point and three-space and the words line and plane be interchanged. ' 

* 12. The theorems of alignment for a space of n dimensions. We 
have already called attention to the fact that Assumption E3',- 
whereby we limited ourselves to the consideration of a space of only 

* This section may be omitted on a first reading. 



30 THEOEEMS OF ALIGNMENT AND DUALITY [CHAP. I 

three dimensions, is entirely arbitrary. This section is devoted to the 
discussion of the theorems of alignment, i.e. theorems derivable from 
Assumptions A and E 0, for a space of any number of dimensions. 
In this section, then, we make use of Assumptions A and E only. 

DEFINITION. If ^, P 19 -Z?, , P n are n + 1 points not on the same 
(n l)-space, and S n _ l is an ( l)-space on J%, P 2> - , P n) the class 
S n of all points such that every point of S n is on a line with J% and 
some point of S n _ 1 is called the n-space determined by J^ and S B-1 . 

As a three-space has already been defined, this definition clearly 
determines the meaning of "w-space" for every positive integral value 
of n. We shall use S n as a symbol for an %-space, calling a plane a 
2-space, a line a 1 -space, and a point a 0-space, when this is convenient. 
S is then a symbol for a point. 

DEFINITION. An S r is on an S t and an S^ is on an S r (r<t) 9 pro- 
vided that every point of S r is a point of S t . 

DEFINITION. Jo points are said to be independent, if there is no S^g 
which contains them all. 

Corresponding to the theorems of J 6-9 we shall now establish 
the propositions contained in the following Theorems S n l, S n 2, 
S ra 3. As these propositions have all been proved for the case n = 3, 
it is sufficient to prove them on the hypothesis that they have already 
been proved for the cases n = 3, 4, - -, n 1 ; i.e. we assume that the 
propositions contained in Theorem S n _ 1 l, a, &, c, d, e, /have been 
proved, and derive Theorem S n l, #,,/ from them. By the prin- 
ciple of mathematical induction this establishes the theorem for any n. 

THEOREM S n l. Let the n-space S n "be defined ly the point R and the 



a. There is an n-space on any n + 1 independent points. 
1). Any line on two points of S n has one point in common with R n _ a , 
and is on S n . 

c. Any S r (r < n) on r + 1 independent points of S n is on S n . 

d. Any S r (r < n) on r + 1 independent points of S n has an S rnl in 
common with R n _ lf provided all r + 1 points are not on R n _ r 

e. Any line I on two points of S n has one point in common with 
any S n _ I on S m . 

/ -Sf .T and T n _ a (T not on T n _^ are any point and any 
(n l.)-space respectively of the n-space determined ly R and R w-1 , 
the latter n-space is the same as that determined ly T and T w _ r 



12] SPACE OF JV DIMENSIONS 31 

Proof, a. Let the n + 1 independent points be JJ, P v - - . , P n . Then 
the points JJ, P 2 , -, JJ are independent; for, otherwise, there would 
exist an S n _ 2 containing them all (definition), and this S n _ 2 with J? 
would determine an S a-1 containing all the points JJ, j, - , %, con- 
trary to the hypothesis that they are independent. Hence, by Theorem 
S n _ 1 l a, there is an S^ on the points JJ, JJ, , Jg; and this S H ^ 
with JJ determines an w-space which is on the points JJ, JJ, J?, ,.. 

&. If the line I is on R or R n _ 1? the proposition is evident from the 
definition of S n . If I is not on R or R n _ 1} let A and J5 be the given 
points of I which are on S n . The lines R ^ and R Q B then meet R^ 
in two points A! and B J respectively. The line I then meets the two 
lines J3'R , R ^'; and hence, by Assumption A3, it must meet the 
line A'B* in a point P which is on R n-1 by Theorem S K _ 1 15. To 
show that every point of I is on S n , consider the points A, A r , P. Any 
line joining an arbitrary point Q of I to R , meets the two lines PA 
and A A 1 , and hence, by Assumption A3, meets the third line A 1 P. 
But every point of AlP is on R n _ l (Theorem S n-1 l 6), and hence Q 
is, by definition, a point of S n . 

c. This may be proved by induction with respect to r. For r 1 it 
reduces to Theorem S B 1 J. If the proposition is true for r = k 1,' all 
the points of an S k on k + 1 independent points of S n are, by definition 
and Theorem S^l/, on lines joining one of these points to the points 
of the S k _ l determined by the remaining k points. But under the 
hypothesis of the induction this S k _ 1 is on S n , and hence, by Theorem 
S n l 6, all points of S k are on S n . 

d. Let r+1 independent" points of S tt be J, J?, - -, P r and let j?J be 
not on R n _ r Each of the lines JJ P k (k = 1, , r) has a point Q k in 
common with R n _ l (by S B 1 J). The points Q 19 Q z , - - -, Q ? . are inde- 
pendent; for if not, they would all be on the same S ? ._ 2 , which, 
together with J^, would determine an S,^ containing all the points 
PI (by S ri-1 l 6). Hence, by S r-1 l a, there is an S r _ : on j, Q*,- ,Q r 
which, by c, is on both S r and S n . 

e. We will suppose, first, that one of the given points is R . Let 
the other be A. By definition I then meets R u-1 in a point A', and, by 
5,^15, in only one such point. If R is on S n _ 1} no proof is required 
for this case. Suppose, then, that R is not on S n _ 1 , and let C be any 
point of S n . r The line R (7 meets R^ in a point C r (by definition). 
By d, S n _ 1 has in common with R tt _! an (n 2)-space, S n _ 2 , and, by 



32 THEOREMS OF ALIGNMENT AND DUALITY [CHAP. I 

Theorem S W _ 1 1^, this has in common with the line A f C f at least 
one point D f . All points of the line D f C are then on S a-1 , by S n-1 l b. 
Now the line I meets the two lines C'D f and CC r ; hence it meets the 
line CD ] (Assumption A3), and has at least one point on S n _ r 

We will now suppose, secondly, that both of the given points are 
distinct from R . Let them be denoted by A and J3, and suppose that 
R is not on S tt _ r By the case just considered, the lines R Q A and 
R Q B meet S n-1 in two points A! and B 1 respectively. The line Z, which 
meets R ^' and R Q B r must then meet A'B 1 in a point which, by 
Theorem S n-1 lJ, is on S B-1 . 

Suppose, finally, that R is on S tt-1 , still under the hypothesis that Z 
is not on R . By d, S n _ l meets R n-1 in an (n 2)-space Q rt _ 2 , and 
the plane R Z meets R B-1 in a line I 1 . By Theorem S n _ l ie 9 V and 
Q n _ 2 have in common at least one point P. Now the lines Z and R P 
are on the plane R Z, and hence have in common a point Q (by Theorem 
S a l e = Theorem 5). By S W-1 1 b the point Q is common to S n _ 1 and Z. 

/. Let the %-space determined by T and T n-1 be denoted by T w . 
Any point of T n is on a line joining T with some point of T n _ r 
Hence, by &, every point of T n is on S n . Let P be any point of S n 
distinct from T . The line T P meets T n-1 in a point, by e. Hence 
every point of S B is a point of T B . 

GOROLLAEY. On n + 1 independent points there is one and but one S n . 

This is a consequence of Theorem S n l a and S n l/. The formal 
proof is left as an exercise. 

THEOEEM S n 2. An S r and an S k having in common an S p , but 
not an S p+1 , are on a common S r+ ^._ p and are not both on the so/me 
s > if n<r + kp. 

Proof. If Jc~p, S k is on S r . If Jc >p, let JJ be a point on S k not 
on S 9 . Then 1J and S r determine an S r+1 , and P^ and S p an S |)+l> 
such that S p+l is contained in S r+1 and S t . If Is > p + 1, let P% be a 
point of S k not on S p+1 . Then P z and S r+1 determine an S r+2 , while 
J> and S p+1 determine an S^ +2 , which is on S r+2 and S r This process 
can be continued until there results an S^, containing all the points 
of S k . By Theorem S n l, Cor., we have i=sJcp. At this stage in the 
process we obtain an S r+ j.^ which contains both S r and S k . 

The argument just made shows that 1J, JJ, , P k _ p , together with 
any set Q v Q z> . . ., Q r+1 , of r + 1 independent points of S r , constitute 



12] SPACE OF N DIMENSIONS S3 

a set of r + k p + 1 independent points, each of which is either in 
S, or S A . If S r and S k were both on an S n , where n< r + & p, these 
could not be independent. 

THEOKEM S W 3. ^w S r and an S k contained in an S tt ar &0#i 00, the 
same S,. +jt _ w . 

Proof. If there were less than r + k n+1 independent points 
common to S, and S^ say r + kn points, they would, by Theorem S n 2, 
determine an S q , where q = r + k (r + k n 1) = n -f 1. 

Theorems S n 2 and S M 3 can be remembered and applied very easily 
by means of a diagram in which S M is represented by n + 1 points. 
Thus, if n = 3, we have a set of four points. That any two S 's have 
an S x in common corresponds to the fact that any two sets of three 
must have at least two points in common. In the general case a set 
of r + 1 points and a set of k + 1 selected from the same set of n + 1 
have in common at least r + k n + 1 points, and this corresponds 
to the last theorem. This diagram is what our assumptions would 
describe directly, if Assumption E were replaced by the assumption: 

JSvery line contains two and only tivo points. 

If one wishes to confine one's attention to the geometry in a space 
of a given number of dimensions, Assumptions E 2, E 3, and E 3' may 
be replaced by the following : 

En. Not all points are on the same S XJ if k < n. 

En'. If S is an S n , all points are on S. 

Eor every S n there is a principle of duality analogous to that which, 
we have discussed for n = 3. In S n the duality is between S k and S n _ t _ l 
(counting a point as an.$ ), for all &'s from to n 1. If n is odd, 
there is a self-dual space in S n ; if n is even, S ff contains no self-dual 
space. 

EXERCISES 

1. State and prove the theorems of duality in S 5 ; in S tt . 

2. If m + 1 is the number of points on a line, how many S A /s are there in 
anS B ? 

* 3. State the assumptions of extension by which to replace Assumption En 
and En' for spaces of an infinite number of dimensions. Make use of the 
transfmite numbers. 

* Exercises marked * are of a more advanced or difficult character. 



CHAPTER II 

PROJECTION, SECTION, PERSPECTIVITY. ELEMENTARY 
CONFIGURATIONS 

13. Projection, section, perspectivity. The point, line, and plane 
are the simple elements of space * ; we have seen in the preceding 
chapter that the relation expressed by the word on is a reciprocal 
relation that may exist between any two of these simple elements. 
In the sequel we shall have little occasion to return to the notion of 
a line as being a class of points, or to the definition of a plane ; but 
shall regard these elements simply as .entities for which the relation 
" on " has been denned. The theorems of the preceding chapter are to 
be regarded as expressing the fundamental properties of this relation.! 
We proceed now to the study of certain sets of these elements, and 
begin with a series of definitions. 

DEFINITION. A figure is any set of points, lines, and planes in space. 
A plane figure is any set of points and lines on the same plane. A 
point figure is any set of planes and lines on the same point. 

It should be observed that the notion of a point figure is the space 
dual of the notion of a plane figure. In the future we shall fre- 
quently place dual definitions and theorems side by side. By virtue 
of the principle of duality it will be necessary to give the proof of 
only one of two dual theorems. 

DEFINITION. Given a figure F DEFINITION. Given a figure F 
and a point P\ every point of F and a plane TT; every plane of F 
distinct from P determines with distinct from TT determines with 
P a line, and every line of F not TT a line, and every line of F not 
on P determines with P a plane; on TT determines with TT a point; 
the set of these lines and planes the set of these lines and points 
through P is called the projection on TT is called the section $ of F 

* The word space is used in place of the three-space in which are all the elements 
considered. 

t We shall not in future, however, confine ourselves to the " on " terminology, 
but shall also use the more common expressions. 

| A section by a plane is often called a plane section. 

34 



13] PEOJECTION, SECTION", PEESPECTIYITY 35 

of F from P, The individual lines by TT. The individual lines and 

and planes of the projection are points of the section are also 

also called the projectors of the called the traces of the respective 

respective points and lines of F. planes and lines of F. 

If F is a plane figure and the point P is in the plane of the figure, the 
definition of the projection of F from P has the following plane dual : 

DEFINITION. Given a plane figure F and a line I in the plane of F; 
the set of points in which the lines of F distinct from I meet I is 
called the section of F by Z. The line I is called a transversal, and 
the points are called the traces of the respective lines of F. 

As examples of these definitions we mention the following: The 
projection of three mutually intersecting nonconcurrent lines from a 
point P not in the plane of the lines consists of three planes through P; 
the lines of intersection of these planes are part of the projection only 
if the points of intersection of the lines are thought of as part of the 
projected figure. The section of a set of planes all on the same line 
by a plane not on this line consists of a set of concurrent lines, the 
traces of the planes. The section of this set of concurrent lines in a 
plane by a line in the plane not on their common point consists of 
a set of points on the transversal, the points being the traces of the 
respective lines. 

DEFINITION. Two figures F 1? F 2 are said to be in (1, 1) correspond- 
ence or to correspond in a one-to-one reciprocal way, if every element 
of F.,^ corresponds (cf. footnote, p. 5) to a unique element of F 2 in such 
a way that every element of F 2 is the correspondent of a unique ele- 
ment of F r A figure is in (1, 1) correspondence with itself, if every 
element of the figure corresponds to a unique element of the same 
figure in such a way that every element of the figure is the corre- 
spondent of a unique element. Two elements that are associated in 
this way are said to be corresponding or homologous elements. 

A correspondence of fundamental importance is described in the 
following definitions: 

DEFINITION. If any two homol- DEFINITION. If any two homol- 
ogous elements of two corre- ogous elements of two corre- 
sponding figures have the same spending figures have the same 
projector from a fixed point 0, trace in a fixed plane o>, such 
such that all the projectors are that all the traces of either 



36 PROJECTION", SECTION, PEESPECTIVITY [CHAP. II 

distinct, the figures are said to figure are distinct, the figures are 
be perspective from 0. The point said to be perspective from co, 
is called the center of perspec- The plane w is called the plane 
tivity. of perspectivity. 

DEFINITION. If any two homologous lines in two corresponding 
figures in the same plane have the same trace on a line I, such 
that all the traces of either figure are distinct, the figures are said 
to be perspective from 1. The line I is called the axis of perspectivity. 

Additional definitions of perspective 'figures will be given in the 
next chapter (p. 56). These are sufficient for our present purpose. 

DEFINITION. To project a figure in a plane a from a point onto a 
plane a r , distinct from a, is to form the section by cc r of the projection 
of the given figure from 0. To project a set of points of a line I from 
a point onto a line V 9 distinct from I but in the same plane with I 
and 0, is to form the section by l f of the projection of the set of points 
from 0. 

Clearly in either case the two figures are perspective from 0, pro- 
vided is not on either of the planes a, a 1 or the lines I, l r . 

EXERCISE 

What is the dual of the process described in the last definition ? 

The notions of projection and section and perspectivity are fun- 
damental in all that follows.* They will be made use of almost 
immediately in deriving one of the most important theorems of pro- 
jective geometry. We proceed first, however, to define an important 
class of figures. 

14. The complete n-point, etc. DEFINITION. A complete n-point in 
space or a complete space n-point is the figure formed by n points, no 
four of which lie in the same plane, together with the n(n 1)/2 
lines joining every pair of the points and the n(n l)(n 2)/6 planes 
joining every set of three of the points. The points, lines, and planes 
of this figure are called the vertices, edges, and faces respectively of 
the complete 7i-point 

* The use of these notions in deriving geometrical theorems goes back to early 
times. Thus, e.g., B. Pascal (1623-1662) made use of them in deriving the theorem 
on a hexagon inscribed in a conic which bears his name. The systematic treatment 
of these notions is due to Ponceletj cl his Traits des proprie'te's proiectives des 
figures, Paris, 1822. 



14] JV-POINT, iV-PLANE, JV-LIXE 37 

The simplest complete %-point in space is the complete space 
four-point. It consists of four vertices, six edges, and four faces, 
and is called a tetrahedron. It is a self-dual figure. 

EXERCISE 

Define the complete n-plane in space by dualizing the last definition. The 
planes, lines, and points of the complete n-plane are also called the faces, 
edges, and vertices of the n-plane. 

DEFINITION. A complete n-point in a plane or a complete plane 
n-point is the figure formed by n points of a plane, no three of 
which are collinear, together with the n(n 1)/2 lines joining every 
pair of the points. The points are called the vertices and the lines 
are called the sides of the n-point. The plane dual of a complete 
plane ^-point is called a complete plane n-line. It has n sides and 
n(n V)/2 vertices. The simplest complete plane n-point consists of 
three vertices and three sides and is called a triangle. 

DEFINITION. A simple space nyoint is a set of n points J^, J^, P v ,P n 
taken in a certain order, in which no four consecutive points are 
coplanar, together with the n lines J?J, JJJJ, , PJE( joining suc- 
cessive points and the n planes J%J%Z, -, J^^ determined by 
successive lines. The points, lines, and planes are called the vertices, 
edges, and faces respectively of the figure. The space dual of a simple 
space 7i-point is a simple space n-plane. 

DEFINITION. A simple plane n-point is a set of n points P^, P%, P%, - JJ 
of a plane taken in a certain order in which no three consecutive points 
are collinear, together with the n lines JJ.2J, P^, -, P n l^ joining suc- 
cessive points. The points and lines are called the vertices and sides 
respectively of the figure. The plane dual of a simple plane n-point is 
called a simple plane n-line. 

Evidently the simple space n-point and the simple space n-plane are 
identical figures, as likewise the simple plane n-point and the simple 
plane n-line. Two sides of a simple n-line which meet in one of its 
vertices are adjacent. Two vertices are adjacent if in the dual relation. 
Two vertices of a simple n-point J%P% - P n (n even) are opposite if, in 
the order J^ P n , as many vertices follow one and precede the other 
as precede the one and follow the other. If n is odd, a vertex and a 
side are opposite if, in the order .TJJfJ - P n , as many vertices follow the 
side and precede the vertex as follow the vertex and precede the side. 



38 



PROJECTION, SECTION, PEESPECTIVITY [CHAP.II 



The space duals of the complete plane %-point and the complete plane 
%-line are the complete n-plane on a point and the complete n-line on a 
point respectively. They are the projections from a point, of the plane 
^-line and the plane ^-point respectively. 

15. Configurations. The figures defined in the preceding section 
are examples of a more general class of figures of which we will now 
give a general definition. 

DEFINITION. A figure is called a configuration, if it consists of a 
finite number of points, lines, and planes, with the property that each 
point is on the same number & 12 of lines and also on the same num- 
ber a ia of planes ; each line is on the same number & 21 of points and the 
same number # 23 of planes ; and each plane is on the same number a 81 
of points and the same number a 32 of lines. 

A configuration may conveniently be described by a square matrix : 



1 point 

2 line 

3 plane 



1 

point 


2 

line 


3 

plane 


31 32 33 



In this notation, if we call a point an element of the first kind, a 
line an element of the second kind, and a plane one of the third kind, 
the number a (i = j) gives the number of elements of the /th kind 
on every element of the ith kind. The numbers a iv a 22t a ss give the 
total number of points, lines, and planes respectively. Such a square 
matrix is called the symbol of the configuration. 

A tetrahedron, for example, is a figure consisting of four points, 
six lines, and four planes ; on every line of the figure are two points 
of the figure, on every plane are three points, through every point 
pass three lines and also three planes, every plane contains three lines, 
and through every line pass two planes. A tetrahedron is therefore 
a configuration of the symbol 




COlSTFIG-UBATIOlSrS 



39 



The symmetry shown in this symbol is due to the fact that the figure 
in question is self-dual. A triangle evidently has the symbol 




Since all the numbers referring to planes are of no importance in 
case of a plane figure, they are omitted from the symbol for a plane 
configuration. 

In general, a complete plane ?i-point is of the symbol 



n 
2 


w-1 



and a complete space w-point of the symbol 



2 Jn(n-l) w-2 

3 3 t n (n-l)(a-2) 



Further examples of configurations are figs. 14 and 15, regarded as 
plane figures. 

EXERCISE 

Prove that the numbers in a configuration symbol must satisfy the condition 

(* / = 1, 2, 8) 



16. The Desargues configuration. A very important configuration 
is obtained by taking the plane section of a complete space five-point. 
The five-point is clearly a configuration with the symbol 




and it is clear that the section by a plane not on any of the vertices 
is a configuration whose symbol may be obtained from the one just 
given by removing the first column and the first row. This is due 
to the fact that every line of the space figure gives rise to a point in 



40 



PBO JECTION, SECTION, PEESPECTIYITY [CHAP, n 



the plane, and every plane gives rise to a line. The configuration in 
the plane has then the symbol 




We proceed to study in detail the properties of the configuration just 
obtained. It is known as the configuration of Desargues. 

We may consider the vertices of the complete space five-point as con- 
sisting of the vertices of a triangle A, B 9 C and of two points O l} 2 




Fio. 14 

not coplanar with any two vertices of the triangle (fig. 14). The sec- 
tion by a plane a not passing through any of the vertices will then 
consist of the following : 

A triangle A^B^C^ the projection of the triangle ABC from : on a. 

A triangle J 2 2? 2 <7 2 , the projection of the triangle ABC from 2 on a. 

The trace of the line 0^0 y 

The traces A zi B^ <7 3 of the lines BC, CA, AB respectively. 

The trace of the plane ABC, which contains the points A z , B^ C y 

The traces of the three planes AO^O^ BO^O^ COft^, which contain 
respectively the triples of points OA^A^ OB^B^ OC^C y 

The configuration may then be considered (in ten ways) as consist- 
ing of two triangles A^B^ and A^B^C V perspective from a point and 



THEOREM OF DESAEGUES 41 

having homologous sides meeting in three collinear points J 3 , B B , <7 8 . 
These considerations lead to the following fundamental theorem: . 

THEOREM 1. THE THEOREM OF DESARGUES.* If two triangles in the 
same plane are 'perspective "from a point, the three pairs of homologous 
sides meet in collinear points; i.e. the triangles are perspective from 
a line. '(A, E) 

Proof. Let the two triangles be A^B^ and A 2 J3 Z C 2 (fig. 14), the 
lines A^A^ B^^ ^aA meeting in the point 0. Let B^A V B^A 2 inter- 
sect in the point <? 3 ; A^C V A 2 C Q in B^\ B^C V B Z C Z in A z . It is required 
to prove that A z , B B , <? s are coUinear, Consider any line through 
which is not in the plane of the triangles, and denote by O v 3 any 
two distinct points on this line other than 0. Since the lines A 2 Z 
and A^O^ lie in the plane (A^A^ 1 2 ), they intersect in a point A. 
Similarly, B^O^ and J? 2 2 intersect in a point B, and likewise C^ and 
C 2 2 in a point C. Thus ABCO^O^ together with the lines and planes 
determined by them, form a complete five-point in space of which the 
perspective triangles form a part of a plane section. The theorem 
is proved by completing the plane section. Since AB lies in a plane 
with AJ& V and also in a plane with A Z B 2 , the lines AJ& V A Z B 2) and 
AB meet in 9 . So also A 1 C V A Z C Z) and AC meet in B B ; and B t C v 
B Z C 2 , and BC meet in A y Since A z) B z , C s lie in the plane ABC and 
also in the plane of the triangles A^B^C^ and A^B^C^ they are collinear. 

THEOREM l f . If two triangles in the same plane are perspective 
from a line, the lines joining pairs of homologous vertices are, con- 
current; i.e. the triangles are perspective from a point. (A, E) 

This, the converse of Theorem 1, is also its plane dual, and hence 
requires no further proof. 

COROLLARY. If two triangles not in the same plane are perspective 
from a point, the pairs of homologous sides intersect in collinear 
points; and conversely. (A, E) 

A more symmetrical and for many purposes more convenient nota- 
tion for the Desargues configuration may be obtained as follows: 
Let the vertices of the space five-point be denoted by ^, J^, P t) P^ J 
(fig. 15). The trace of the line P^ in the plane section is then ' 
naturally denoted by ^ in general, the trace of the line JJJ^ by 7?. 
(i, j = 1, 2, 3, 4, 5, i = /). Likewise the trace of the plane j?J^5 ma 7 
* G-irard Desargues, 1593-1662. 



42 PKQJECTICH8", SECTION, PER8PECTIVITY [CHAP, n 

be denoted by l vt (i, j, k = 1, 2, 3, 4, 5). This notation makes it pos- 
sible to te]l at a glance which lines and points are united. Clearly a 
point is on a line of the configuration if and only if the suffixes of 
the point are both among the suffixes of the line. Also the third 
point on the line joining JJ and J^ is the point JJ, ; two points are 
on the same line if and only if they have a suffix in common, etc. 




EXERCISES 

1. Prove Theorem V without making use of the principle of duality. 

2. If two complete w-points in different planes are perspective from a point, 
the pairs of homologous sides intersect in collinear points. "What is the dual 
theorem ? What is the corresponding theorem concerning any two plane figures 
in. different planes ? 

3. State and prove the converse of the theorems in Ex. 2. 

4. If two complete ra-points in the same plane correspond in such a way 
that homologous sides intersect in points of a straight line, the lines joining 
homologous vertices are concurrent ; i.e. the two w-points are perspective from 
a point. Dualize. 

5. What is the figure formed by two complete n-points in the same plane 
when they are perspective from a point? Consider particularly the cases n = 4 and 
n = 5. Show that the figure corresponding to the general case is a plane section 
of a complete space (n + 2)-point. Give the configuration symbol and dualize. 

6. If three triangles are perspective from the same point, the three axes of 
perspectivity of the three pairs of triangles are concurrent ; and conversely. 
Dualize, and compare the configuration of the dual theorem with the case n = 4 
of Ex. 5 (cf. fig. 15, regarded as a plane figure). 



17] 



PEESPECTIVE TETEAHEDEA 



43 



17. Perspective tetrahedra. As an application of tlie corollary of 
the last theorem we may now derive a theorem in space analogous to 
the theorem of Desargues in the plane. 

THEOREM 2. If two tetrahedra are perspective from a point, tlie six 
p airs of homologous edges intersect in coplanar points, and the four 
pairs of homologous faces intersect in coplanar lines ; i.e. the tetra- 
are perspective from a plane. (A, E) 




:*' 



FIG. 16 

Proof. Let the two tetrahedra be -Zy^J-ZJ and JJ'JJ'JJ'JEJ', and let 
the lines JJJJ', J^', jZ^', JJ-Z^ meet in the center of perspectivity 0. 
Two homologous edges P % P 3 and J?'J' then clearly intersect ; call the 
point of intersection P v . The points ^ 2 , J? s , JJ S lie on the same line, 
since the triangles JJJSJJ and -ZJ'Ia'JJ' are perspective from (The- 
orem 1, Cor.). By similar reasoning applied to the other pairs of 
perspective triangles we find that the following triples of points are 
collinear: 

The first two triples have the point P^ in common, and hence 
determine a plane;, each of the other two triples has a point in 



44 PKOJECTION, SECTION, PEESPECTIYITY [CHAP.II 

common with each, of the first two. Hence all the points P %1 lie in 
the same plane. The lines of the four triples just given are the lines 
of intersection of the pairs of homologous faces of the tetrahedra. 
The theorem is therefore proved. 

THEOREM 2'. If two tetrahedra are perspective from a plane, the 
lines joining pairs of homologous vertices are concurrent, as likewise 
the planes determined ly pairs of homologous edges ; i.e. the tetrahedra 
are perspective from a point (A, E) 

This is the space dual and the converse of Theorem 2. 

EXERCISE 

"Write the symbols for the configurations of the last two theorems. 

18. The quadrangle-quadrilateral configuration. 

DEFINITION. A complete plane DEFINITION. A complete plane 
four-point is called a complete four-line is called a complete 
quadrangle. It consists of four quadrilateral It consists of four 
vertices and six sides. Two sides sides and six vertices. Two ver- 
not on the same vertex are called tices not on the same side are 
opposite. The intersection of two called opposite. The line joining 
opposite sides is called a diag- two opposite vertices is called a 
onal point. If the three diagonal diagonal line. If the three diag- 
points are not collinear, the tri- onal lines are not concurrent, the 
angle formed by them is called triangle formed by them is called 
the diagonal triangle of the the diagonal triangle of the 
quadrangle.* quadrilateral* 

The assumptions A and E on which all our reasoning is based do 
not suffice to prove that there are more than three points on any line. 
In fact, they are all satisfied by the triple system (1), p. 3 (cf. fig. 17). 
In a case like this the diagonal points of a complete quadrangle, ,ajce 
^collinear.and.the di^onSTlmes of sT^qm^ete quadrilateral concur- 
rent^ as may readily be verified. Two perspective triangles cannot 
exist in such a plane, and hence the Desargues theorem becomes 

* In general, the intersection of two sides of a complete plane w-point which do 
not have a vertex in common is called a diagonal poM of the ra-point, and the line 
joining two vertices of a complete plane n-line which do not lie on the same side 
is called a diagonal line of the 7i-line. A complete plane n~point (n-line) then has 
n(n - 1) (n - 2) (n -3)/8 diagonal points (lines). Diagonal points and lines are 
sometimes called false vertices and false sides respectively. 



ifl ASSUMPTION H 








trivial. Later on we shall add an assumption* which excludes all 
such cases as this, and, in fact, provides for the existence of an in- 
finite number of points on a line. A part of what is contained in 
this assumption is the following: 

ASSUMPTION H . Tlie diagonal 
points of a complete quadrangle 
are noncollinear. 

Many of the important theorems 
of geometry, however, require the 
existence of no more than a finite 
number of points. "We shall there- 
fore proceed without the use of w P IG 17 

further assumptions than A and E, 

understanding that in order to give our theorems meaning there must 
~be postulated, the existence of the points specified in their hypotheses. 
In most cases the existence of a sufficient number of points is 
insured by Assumption H , and the reader who is taking up the 
subject for the first time may well take it as having been added 
to A and E. It is to be used in the solution of problems. 

We return now to a further study of the Desargues configuration. 
A complete space five-point may evidently be regarded (in five ways) 
as a tetrahedron and a complete four-line at a point A plane section 
of a four-line is a quadrangle and the plane section of a tetrahedron 
is a quadrilateral. It follows that (in five ways) the Desargues con- 
figuration may be regarded as a quadrangle and a quadrilateral. 
Moreover, it is clear that the six sides of the quadrangle pass through 
the six vertices of the quadrilateral. In the notation described on 
page 41 one such quadrangle is JJ 2 , ^ 3 , JJ 4 , JJ 5 and the corresponding 
quadrilateral is Z 234 , Z 235 , ? 2 45> Z 845 . 

The question now naturally arises as to placing the figures thus ob- 
tained in special relations. As an application of the theorem of De- 
sargues we will show how to construct f a quadrilateral which has the 
same diagonal triangle as a given quadrangle. "We will assume in our 
discussion that the diagonal points of any quadrangle form a triangle. 

* Merely saying that there are more than three points on a line does not insure 
that the diagonal points of a quadrangle are noncollinear. Cases where the diagonal 
points are collinear occur whenever the number of points on a line is 2 + ! 

t To construct a figure is to determine its elements in terms of certain given 
elements. 



46 PROJECTION, SECTION", PEKSPECTIVITY [CHAP, n 

Let PV P^ p *> % be the vertices of the given complete quadrangle, 
and let D 12 , Z> 13 , D u be the vertices of the diagonal triangle, D 12 being 
on the side Pf^ > 1B on the side '%P B , and D 14 on the side JJJJ (fig. 18). 
We observe first that the diagonal triangle is perspective with each of the 
four triangles formed ly a set of three of the vertices of the quadrangle, 
the center of perspectivity being in each case the fourth vertex. This 
gives rise to four axes of perspectivity (Theorem 1), one corresponding 
to each vertex of the quadrangle.* These four lines clearly form the 
sides of a complete quadrilateral whose diagonal triangle is Z> 12 , D 13 , Z> 14 . 




It may readily be verified, by selecting two perspective triangles, 
that the figure just formed is, indeed, a Deja^^ This 

special case of the Desargues configuration is called the qruadrangle- 
qiiadrilateral configuration.^ 

EXERCISES 

1. If p is the polar of P with regard to the triangle ABC, then P is the 
pole of p with regard to the same triangle ; that is, P is obtained from p by 
a construction dual to that used in deriving p from P. From this theorem it 
follows that the relation between the quadrangle and quadrilateral in this 

* The line thus uniquely associated with a vertex is called the polar of the point 
with respect to the triangle formed by the remaiDing three vertices. The plane dual 
process leads to a point associated with any line. This point is called the pole of the 
line with respect to the triangle. 

t A further discussion of this configuration and its generalizations will be found 
in the thesis of H. F. McNeish. Some of the results in this paper are indicated in 
the exercises. 



18,19] QUADBAjSTGULAK SETS 47 

configuration is mutual ; that is, if either is given, the other is determined. 
For a reason ^rhich will be evident later, either is called a covariant of the 
other. 

2. Show that the configuration consisting of two perspective tetrahedra, 
their center and plane of perspectivity, and the projectors and traces may be 
regarded in six ways as consisting of a complete 5-point P 12 , P 13 , P u , P 15 , P 16 
and a complete 5-plane ir S456 , *- 3456 , 7r 233G , 7r 2S46 , 7r 234 -, the notation being 
analogous to that used on page 41 for the Desargues configuration. Show- 
that the vertices of the 5-plane are on the faces of the 5-point. 

3. If P 15 P 2 , P g , P 4 , P 5 , are vertices of a complete space 5-point, the ten 
points Dy, in which an edge p {j meets a face P l P l P m (i,/, fc, I, m all distinct), 
are called diagonal points. The tetrahedra P 2 P 8 P 4 P 5 and -Z^-^is^w^iB are per- 
spective with P x as center. Their plane of perspectivity, v 19 is called the polar 
of P x with regard to the four vertices. In like manner, the points P 2 , P 3 , P 4 , P 5 
determine their polar planes 7r 2 , 7r 3 , 7r 4 , ir e . Prove that the 5-point and the polar 
5-plane form the configuration of two perspective tetrahedra ; that the plane 
section of the 5-point by any of the five planes is a quadrangle-quadrilateral 
configuration ; and that the dual of the above construction applied to the 5-plane 
determines the original 5-point. 

4. If P is the pole of IT with regard to the tetrahedron A^A 2 A Z A^ then is v 
the polar of P with regard to the same tetrahedron ? 

19. The fundamental theorem on quadrangular sets. 

THEOREM 3. If two complete quadrangles %%J%% and P^PjPjPl 
correspond P^ to P^, 1% to P^ etc. in suc7i a way that five of the 
pairs of homologous sides intersect in points of a line I, then the sixth 
pair of homologous sides will intersect in a point of I. (A, E) 

This theorem holds whether the quadrangles are in the same or 
in different planes. 

Proof. Suppose, first, that none of the vertices or sides of one of 
the quadrangles coincide with any vertex or side of the other. Let 
%Pv %J%, -ZJ-Z2, JJJJ, P^PI be the five sides which, by hypothesis, 
meet their homologous sides P^Pj, P^Pj, PJPJ, PJPJ, P^Pj in points 
of I (fig. 19). We must show that P Z P^ and IgP meet in a point 
of I. The triangles ^P^ and P^P* are, by hypothesis, perspec- 
tive from l\ as also the triangles P^P and P^P^P^. Each pair is 
therefore (Theorem I/) jpersgective from appoint, and this point is in 
each case the intersection^^ of the lines JJZJ 7 and P^Ef. Hence the 
triangles P^P^ and ^^^ are perspective from and their pairs 
of homologous sides intersect in the points of a line, which is evi- 
dently I, since it contains two points of L But P^ and $P are 



48 



PBOJECTIOST, SECTION, PEESPECTIYITY [CHAP.H 



two homologous sides of these last two triangles. Hence they inter- 
sect in a point of the line I. 

If a vertex or side of one quadrangle coincides with a vertex or 
side of the other, the proof is made by considering a third quadrangle* 
whose vertices and sides are distinct from those of both of the others, 
and which has five of its sides passing through the five given points 




of intersection of homologous sides of the two given quadrangles. By 
the argument above, its sixth side will meet the sixth side respectively 
of each of the two given quadrangles in the same point of I. This 
completes the proof of the theorem. 



1. It should be noted that the theorem is still valid if the line I con- 
tains one or more of the diagonal points of the quadrangles. The case in which 
I contains two diagonal points is of particular importance and will be discussed 
in Chap. IV, 31. 

NOTE 2. It is of importance to note in how far the quadrangle P^P^P^P^ 
is determined when the quadrangle P i P 2 P 8 P 4 and the line I are given. It may 
be readily verified that in such a case it is possible to choose any point P{ to 
correspond to any one of the vertices P t , P 2 , P 8 , P 4 , say P x ; and that if m is 
any line of the plane IP{ (not passing through Pf ) which meets one of the sides, 
say a, of P^PJP^P^ (not passing through P x ) in a point of I, then m may be 
chosen as the side homologous to a. But then the remainder of the figure is 
uniquely determined. 



* This evidently exists whenever the theorem is not trivially obvious. 



19 ] QUADBAKG-ULAK SETS 49 

THEOREM 3'. If two complete quadrilaterals a^a^ and a(a^a(a[ 
correspond a l to a[, 3 to a! 2) etc. in such a way that five of the lines 
joining homologous vertices pass through a point P, the line joining the 
sixth pair of homologous vertices will also pass through P. (A, E) 

This is the plane dual of Theorem 3 regarded as a plane theorem. 

DEFINITION. A set of points in which the sides of a complete quad- 
rangle meet a line I is called a quadrangular set of points. 

Any three sides of a quadrangle either, form a triangle or meet in 
a vertex ; in the former case they are saifl. to form a triangle triple, 
in the latter > point triple of lines. In a quadrangtilar set of points 
on a line I any three points in which the lines of a triangle triple meet I 
is called a triangle triple of points in the set ; three points in which 
the lines of a point triple meet I are called a point triple of points. 
A quadrangular set of points will be denoted by 

Q(ABC, DBF), 

where ABC is a point triple and DEF is a triangle triple, and 
where A and D, B and E, and and F are respectively the inter- 
sections with the line of the set of the pairs of opposite sides of 
the quadrangle. 

The notion of a quadrangular set is of great importance in much 
that follows. It should be noted again in this connection that one 
or two * of the pairs A, D or B, U or (7, F may consist of coincident 
points ; this occurs when the line of the set passes through one or 
two of the diagonal points.f 

We have just seen (Theorem 3) that if we have a quadrangular 
set of points obtained from a given quadrangle, there exist other 
quadrangles that give rise to the same quadrangular set In the 
quadrangles mentioned in Theorem 3 there corresponded to every 
triangle triple of one a triangle triple of the other. 

DEFINITION. When two quadrangles giving rise to the same 
quadrangular set are so related with reference to the set that to a 
triangle triple of one corresponds a triangle triple of the other, the 

* All three may consist of coincident points in a space in which the diagonal points 
of a complete quadrangle are collinear. 

t It should he kept in mind that similar remarks and a similar definition may he 
made to the effect that the lines joining the vertices of a quadrilateral to a point P 
form a quadrangular set of lines, etc. (cf. 30, Chap. IV). 



50 



PKOJECTION, SECTION, PEBSPECTIVITY [CHAP, n 



quadrangles are said to be similarly placed (fig. 20); if a point triple 
of one corresponds to a triangle triple of the other, they are said to 
be oppositely placed (fig. 21). 

It will be shown later (Chap. IV) that quadrangles oppositely 
placed with respect to a quadrangular set are indeed possible. 




FIG. 20 




FIG. 21 

With the notation for quadrangular sets defined above, the last 
theorem leads to the following 

COROLLARY. If all "but one of the points of a quadrangular set Q (ABC, 
DBF) are given, the remaining one is uniquely determined. (A, E) 

For two quadrangles giving rise to the same quadrangular set 
with the same notation must be similarly placed, and must hence 
be in correspondence as described in -the theorem. 



19,20] DESAEGUES CONFIGURATION 51 

The quadrangular set which is the section by a 1-space of a complete 4-point 
in a 2-space, the Desargues configuration which is the section by a 2-space of 
a complete 5-point in a 3-space, the configuration of two perspective tetra- 
hedra which maybe considered as the section by a 3-space of a complete 6-point 
in a 4-space are all special cases of the section by an n-space of a complete 
(n + 3)-point in an (n + l)-space. The theorems which we have developed for 
the three cases here considered are not wholly parallel. The reader will find 
it an entertaining and far from trivial exercise to develop the analogy in full. 

EXERCISES 

1. A necessary and sufficient condition that three lines containing the ver- 
tices of a triangle shall be concurrent is that their intersections P, Q, It with 
a line I form, with intersections E, F, G of corresponding sides of the triangle 
with I, a quadrangular set Q(PQR, EFG).' 

2. If on a given transversal line two quadrangles determine the same quad- 
rangular set and are similarly placed, their diagonal triangles are perspective 
from the center of perspectivity of the two quadrangles. 

3. The polars of a point P on a line I with regard to all triangles which 
meet I in three fixed points pass through a common point P' on L 

4. In a plane TT let there be given a quadrilateral a 19 2 , a s , a 4 and a point 
not on any of these lines. Let A 19 A 2 , A%, A 4 be any tetrahedron whose four 
faces pass through the lines a v 2 , a 3 , a respectively. The polar planes of 
with respect to all such tetrahedra pass through the same line of TT. 

20. Additional remarks concerning the Desargues configuration. 

The ten edges of a complete space five-point may be regarded (in 
six ways) as the edges of two simple space five-points. Two such 
five-points are, for example, JJjZyjJJJj and P^P^P^. Corresponding 
thereto, the Desargues configuration may be regarded in six ways 
as a pair of simple plane pentagons (five-points). In our previous 
notation the two corresponding to the two simple space five-points 
just given are ^J^^^^i an< i -^is-^B^a^L^Jr Every vertex* of each 
of these pentagons is on a side of the other. 

Every point, P^ for instance, has associated with it a unique line 
of the configuration, viz. 1 M5 in the example given, whose notation 
does not contain the suffixes occurring in the notation of the point. 
The line may be called the polar of the point in the configuration, 
and the point the pole of the line. It is then readily seen that the 
polar of any point is the axis of perspectivity of two triangles 
whose center of perspectivity is the point. In case we regard the 
configuration as consisting of a complete quadrangle and complete 



52 PEOJBCTIOK, SECTION, PERSPECTIVITY [CHAP, n 

quadrilateral, it is found that a pole and polar are homologous vertex 
and side of the quadrilateral and quadrangle. If we consider the 
configuration as consisting of two simple pentagons, a pole and polar 
are a vertex and its opposite side, e.g. ^ and -? 4 ^ 5 . 

The Desargues configuration is one of a class of configurations 
having similar properties. These configurations have been studied 
by a number of writers.* Some of the theorems contained in these 
memoirs appear in the exercises below. 

EXERCISES 

In discussing these exercises the existence should be assumed of a sufficient number 
of points on each line so that the figures in question do not degenerate. In some cases 
it may also be assumed that the diagonal points of a complete quadrangle are not 
cottinear. Without these assumptions our theorems are true, indeed, but trivial. 

1. What is the peculiarity of the Desargues configuration obtained as the 
section of a complete space five-point by a plane which contains the point of 
intersection of an edge of the five-point with the face not containing this edge ? 
also by a plane containing two or three such points ? 

2. Given a simple pentagon in a plane, construct another pentagon in the 
same plane, whose vertices lie on the sides of the first and whose sides con- 
tain the vertices of the first (cf. p. 51). Is the second uniquely determined 
when the first and one side of the second are given? 

S, 3. If two sets of three points A, B, C and A', B' } C' on two coplanar lines 
Z and V respectively are so related that the lines A A', BB', CC f are concurrent, 
then the points of intersection of the pairs of lines AB f and BA', BC' and CB', 
CA' and A C' are collinear with the point ZZ'. The line thus determined is called 
the polar of the point (A A', BB') with respect to I and Z'. Dualize. 

4. Using the theorem of Ex. 3, give a construction for a line joining any 
given point in the plane of two lines Z, V to the point of intersection of Z, V 
without making use of the latter point. 

5. Using the definition in Ex. 3, show that if the point P' is on the polar p 
of a point P with respect to two lines Z, Z', then the point P is on the polar p' 
of P' with respect to Z, V. 

6. If the vertices A 19 A^ -4 3 , A of a simple plane quadrangle are respec- 
tively on the sides a z , a 3 , a s , a 4 of a simple plane quadrilateral, and if the inter- 
section of the pair of opposite sides A^A^ A%A is on the line joining the pair 
of opposite points a^, <z 2 a 8 , the remaining pair of opposite sides of the quad- 
rangle will meet on the line joining the remaining pair of opposite vertices of 
the quadrilateral. Dualize. 

* A. Cayley, Collected Works, Vol. I (1846), p. 317. G. Veronese, Mathema- 
tische Annalen, Vol. XIX (1882). Further references will be found in a paper by 
W. B. Carver, Transactions of the American Mathematical Society, Vol. VI (1905), 
p. 534. 



20] EXEECISES 53 

7. If two complete plane n-points A^ A 2 , - , A n and A{, A^ , A' n are 
so related that the side A^A^ and the remaining 2 (n 2) sides passing through 
^ and ^L 2 meet the corresponding sides of the other rc-point in points of a line Z, 
the remaining pairs of homologous sides of the two n -points meet on I and the 
two n-points are perspective from a point. Dualize. 

8. If five sides of a complete quadrangle A^A^A^A^ pass through five 
vertices of a complete quadrilateral a 1 a a a 8 a 4 in such a way that A^A^ is on 
<2 3 a 4 , A 2 A S on a 4 a 1? etc., then the sixth side of the quadrangle passes through 
the sixth vertex of the quadrilateral. Dualize. 

9. If on each of three concurrent lines a, &, c two points are given, A lt A 3 
on a; jB 1? B 2 on &; C v C 2 on c, there can be formed four pairs of triangles 
AfBjC L (i,j, k = 1, 2) and the pairs of corresponding sides meet in six points 
which are the vertices of a complete quadrilateral (Veronese, Atti dei Lincei, 
1876-1877, p. 649). 

10. With nine points situated in sets of three on three concurrent lines 
are formed 36 sets of three perspective triangles. For each set of three dis- 
tinct triangles the axes of perspectivity meet in a point; and the 36 points 
thus obtained from the 36 sets of triangles lie in sets of four on 27 lines, 

o/> q 

giving a configuration (Veronese, loc. cit.). 

11. A plane section of a 6-point in space can be considered as 3 triangles 
perspective in pairs from 3 collinear points with corresponding sides meeting 
in 3 collinear points. 

12. A plane section of a 6-point in space can be considered as 2 perspective 
complete quadrangles with corresponding sides meeting in the vertices of a 
complete quadrilateral. 

13. A plane section of an n-point in space gives the configuration * 




which may be considered (in n C n _ k ways) as a set of (n ) fc-points perspective 



in pairs from n _x,C a points, which form a configuration 
the points of intersection of corresponding sides form a coi 


_*A n fc 2 

& " n-*^ 


an< 


ifiguration 




/A fc-2 
3 k C s 



14. A plane section of a 7-point in space can be considered (in 120 ways) 
as composed of three simple heptagons (7-points) cyclically circumscribing 
each other. 

15. A plane section of an 11-point in space can be considered (in 19 ways) 
as composed of four 11-points cyclically circumscribing each other. 

16. A plane section of an n-point in space for n prime can be considered 

(in [n 2 ways) as n ~ simple n-points cyclically circumscribing each other. 

2 

* The symbol n C r is used to denote the number of combinations of n things 
taken r at a time. 



54 PROJECTION, SECTION, PERSPECTIVES [CHAP, n 

17. A plane section of a 6-point in space gives (in six ways) a 5-point whose 
sides pass through the points of a configuration 

18. A plane section of an n -point in space gives a complete (n 1) -point 
whose sides pass through the points of a configuration 





* 19. The n-space section of an ??2-point (m z. n + 2) in an (n + l)-space can be 
considered in the w-space as (m &) ^-points (in m (? m _ x ways) perspective in pairs 
from the vertices of the ?i-space section of one (m I')-point ; the r-spaces of 
the /j-point figures meet in (r 1) -spaces (r = l,2, ,w 1) which form the 
rc-space section of a -point. 

* 20. The figure of two perspective (n + l)-points in an w-space separates 
(in n + 3 ways) into two dual figures, respectively an (n + 2)-point circum- 
scribing the figure of (n + 2) (n l)-spaces. 

* 21. The section by a 3-space of an n-point in 4-space is a configuration 



M c, 


71-2 


-2<?2 


3 


M C 3 


n-3 


6 


4 


,A 



The plane section of this configuration is 




22. Let there be three points on each of two concurrent lines l v / 2 . The 
nine lines joining points of one set of three to points of the other determine 
six triangles whose vertices are not on / t or l z . The point of intersection of ^ 
and Z 2 has the same polar with regard to all six of these triangles. 

23. If two triangles are perspective, then are perspective also the two 
triangles whose vertices are points of intersection of each side of the given 
triangles with a line joining a fixed point of the axis of perspectivity to the 
opposite vertex. 

*24. Show that the configuration of the two perspective tetrahedra of 
Theorem 2 can be obtained as the section by a 3-space of a complete 6-point 
in a 4-space. 

*25. If two 5-points in a 4-space are perspective from a point, the corre- 
sponding edges meet in the vertices, the corresponding plane faces meet in the 
lines, and the corresponding 3-space faces in the planes of a complete 5-plane 
in a 3-space. 

* 26. If two (n -f l)-points in an ra-space are perspective from a point, 
their corresponding r-spaces meet in (r 1) -spaces which lie in the same 
(n l)-space (r=rl, 2 , 1) and form a complete configuration of 
(n + 1) (n 2) -spaces in (n 1) -space. 



CHAPTER III 

PROJECTIVITIES OF THE PRIMITIVE GEOMETRIC FORMS OF 
ONE, TWO, AND THREE DIMENSIONS 

21. The nine primitive geometric forms. 

DEFINITION. A. pencil of points DEFINITION. A. pencil of planes 
or a range is the figure formed by or an axial pencil * is the figure 
the set of all points on the same formed by the set of all planes on 
line. The line is called the axis the same line. The line is called 
of the pencil. the axis of the pencil. 

As indicated, the pencil of points is the space dual of the pencil 
of planes. ^/ 

DEFINITION. A pencil of lines or a flat pencil is the figure formed 
by the set of all lines which are at once on the same point and the 
same plane ; the point is called the vertex or center of the pencil 

The pencil of lines is clearly self-dual in space, while it is the 
plane dual of the pencil of points. The pencil of points, the pencil 
of lines, and the pencil of planes are called the primitive geometric 
forms of the first grade or of one dimension. 

DEFINITION. The following are known as the primitive geometric 
forms of the second grade or of two dimensions : 

The set of all points on a plane The set of all planes on a point 
is called a plane of points. The is called a bundle of planes. The 
set of all lines on a plane is called set of all lines on a point is called 
a plane of lines: The plane is a bundle of lines . The point is 
called the "base of the two forms, called the center of the bundles. 
The figure composed of a plane The figure composed of a bundle 
of points and a plane of lines of lines and a bundle of planes 
with the same base is called a with the same center is called 
planar field. simply a "bundle. 

DEFINITION. The set of all planes in space and the set of all points' 
in space are called the primitive geometric forms of the third grade 
or of three dimensions. 

* The pencil of planes is also called "by some writers a sheaf. 
55 



56 PRIMITIVE GEOMETRIC FORMS [CHAP, in 

There are then, all told, nine primitive geometric forms in a space 
of three dimensions.* 

22. Perspectivity and projectivity. In Chap. II, 13, we gave a 
definition of perspectivity. This definition we will now apply to the 
case of two primitive forms and will complete it where needed. We 
note first that, according to the definition referred to, two pencils of 
points in the same plane are perspective provided every two homol- 
ogous points of the pencils are on a line of a flat pencil, for they 
then have the same projection from a point. Two ^planes of points 
(lines) are perspective, if every two homologous elements are on a 
line (plane) of a bundle of lines (planes). Two pencils of lines in the 
same plane are perspective, if every two homologous lines intersect 
in a point of the same pencil of points. Two pencils of planes are 
perspective, if every two homologous planes are on a point of a pencil 
of points (they then have the same section by a line). Two bundles of 
lines (planes) are perspective, if every two homologous lines (planes) 
are on a point (line) of a plane of points (lines) (they then have the 
same section by a plane), etc. Our previous definition does not, how- 
ever, cover all possible eases. In the first place, it does not allow for 
the possibility of two forms of different kinds being perspective, such 
as a pencil of points and a pencil of lines, a plane of points and a 
bundle of lines, etc. This lack of completeness is removed for the 
case of one-dimensional forms by the following definition. It should 
be clearly noted that it is in complete agreement with the previous 
definition of perspectivity ; as far as one-dimensional forms are con- 
cerned it is wider in its application. 

DEFINITION. Two one-dimensional primitive forms of different kinds, 
not having a common axis, are perspective, if and only if they corre- 
spond in such a (1, 1) way that each element of one is on its homol- 
ogous element in the other ; two one-dimensional primitive forms of 
the same kind are perspective, if and only if every two homologous 
elements are on an element of a third one-dimensional form not 
having an axis in common with one of the given forms. If the third 
form is a pencil of lines with vertex P, the perspectivity is said to be 

* Some writers enumerate only six, by defining the set of all points and lines on 
a plane as a single form, and by regarding the set of all planes and lines at a point 
and the set of all points and planes in space each as a single form. We have fol- 
lowed the usage of Enriques, Vorlesungen iiber Projektive Geometric. 



22] PEESPECTIYITY 57 

central with center P; if the third form is a pencil of points or a pencil 
of planes with axis Z, the perspectivity is said to be axial with axis Z. 

As examples of this definition we mention the following: Two 
pencils of points on skew lines are perspective, if every two homol- 
ogous elements are on a plane of a pencil of planes ; two pencils of 
lines in different planes are perspective, if every two homologous 
lines are on a point of a pencil of points or a plane of a pencil of 
planes (either of the latter conditions is a consequence of the other) ; 
two pencils of planes are perspective, if every two homologous planes 
are on a point of a pencil of points or a line of a pencil of lines (in 
the latter case the axes of the pencils of planes are coplanar). A pen- 
cil of points and a pencil of lines are perspective, if every point is on 
its homologous line, etc. 

It is of great importance to note that our definitions of perspective 
primitive forms are dual throughout; ie. that if two forms are per- 
spective, the dual figure will consist of perspective forms. Hence any 
theorem proved concerning perspectivities can at once be dualized ; in 
particular, any theorem concerning the perspectivity of two forms of 
the same kind is true of any other two forms of the same kind. 

We use the notation [P] to denote a class of elements of any kind 
and denote individuals of the class by P alone or with an index or 
subscript. Thus two ranges of points may be denoted by [P] and [Q]. 
To indicate a perspective correspondence between them we write 



The same symbol, ^, is also used to indicate a perspectivity between 
any two one-dimensional forms. If the two forms' are of the same 
kind, it implies that there exists a third form such that every pair 
of homologous elements of the first two forms is on an element of 
the third form. The third form may also be exhibited in the notation 
by placing a symbol representing the third form immediately over 
the sign of perspectivity, ^. 
Thus the symbols 



denote that the range [P] is perspective by means of the center A with 
the range [Q], that each Q is on a line r of the flat pencil [>], and 
that the pencil [r] is perspective by; the axis a with the flat pencil [*]. 



58 PEIMITIVE GEOMETRIC FOEMS [CHAP, in 

A class of elements containing a finite number of elements can 
be indicated by the symbols for the several elements. When this 
notation is used, the symbol of perspectivity indicates that elements 
appearing in corresponding places in the two sequences of symbols 

are homologous. Thus 

123 4: = AJ3C D 

A 

implies that 1 and A, 2 and B, 3 and <7, 4 and D are homologous. 

DEFINITION.* Two one-dimensional primitive forms [cr] and [V] (of 
the same or different kinds) are said to be projective, provided there 
exists a sequence of forms [r], |V] ? , [r (n) ] such that 

[<r] == [r] = [r'] = -== [T ( >] = [>']. 
. LJ A LJ A L J A A L J A L J 

The correspondence thus established between [cr] and [V] is called 
a protective correspondence or projectivity , or also a protective trans- 
formation. Any element cr is said to be projected into its homologous 
element cr' by the sequence of perspectivities. 

Thus a projectivity is the resultant of a sequence of perspectivities. 
It is evident that [cr] and JV ] may be the same form, in which case 
the projectivity effects a permutation of the elements of the form. 
For example, it is proved later in this chapter that any four points 
A, B, C, D of a line can be projected into B, A } D, C respectively. 

A projectivity establishes a one-to-one correspondence between the 
elements of two one-dimensional forms, which correspondence we may 
consider abstractly without direct reference to the sequence of perspec- 
tivities by which it is defined. Such a correspondence we denote by 



Projectivities we will, in general, denote by letters of the Greek 
alphabet, such as TT. If a projectivity TT makes an element a- of a 
form homologous with an element a 1 of another or the same form, 
we will sometimes denote this by the relation ^(^)== .';,. Jn this 
case we may say the* projectivity transforms cr into <r'. Here the 
symbol TT( ) is used as a functional symbol "f acfiEg orTthe variable J 
cr, which represents any one of the elements of a given form. 

* This is Poncelet's definition of a projectivity. 
t Just like F(x), sin(a;), log(fc), etc. 

J The definition of variable is " a symbol x which represents any one of a class 
of elements [cc]." It is in this sense that we speak of " a variable point." 



23] PEOJECTIYITY 59 

23. The projectivity of one-dimensional primitive forms. The 

projectivity of one-dimensional primitive forms will be discussed 
with reference to the projectivity of pencils of points. The corre- 
sponding properties for the other one-dimensional primitive forms 
will then follow immediately by the theorems of duality (Theorems 
11-13, Chap. I). 

THEOREM 1. If A, B } C are three points of a line I and A 1 , B r , C f 
three points of another line V 9 then A can "be projected into A r , B into 
B f , and C into C 1 ly means of two centers of perspectivity. (The lines 
may be in the same or in different planes.) (A, E) 

Proof. If the points in any one of the pairs AA\ BE 1 , or CC' are 
coincident, one center is sufficient, viz., the intersection of the lines 
determined by the other 
two pairs. If each of these 
pairs consists of distinct 
points, let 8 be any point 
of the line AA r , distinct 
from A and A 1 (fig. 22). 
From 8 project A, B, C 
on any line V r distinct 
from I and V 9 but con- 
taining A f and a point 
of I If B", C" are the 
points of Z" correspond- 
ing to B, C respectively, 

the point of intersection S 1 of the lines B f B rf and C f O fr is the second 
center of perspectivity. This argument holds without modification, 
if one of the points A 9 B, C coincides with one of the points A r , B ! , O r 
other than its corresponding point. 

COROLLARY 1. IfA,B,C and A!, B ! 9 C 1 are on the same line, three 
centers of perspectivity are sufficient to project A, B, C into A 1 , B f , C f 
respectively. (A, E) 

COROLLARY 2. Any three distinct elements of a one-dimensional 
primitive form are protective with any three distinct elements of 
another or the same one-dimensional primitive form. (A, E) 

For, when the two forms are of the same kind, the result is ob- 
tained from the theorem and the first corollary directly from the 




60 PRIMITIVE GEOMETEIO FOKMS [CHAP.III 

theorems of duality (Theorems 1113, Chap. I). If they are of differ- 
ent kinds, a projection or section is sufficient to reduce them to the 
same kind. 

THEOREM 2. The projectivity ABCD -j^B ADC holds for any four 
distinct $>oints A, B, C, D of a line. (A, E) 

Proof. From a point S, not on the line I = AB, project ABCD into 
AB r C f D f on a line V through A and distinct from I (fig. 23). From D 
project AB'C f D f on the line SB. The last four points will then project 
into BADC by means of the center C f . In fig. 23 we have 

S D C r 

ABCD~=AB r C f D r = BB f C"S = BADC. 

A A A 

It is to be noted that a geometrical order of the points ABCD has no bearing 
on the theorem. In fact, the notion of such order has not yet been introduced 

into our geometry and, indeed, cannot 
be introduced on the basis of the 
present assumptions alone. The theo- 
rem merely states that the correspond- 
ence obtained ly interchanging any two 
of four collinear points and also inter- 
changing the remaining two is projectwe. 
The notion of order is, however, im- 
plied in our notation of projectivity 
and perspectivity. Thus, for example, 
we introduce the following definition : 

DEFINITION. Two ordered pairs of elements of any one-dimensional 
form are called a throw;"]! the pairs are AB, CD, this is denoted by 
T(AB, CD). Two throws are said to be equal, provided they are 
projective ; in symbols, T(AB, CD) = T (A f B f , C'D f ), provided we have 




The last theorem then states the equality of throws : 
J(AB, CD) = T(BA, DC) = T(CD, AB) = T(DC, 



The results of the last two theorems may be stated in the follow- 
ing form : 

THEOREM 1'. If 1 , %, 3 are elements of any one-dimensional prim- 
itive form, there exist projective transformations 'which will effect any 
one of the six permutations of these three elements. 



PEOJECTIYITY 



61 



THEOREM 2'. If 1, 2, 3, 4 are any four distinct elements of a one- 
dimensional primitive form, there exist ^rojectixe transformations 
which will transform 1234 into an y ne f ^ ie following permuta- 
tions of itself: 1234, 143, 3412, 4321. 

A projective transformation has been defined as the resultant of any 
sequence of perspectivities. We proceed now to the proof of a chain 
of theorems, "which lead to the fundamental result that any projective 
transformation between two distinct one-dimensional primitive forms 
of the same kind can be obtained as the resultant of two perspectivities. 

THEOREM 3. If [P], [P'], [P' f ] are pencils of points on three distinct 

S S 1 

concurrent lines I, V, l n respectively, sucli that [P] = [P f ] and [P r ] = 

S" 
"], then likewise [P] = [P"], and the three centers of perspectivity 



S, S f S" are collinear. (A, E) 




TIG-. 24 

Proof. Let be the common point of the lines 1 9 V, l n . 1S.P V P^P Z 
are three points of [P], and ^J^'J?' and JJ /7 J5"JJ f/ the corresponding 
points of [P'], [P n ] (fig. 24), it is clear that the triangles P^P^, 
P^PI, P^P^ !I are perspective from 0* By Desa-jg^e^Jbh^eorgm 
(Theorem 1, Chap. II) homologous sides of ^any pair, of these three 
triangles meet in ^nll jnaar ? ppiTrf^ Thp finn elusion of the theorem then 



follows readily from the hypotheses. 

* If the points in each of these sets of three are collinear, the theorem is obvious 
and the three centers of perspectivity coincide. 



62 



PRIMITIVE GEOMETEIC FOEMS 



[CHAP. Hi 



COROLLARY. If n concurrent lines l v l v Z 8 , 



S, 



n-l,n 

(A,E) 



, l n are connected by 
if I 



perspectivities L ^_, A L - 2J A L - aj A 
are distinct lines, then we have [J^] == | 

Proof. This follows almost immediately from the theorem, except 
when it happens that a set of four successive lines of the set Z 1 Z 2 Z S - l n 
are such that the first and third coincide and likewise the second and 
fourth. That this case forms no exception to the corollary may be 
shown as follows : Consider the perspectivities connecting the pencils 
of points on the lines l v Z 2 , Z 3 , Z 4 on the hypothesis that Z 1 = Z 8 , Z a = Z 4 
(fig. 25.) Let l v l z meet in <9, and let the line 3^S M meet Z x in A v 




and l z in -4 2 ; let A 3 = A l and J^ 4 be the corresponding points of Z 3 and 
Z 4 respectively. Further, let JS V B Z9 B and C v C z> <7 8 , (7 4 be any 
other two sequences of corresponding points in the perspectivities. 
Let S A be determined as the intersection of the lines A^A.^ and B^B^ 
The two quadrangles S 12 S^B 2 C 2 and $ 41 34 J? 4 <7 4 have five pairs of 
homologous sides meeting l t =l s in the points OA^B^B^Cy /Hence 
the side S^C^ meets ^ in C^ (Theorem 3, Chap. II). 

THEOREM 4. J/ K] K] [^] ^ re pencils of points on distinct 

cr Sf 

Kwe^ Z y Z 2 , Z respectively, such that K] = [^]==K], awd ^ [P'] is 
the pencil of points on any line V containing the intersection of l v I 
and also a point of 1 2 , "but not containing $ 2 , then there exists a point 

S f S 
v such that [JJ] = [P ! ] = [J>]. (A, E) 



on 



23] PEOJEGTIYITY 63 

Proof. Clearly we have 



But by tlie preceding theorem and the conditions on V we have 

# 
[-ZJ] = [P'], -where S a ' is a point of ^^ Hence we have 

K] = [^'] = K]- 

This theorem leads readily to the next theorem, which is the result 
toward which we have been working. We prove first the following 
lemmas : 

LEMMA 1. Any axial perspectivity between the points of two skeiv 
lines is equivalent to (and may le replaced ly) two central perspectives. 
(A,E) 

For let [P], [P'] be the pencils of points on the skew lines. Then 
if S and S r are any two points on the axis s of the axial perspectivity, 
the pencils of lines $[P], S r [P f ] * are so related that pairs of homol- 
ogous lines intersect in points of the line common to the planes of the 
two pencils [P] and S f [P f ], since each pair of homologous lines lie, 
by hypothesis, in a plane of the axial pencil s[P]~s[P f ]. 

LEMMA 2. Any projectivity "between pencils of points may be defined 
"by a sequence of central perspectivities. 

For any noncentral perspectivities occurring in the sequence defining 
a projectivity may, in consequence of Lemma 1, be replaced by sequences 
of central perspectivities. 

THEOREM 5. If two pencils of points [P] and [P ; ] on distinct lines 
are projective, there exists a pencil of points [Q] and two points S, S r 

S S f 
such that we haw [P] = [Q] = [P']. (A, E) 

Proof. By hypothesis and the two preceding lemmas we have a 
sequence of perspectivities 

& & S S S n 



* Given a class of elements [P]; the symbol S[P] is used to denote the class 
of elements SP determined by a given element S and any element of [P]. Hence, 
if [P] is a pencil of points and 8 a point not in [P], S [P] is a pencil of lines with 
center 8 ; if s is a line not on any P, s [P] is a pencil of planes with axis s. 



64 PEIMITIVE GEOMETEIC POEMS [CHAP, in 

We assume the number of these perspectivities to be greater than two, 
since otherwise the theorem is proved. By applying the corollary of 
Theorem 3, when necessary, this sequence of perspectivities may be 
so modified that no three successive axes are concurrent. We may 
also assume that no two of the axes Z, Z p Z 2 , Z s , , V of the pencils 
[P], [J], [JJ], [JJ], - [P'J are coincident; for Theorem 4 may evidently 
be used to replace any l t (= J t ) by a line ZJf (= Z t ). Now let l[ be the 
line joining the points ZZ t and Z 2 Z 8 , and let us suppose that it does not 
contain the center 2 (fig. 26). If then [2J 7 ] is the pencil of points 
on l[ t we may (by Theorem 4) replace the given sequence of per- 

orf or or cr 

spectivities by [P] == [P^] = [P 2 ] = [P s ] = and this sequence 

may in turn be replaced by 

t sy si s, 




(Theorem 3). If $ 2 is on the line 
joining ll : and Z 2 Z 3 , we may replace 
Z x by any line I" through the inter- 
l section of Z X Z 2 which meets Z and 
y IG t 26 does nofc contain the point 3 l (The- 

orem 4). The line joining Z 2 Z 3 to 

ZZ[' does not contain the point 3% which replaces 3 S . For, since ^ 2 is 
on the line joining Z 3 Z 2 to ll v the points Z 3 Z 2 and ZZ X are homologous 
points of the pencils [P 8 ] and [P] ; and if 3% were on the line join- 
ing Z 3 Z 2 to ll[ r , the point Z 3 Z 2 would also be homologous to ll[ f . We 
may then proceed as before. By repeated application of this process 
we can reduce the number of perspectivities one by one, until finally 
we obtain the pencil of points [Q] and the perspectivities 

S S r 



' As a consequence we have the important theorem : 

THEOEEM 6. Any two projective pencils of points on skew lines are 
axially perspective. (A, E) 

Proof. The axis of the perspectivity is the line 33 r of the last 
theorem. 

24. General theory of correspondence. Symbolic treatment. In 
preparation for a more detailed study of projective (and other) corre- 
spondences, we will now develop certain general ideas applicable to 



CORRESPONDENCE 65 

all one-to-one reciprocal correspondences as defined in Chap. II, 13, 
p. 35, and show in particular how these ideas may be conveniently 
represented in symbolic form.* As previously indicated (p. 58), we 
will represent such correspondences in general by the letters of the 
Greek alphabet, as A, B, F, . The totality of elements affected 
by the correspondences under consideration forms a system which we 
may denote by S. If, as a result of replacing every element of a system 
$! by the element homologous to it in a correspondence A, the sys- 
tem Sj is transformed into a system S 2 , we express this by the relation 
A(S 1 ) = S a . In particular, the element homologous with a given ele- 
ment P is represented by A (P). 

I. If two correspondences A, B are applied successively to a sys- 
tem Sj, so that we have A (S t ) = S 2 and B (S 2 ) = S 3 , the single corre- 
spondence F which transforms S x into S 3 is called the resultant or 
product of A by B; in symbols S s = B (S 2 ) = B (A(S 1 )) = BA (SJ, or, 
more briefly, B A = F. Similarly, for a succession of more than two 
correspondences. 

II. Two successions of correspondences A m A m ^i - A x and 
BgB^-L - B! have the same resultant, or their products are equal, 
provided they transform S into the same S'; in symbols, from the 

A W A W _ 1 - - A,(S) = B A.! - - - B,(S) 
follows A^A^ - . - A x = B ff B^. . . B r 



III. The correspondence which makes every element of the sys- 
tem correspond to itself is called the identical correspondence or simply 
the identity, and is denoted by the symbol 1. It is then readily seen 
that for any correspondence A we have the relations 



IV. If a correspondence A transforms a system S x into S 2 , the corre- 
spondence which transforms S 3 into S x is called the inverse of A and is 
represented by A" 1 ; i.e. if we have A (S x ) = S 2 , then also A"" 1 (S 2 ) = S r 
The inverse of the inverse of A is then clearly A, and we evidently 
have also the relations 

AA- I =AT I A=I. 

* In this section we have followed to a considerable extent the treatment given 
"by H. Wiener, Berichte der K. sachsischen Gesellschaft der Wissenschaften, Leipzig. 
Vol. XLH (1890), pp. 249-252. 



66 PEIMITIYE GEOMETBIC FOBMS [CHAP. Ill 

Conversely, if A, A' are two correspondences such that we have 
AA' = 1, then A' is the inverse of A. Evidently the identity is its 
own inverse. 

V. The product of three correspondences A, B, F always satisfies 
the relation (FB) A = F (BA) (the associative law). For from the 
relations A(S 1 )=S 2 , B(S 2 )=S 3 , F(S 3 )=S 4 foUows at once BA(S 1 )=S 3 , 
whence F(BA) (S~) =S 4 ;" and also FB (S 2 ) = S 4 , and hence (FB) A (SJ 
= S 4 , which proves the relation in question. More generally, in any 
product of correspondences any set of successive correspondences may 
be inclosed in parentheses (provided their order be left unchanged), 
or any pair of parentheses may be removed; in other words, in a 
product of correspondences any set of successive correspondences may 
be replaced by their resultant, or any correspondence may be replaced 
by a succession of which the given correspondence is the resultant. 

VI. In particular, we may conclude from the above that the inverse 
of the product M - - BA is A" 1 B"" 1 M" 1 , since we evidently have 
the relation M - - - BAA^B" 1 . . . M"^! (cf. IV). 

VII. Further, it is easy to show that from two relations A = B and 
F = A follows AF = BA and FA = AB. In particular, the relation 
A = B may also be written AB~ x = 1, B~ x A = 1, B A~ l = 1, or A~ *B = 1. 

VIII. Two correspondences A and B are said to be commutative 
if they satisfy the relation BA = AB. 

IX. If a correspondence A is repeated n times, the resultant is writ- 
ten AA A = A n . A correspondence A is said to be of period n, if n 
is the smallest positive integer for which the relation A n = 1 is satisfied. 
When no such integer exists, the correspondence has no period ; when 
it does exist, the correspondence is said to be periodic or cyclic. 

X. The case n = 2 is of particular importance. A correspondence 
of period two is called involutoric or reflexive. 

25. The notion of a group. At this point it seems desirable to 
introduce the notion of a group of correspondences, which is funda- 
mental in any system of geometry. We will give the general abstract 
definition of a group as follows : * 

DEFINITION. A class G of elements, which we denote by a, &, 
c, -, is said to form a group with respect to an operation or law of 

* We have used here substantially the definition of a group given by L. E. Dickson, 
Definitions of a Group and a Pield by Independent Postulates, Transactions of the 
American Mathematical Society, Vol. VI (1905), p. 199. 



25,26] GROUPS 67 

combination o, acting on pairs of elements of G, provided the fol- 
lowing postulates are satisfied : 

G 1, For every pair of (equal or distinct) elements a, b of G, the 
result a o & of acting with the operation o on the pair in the order 
given * is a uniquely determined element of G. 

G 2. The relation (a o &) o c = a o (b o c) holds for any three (equal or 
distinct] elements a, 1), c of G. 

G3. There occurs in G an element i, such that the relation aoi = a 
holds for ever?/ clement a of G. 

G4. For every element a in G there exists an element a 1 satisfying 
the relation a o a'= i. 

From the above set of postulates follow, as theorems, the following : 

The relations a o a r = i and a o i = a im%)ly respectively the relations 
a'o a i and ioa~ a. 

An element i of G is called an identity element, and an element a 1 
satisfying the relation a o a 1 = i is called an inrerse element of a. 

There is only one identity element in G. 

For every element a of & there is only one inverse. 

We omit the proofs of these theorems. 

DEFINITION. A group which satisfies further the following postulate 
is said to be commutative (or dbelian) : 

G 5. TJie relation a o 5 = I o a is satisfied for every pair of de- 
ments a, b in G. 

26. Groups of correspondences. Invariant elements and figures. 

The developments of the last two sections lead now immediately 
to the theorem: 

A set of correspondences forms a group provided the set contains 
the inverse of any correspondence in the set and provided the resultant 
of any two correspondences is in the set 

Here the law of combination o of the preceding section is simply 
the formation of the resultant of two successive correspondences. 

DEFINITION. If a correspondence A transforms every element of a 
given figure F into an element of the same figure, the figure F is said 
to be invariant under A, or to be left invariant by A. In particular, 

* I.e. ao 5 and & o a are not necessarily identical. The operation o simply defines 
a correspondence, whereby to every pair of elements a, 6 in G in a given order corre- 
sponds a unique element j this element is denoted by a o &. 



68 PEIMITIVE G-EOMETEIC FOEMS [CHAP, in 

an element which is transformed into itself by A is said to be an 
invariant element of A; the latter is also sometimes called a double 
element or a fixed element (point, line, plane, etc.). 

We now call attention to the following general principle : 
The set of all correspondences in a group G which learn a given 
figure invariant forms a group. 

This follows at once from the fact that if each of two corre- 
spondences of G leaves the figure invariant, their product and their 
inverses will likewise leave it invariant ; and these are all in G, since, 
by hypothesis, G is a group. It may happen, of course, that a group 
defined in this way consists of the identity only. 

These notions are illustrated in the following section : 
27. Group properties of projectivities. From the definition of a pro- 
jectivity between one-dimensional forms follows at once 

THEOREM 7. The inverse of any projectivity and the resultant of 
any two projectivities are projectivities. 

On the other hand, we notice that the resultant of two perspec- 
tivities is not, in general, a perspectivity ; if, however, two perspec- 
tivities connect three concurrent lines, as in Theorem 3, their resultant 
is a perspectivity. A perspectivity is its own inverse, and is therefore 
reflexive. As an example of the general principle of 26, we have 
the important result : 

THEOREM 8. The set of all projectivities leaving a given pencil of 
points invariant form a group. 

If the number of points in such a pencil is unlimited, this group con- 
tains an unlimited number of projectivities. It is called the general 
projective group on the line. Likewise, the set of all projectivities on a 
line leaving the figure formed by three distinct points invariant forms a 
subgroup of the general group on the line. If we assume that each per- 
mutation (cf. Theorem I/) of the three points gives rise to only a single 
projectivity (the proof of which requires an additional assumption), 
this subgroup consists of six projectivities (including, of course, the 
identity). Again, the set of all projectivities on a line leaving each of two 
given distinct points invariant forms a subgroup of the general group. 

We will close this section with two examples illustrative of the 
principles now under discussion, in which the projectivities in ques- 
tion are given by explicit constructions. 



27] G-BOUP OF PEOJECTIVITIES 69 

EXAMPLE 1. A group of projectivities leaving eacli of two given 
points invariant. Let M 9 N be two distinct points on a line I, and 
let m, n be any two lines through 2I 9 N respectively and coplanar 
with I (fig. 27). On m let there be an arbitrary given point S. If ^ 
is any other point on m distinct from J/ ; the points 8, S l together 
with the line n define a projectivity TT^ on I as follows : The point 
TTj (A) = A r homologous to any point A of I is obtained by the two 

S 8 l 
perspectivities [A] = [^i]~[-4/], where [A^ is the pencil of points 

on n. Every point jS % then, distinct from J/, defines a unique pro- 
jectivity 7r 4 ; we are to show that the set of all these projectivities TT, 
forms a group. We show first that the product 
of any two TT V 7r 2 is a uniquely determined pro- 
jectivity 7T 3 of the set (fig. 27). 
In the figure, A! = TT^ (A) 
and A n = Tr z (A f ) have been 




M B B' B" A A' A" N 



constructed. The point S s giving A n directly from A by a similar con- 
struction is then uniquely determined as the intersection of the lines 
A n A v m. Let B be any other point of I distinct from M 9 N 9 and let 
J5'= ir^(S) and J?*= ^(B 1 ) be constructed ; we must show that we have 
B"=z*jr z (B). We recognize the quadrangular set Q(If3'A r , NA"B n ) as 
defined by the quadrangle SS 2 B Z A Q . But of this quadrangular set all 
points except B n are also obtained from the quadrangle S^B^A^ 
whence the line S^ determines the point B n (Theorem 3, Chap. II). 
It is necessary further to show that the inverse of any projectivity in 
the set is in the set. For this purpose we need simply determine $ 2 
as the intersection of the line AA Z with m and repeat the former argu- 
ment. This is left as an exercise. Finally, the identity is in the set, 
since it is TT V when S^S. 



70 PRIMITIVE GEOMETRIC FORMS [CHAP.III 

It is to be noted that in this example the points Jf and N are 
double points of each projectivity in the group; and also that if P, P r 
and Q, Q r are any two pairs of homologous points of a projectivity 
we have Q (MPQ, NQ'P ! ). Moreover, it is clear that any projectivity, 
of the group is uniquely determined by a pair of homologous elements, 
and that there exists a projectivity which 
will transform any point A of I into any 
other point B of I, provided only that 
A and B are distinct from 
M and N. By virtue of 
the latter property the 
group is said 
to be transitive. 




A A' A/ A" 

FIG. 28 

EXAMPLE 2. Commutative projectivities. Let M be a point of a 
line Z, and let m, m f be any two lines through H distinct from I, but 
in the same plane with I (fig. 28.) Let 8 be a given point of m } and 
let a projectivity TT I be defined by another point S l of m which deter- 

*? a 

mines the perspectivities [A] = [-4 X ] == [A r ] 9 where [^ x ] is the pencil 

of points on m f . Any two projectivities defined in this way ly points S l 
are commutative. Let 7r 2 be another such projectivity, and construct 
the points A r =^(A) 9 A' f ='jr 2 (A r ) 9 and A[ = Tr z (A). The quadrangle 
JSS^A^A^ gives Q(MAA r , ITA ff A[)- and the quadrangular set determined 
on I by the quadrangle SS^A^ has the first five points of the former 
in the same positions in the symbols. Hence we have vr^A^) = A rr , and 
therefore TT^ = TT^. 

EXERCISES 

1. Show that the set of all projectivities TT^ of Example 2 above forms a 
group, which is then a commutative group. 

2. Show that the projectivity wj of Example 1 above is identical with the 
projectivity obtained by choosing any other two points of m as centers of 
perspectivity, provided only that the two projectivities have one homologous 



27,28] TWO-DBmNSIOjSTAL PEOJEOTIYITIES 71 

pair (distinct from M or 2V) in common. Investigate the general question as 
to how far the construction may be modified so as still to preserve the propo- 
sition that the projectivities are determined by the double points J/, -V and 
one pair of homologous elements. 

3. Discuss the same general question for the projectivities of Example 2. 

4. Apply the method of Example 2 to the projectivities of Example 1. 
"Why does it fail to show that any two of the latter are commutative ? State 
the space and plane duals of the two examples. 

5. ABCD is a tetrahedron and a, ^8, y, S the faces not containing A ,B,C,D 
respectively, and I is any line not meeting an edge. The planes (I A , IB, 1C, ID) 
are projective with the points (la, 1/3, ly, /S). 

6. On each of the ten sides of a complete 5-point in a plane there are three 
diagonal points and two vertices. Write down the projectivities among these 
ten sets of five points each. 

28. Projective transformations of two-dimensional forms. 

DEFINITION. A projective transformation between the elements of 
two two-dimensional or two three-dimensional forms is any one-to- 
one reciprocal correspondence between the elements of the two forms, 
such that to every one-dimensional form of one there corresponds 
a projective one-dimensional form of the other. 

DEFINITION. A collineation is any (1, 1) correspondence between 

-"^j^a^iEWM*--* J \ / JT 

two two-dimensionaior two three-dimensional forms in which to every 
element of one of the forms corresponds an element of the same kind 
in the other form, and in which to every one-dimensional form of one 
corresponds a one-dimensional form of the other. A. projective, colline- 
ation is one in which this correspondence is projective. Unless other- 
wise specified, the term collineation will, in the future, always denote 
a projective collineation.* 

In the present chapter we shall confine ourselves to the discus- 
sion of some of the fundamental properties of collineations. In this 
section we discuss the collineations between two-dimensional forms, 
and shall take the plane (planar field) as typical; the corresponding 
theorems for the other two-dimensional forms will then follow from 
duality. 

The simplest correspondence between the elements of two distinct 
pla r nes TT, ir 1 is a perspective correspondence, whereby any two homol- 
ogous elements are on the same element of a bundle whose center 
is on neither of the planes TT, TT'. The simplest collineation in a plane, 

* In how far a collineation must be projective will appear later. 



72 PRIMITIVE GEOMETEIC FOEMS [CHAP.III 

ie. which transforms every element of a plane into an element of the 
same plane, is the following : 

DEFINITION. A perspectwe^ collineation in a plane is a protective 
collineation leaving invariant every point on a given line o and every, 
line on a given point 0. The line o and the point are called the 
axis and center respectively of the perspective collineation. If the 
center and axis are not united, the collineation is called a planar 
homology; if they are united, a planar elation. 

A. perspective collineation in a plane TT may be constructed as 
follows : Let any line o and any point of TT "be chosen as axis and 
center respectively, and let T^ be any plane through o distinct from TT. 
Let O v 2 be any two points collinear with and in neither of the 
planes TT, 7r r The perspective collineation is then obtained by the 

O l 0* 

two perspectivities [P] === [PJ = [P'], where P is any point of TT and 

.ZJ, P r are points of TT I and TT respectively. Every point of the line o 
and every line through the point clearly remain fixed by the trans- 
formation, so that the conditions of the definition are satisfied, if 
only the transformation is projeetive. But it is readily seen that 
every pencil of points is transformed by this process into a perspec- 
tive pencil of points, the center of perspectivity being the point 0\ 
and every pencil of lines is transformed into a perspective pencil, the 
axis of perspectivity being o. The above discussion applies whether 

or not the point is on the line o, 

THEOREM 9. A perspective col- 
lineation in a plane is uniquely 
defined if the center, axis, and any 
two homologous points (not on the 
axis or center) are given, with the 
single restriction that the homol- 
ogous points must le collinear 
with 0. (A, E) 

Proof. Let 0, o be the center and axis respectively (fig. 29). It is 
clear from the definition that any two homologous points must be 
collinear with 0, since every line through. is invariant ; similarly 
(dually) any two homologous lines must be concurrent with o. Let 
A, A 1 be the given pair of homologous points collinear with 0. The 




28] TWO-DIMEXSIO-X T AL PEOJECTITITIES 73 

point B f homologous to any point B of the plane is then determined. 
We may assume 3 to be distinct from 0, A, A 1 and not to be on o. 
B ! is on the line OB, and if the line AB meets o in C, then, since C 
is invariant by definition, the line AB = AC is transformed into A r C. 
B f is then determined as the intersection of the lines OB and A'C. 
This applies unless B is on the line A A'} in this case we determine 
as above a pair of homologous points not on AA' 9 and then use the 
two points thus determined to construct B 1 . This shows that there 
can be no more than one perspective collineation in the plane with 
the given elements. 

To show that there is one we may proceed as follows : Let ^ be 
any plane through o distinct from TT, the plane of the perspectivity, 
and let O t be any point on neither of the planes TT, TT I . If the line A0 1 
meets ^ in A v the line A ? A^ meets 00^ in a point 2 . The perspec- 
tive collineation determined by the two centers of perspectivity O v 
and the plane ^ then has 0, o as center and axis respectively and A 9 A r 
as a pair of homologous points. 

COROLLARY 1. A perspective collineation in a plane transforms every 
one-dimensional form into a perspective one-dimensional form. (A, E) 

COROLLARY 2. A perspective collineation with center and axis o 
transforms any triangle none of whose vertices or sides are on o *or 
into a perspective triangle, the center of perspectivity of the triangles 
being the center of the collineation and the asois of perspectivity being 
the axis of the collineation. (A, E) 

COROLLARY 3. The only planar collineations (whether required to 
be protective or not] which leave invariant the points of a line o and 
the lines through a point are homologies if is not on o, and 
elations if is on o. ,(A, E) 

Proof. This will be evident on observing that in the first paragraph 
of the proof of the theorem no use is made of the hypothesis that the 
collineation is protective. 

COROLLARY 4. If H is a perspective collineation such that H (0) = 0, 
H(o) = o, E(A) A r 9 H(B) = B f where A, A r , B, B J are collinear with 
a point K of o, then we have Q(OAB, KB'A 1 ). (A, E) 

Proof. If C is any point not on AJ! and H(<7) = C", the lines AC 
and A r C r meet in a point L of o, and BO and B f C r meet in a point M 
of o\ and the required quadrangle is CC'LM (cf. fig. 32, p. 77). 



74 



PRIMITIVE GEOMETRIC FORMS [CHAP, m 



THEOREM 10. Any complete quadrangle of a plane can le trans- 
formed into any complete quadrangle of the same or a different plane 
ly a projectile collineation which, if the quadrangles are in the same 
plane, is the resultant of a finite numler of perspective collineations. 
(A,E) 

Proof. Let the quadrangles be in the same plane and let their ver- 
tices be A, B, 0, D and A', B\ C", D J respectively. We show first that 
there exists a collineation leaving any three vertices, say A', B 1 , C f , of 




, yl / / v 

>\/Vi \ 

' k / 



/ 

/ / 




A' 



T 



0* 



Pio. 80 



the quadrangle A f B ! C f D r invariant and transforming into the fourth, 
D f , any other point Z> 8 not on a side of the triangle A r $ r C r (Gg- 30). Let 
D be the intersection of ^D 3 , ffD 1 and consider the homology with 
center A 1 and ajds JS f C f transforming D 3 into D. Next consider the 
homology with center B 1 and axis C f A r transforming ^into D f . Both 
these homologies exist by Theorem 9. The resultant of these two 
homologies is a eollineation leaving fixed A 1 , B r , C 1 and transforming 
Z> 3 into D f . (It should be noticed that one or both of the homologies 
may be the identity.) 

Let O t be any point on the line containing A and A r and let o be 
any line not passing through A or A*. By Theorem 9 there exists a 



28,29] THEBB-DIMEirSIOSrAL PBOJECTIVITIES 75 

perspective collineation ^ transforming A to A* and having O l and t 
as center and axis. Let B v C v D^ be points such that 



In like manner, let o z be any line through A f not containing B^ or 
B r and let 2 be any point on the line B^B'. Let 7r 3 be the perspec- 
tive collineation with axis 2 , center 2 , and transforming B : to B*. 
Let <7 = 7 and D= w). Here" 



3 



Now let 3 be any point on the line C 2 6 f/ and let 7r 8 be the per- 
spective collineation which has A f B r =o s for axis, 3 for center, and 
transforms <7 2 to <7 ; . The existence of 7r g follows from Theorem 9 as 
soon as we observe that C f is not on the line A f B f , by hypothesis, 
and 6 2 is not on A ! B ! \ because if so, C^ would be on A r B and there- 
fore C would' be on AB. Let 7r 3 (D 2 ) = D 3 . It follows that 



The point D 3 cannot be on a side of the triangle A f B f C f because 
then D 2 would be on a side t of A ! B r C 2 , and hence D^ on a side of 
A r B : G v and, finally, D on a side of -ABC. Hence, by the first para- 
graph of this proof, there exists a projectivity 7r 4 such that 

n-t(A r B r C'Z> 9 ) = A'B'C'D*. 

The resultant tr^ir^ir^rr^ of these four collineations clearly transforms 
A, B, C, D into A f 9 B r , C f , D ! respectively. If the quadrangles are in 
different planes, we need only add a perspective transformation between 
the two planes. 

COROLLAEY. There exist projectile collineations in a plane which 
will effect any one of the possible % permutations of the vertices of 
a complete quadrangle in the plane. (A, E) 

29. Projective collineations of three-dimensional forms. Protective 
collineations in a three-dimensional form have been defined at the 
beginning of 28. 

DEFINITION. A projective collineation in space which leaves inva- 
riant every point of a plane a> and every plane on a point is called a 
perspective collineation. The plane a> is called the plane ofperspectivity; 
the point is called the center. If is on a>, the collineation is said 
to be an elation in space ; otherwise, a homology in space. 



76 PEIMITIYE GBOMETEIC FOKMS [CHAP.III 

THEOREM 11. I/O is any point and a any plane, there exists one 
and only one perspective collineation in space having 0, co for center 
and plane of per spectivity respectively, which transforms any point A 
(distinct from and not on co) into any other point A! (distinct from 
and not on co) collinear with AO. (A, E) 

Proof. We show first that there cannot be more than one per- 
spective collineation satisfying the conditions of the theorem, by 
showing that the point B f homologous to any point B is uniquely 




determined by the given conditions. We may assume B not on a 
and distinct from and A. Suppose first that B is not on the line 
AO (fig. 31). Since BO is an invariant line, B 1 is on B0\ and if 
the line AB meets co in L, the line AB = AL is transformed into 
the line A ! L. Hence B r is determined as the intersection of BO 
and A r L. There remains the case where B is on AO and distinct 
from A and (fig. 32). Let 0, C 1 be any pair of homologous points 
not on AO, and let AC and BC meet to in L and M respectively. 
The line MB = M is transformed into M C r 9 and the point B J is then 
determined as the intersection of the lines BO and MC f . That this 
point is independent of the choice of the pair C, C J now follows 
from the fact that the quadrangle MLCC 1 gives the quadrangular 
set Q(KAA f , OB r B), where K is the point in which AO meets co 
(K may coincide with without affecting the argument). The point 
B* is then uniquely determined by the five points 0, E, A, A 1 , B. 
The correspondence defined by the construction in the paragraph 
above has been proved to be one-to-one throughout. On the line AO 
it is projective because of the perspectivities (fig. 32) 



29] THBEE-DIMENSIO]S T AL PEOJECTIYITIES 77 

C ' 



On OB, any other line through 0, it is projeetive because of the per- 
spectivities (fig. 31) , ^ 

w=w= 

That any pencil of points not through is transformed into a 
perspective pencil, the center of perspectivity being 0, is now easily 
seen and is left as an exercise for the reader. From this it follows 




PIG. 32 

that any one-dimensional form is transformed into a projeetive form, 
so that the correspondence which has been constructed satisfies the 
definition of a projeetive collineation. 

THEOREM 12. Any complete five-point in space can le transformed 
into any other complete Jive-point in space ly a projeetive collineation 
which is the resultant of a finite number of perspective collineations. (A,E) 

Proof. Let the five-points be ABODE and AWC'D'E* respectively. 
We will show first that there exists a collineation leaving A f f C f D f 
invariant and transforming into JS f any point 2S Q not coplanar with 
three of the points A f J3 r C f Z> f . Consider a homology having A f B ! C l as 
plane of perspeetivity and D 1 as center. Any such homology trans- 
forms J into a point on the line JS Q I> f . Similarly, a homology with 
plane A f f l> ! and center C r transforms Iff into a point on the line HfC r . 
If E Q D ! and S r O f intersect in a point JS V the resultant of two homol- 
ogies of the kind described, of which the first transforms H into JS^ 
and the second transforms JE^ into Jff, leaves A f ffC f Z> f invariant and 
transforms JH Q into If. If the lines E^D* and JS f C r are skew, there 
is a line through B 1 meeting the lines E^D 1 and E f C r respectively 



78 PEIMITIVE GEOMETEIC POEMS [CHAP, in 

in two points E^ and JS 2 . The resultant of the three homologies, of 
which the first has the plane A*'C f and center D f and transforms 
JSo to JE V of which the second has the plane A f C f D f and center B 1 
and transforms E l to JE Z , and of which the third has the plane A f 'D r 
and center C r and transforms J 2 to E ! , is a collineation leaving A J 3 f C 1 D t 
invariant and transforming E^ to E 1 . The remainder of the proof is 
now entirely analogous to the proof of Theorem 10. The details are 
left as an exercise. 

COROLLARY. There exist projectile collineations which, will effect 
any one of the possible 120 permutations of the vertices of a complete 
five-point in space. (A, E) 

' EXERCISES 

1. Prove the existence of perspective collineations in a plane without 
making use of any points outside the plane. 

2. Discuss the figure formed by two triangles which are homologous 
under an elation. How is this special form of the Desargues configuration 
obtained as a section of a complete five-point in space? 

3. Given an elation in a plane with center and axis o and two homol- 
ogous pairs A , A' and B, B' on any line through 0, show that we always 
have Q(OAA', OB'B). 

4. What permutations of the vertices of a complete quadrangle leave a 
given diagonal point invariant ? every diagonal point ? 

5. Write down the permutations of the six sides of a complete quadrangle 
brought about by all possible permutations of the vertices. 

6. The set of all homologies (elations) in a plane with the same center 
and axis form a group. 

7. Prove that two elations in a plane having a common axis and center 
are commutative. Will this method apply to prove that two homologies with 
common axis and center are commutative? 

8. Prove that two elations in a plane having a common axis are commu- 
tative. Dualize. Prove the corresponding theorem in space. 

9. Prove that the resultant of two elations having a common axis is an 
elation. Dualize. Prove the corresponding theorem in space. What groups 
of elations are defined by these theorems ? 

10. Discuss the effect of a perspective collineation of space on : (1) a pencil 
of lines ; (2) any plane ; (3) any bundle of lines j (4) a tetrahedron ; (5) a 
complete five-point in space. 

11. The set of all collineations in space (in a plane) form a group. 

12. The set of all projective collineations in space (in a plane) form a group. 

13. Show that under certain conditions the configuration of two perspective 
tetrahedra is left invariant by 120 collineations (cf. Ex. 3, p. 47). 



CHAPTER IV 



HARMOITCC CONSTRUCTIONS AM) THE FUNDAMENTAL THEOREM 
OF PROTECTIVE GEOMETRY 

30. The projectivity of quadrangular sets. We return now to a 
more detailed discussion of the notion of quadrangular sets introduced 
at the end of Chap. II. We there defined a quadrangular set of points 
as the section by a transversal of the sides of a complete quadrangle ; 
the plane dual of this figure we call a quadrangular set of lines;* 
it consists of the projection of the vertices of a complete quadrilateral 
from a point which is in the plane of the quadrilateral, but not on 
any of its sides ; the space dual of a quadrangular set of points we 
call a quadrangular set of planes; it is the figure formed by the 
projection from a point of the 
figure of a quadrangular set 
of lines. We may now prove 
the following im- 
portant theorem : 

THEOREM 1. 
TJie section by a 
transversal of a 
quadrangular 
set of lines is a 
quadrangular 
set of points. 
(A,E) 

Proof. By Theorem 3', Chap. II, p. 49, and the dual of Note 2, on 
p. 48, we may take the transversal I to be one of the sides of a com- 
plete quadrilateral the projection of whose vertices from a point P 
forms the set of lines in question (fig. 33). Let the remaining three 
sides of such a quadrilateral be a, &, <?. Let the points Jo> ca, and al 

* It would be more natural at this stage to call such a set a quadrilateral set of 
lines; the next theorem, however, justifies the term we have chosen, which has the 
advantage of uniformity. 

79 




FIG. 33 



80 



THE FUNDAMENTAL THEOEEM 



[CHAP. IV 



be denoted by A, B, and C respectively. The sides of the quadrangle 
PABCra&et I in the same points as the lines of the quadrangular set 
of lines. 

COROLLARY. A set of collinear points which is projective with a 
quadrangular set is a quadrangular set. (A, E) 

THEOKEM I/. The projection from a point of a quadrangular set of 
points is a quadrangular set of lines. (A, E) 

This is the plane dual of the preceding ; the space dual is : 

THEOREM I/ 7 . The section ly a plane of a quadrangular set of planes 
is a quadrangular set of lines. (A, E) 

COROLLARY. If a set of elements of a primitive one-dimensional form 
is protective with a quadrangular set, it is itself a quadrangular set. 
(A,E) 

31. Harmonic sets. DEFINITION. A quadrangular set Q (123, 124) 
is called a harmonic set and is denoted by H (12, 34). The elements 
3, 4 are called harmonic conjugates with respect to the elements 1, 2 ; 
and 3 (or 4) is called the harmonic conjugate of 4 (or 3) with respect 
to 1 and 2. 

From this definition we see that in a harmonic set of points 
\\(AC, BD), the points A and C are diagonal points of a complete 




FIG. 34 



quadrangle^ while the points B and D are the intersections of the 
remaining two opposite sides of the quadrangle with the line AC 
(fig. 34). Likewise, in a harmonic set of lines H (ac, Id), the lines a 
and c are two diagonal lines of a complete quadrilateral, while the 



31] HAEMO]SriC SETS 81 

lines 5 and d are the lines joining the remaining pair of opposite 
vertices of the quadrilateral to the point of intersection ac of the 
lines a and c (fig. 35). A harmonic set of planes is the space dual 
of a harmonic set of points, and is therefore the projection from a 
point of a harmonic set of lines. 

In case the diagonal points of a complete quadrangle are collinear, any 
three points of a line form a harmonic set and any point is its own harmonic 
conjugate with regard to any two points collinear with it. Theorems on har- 
monic sets are therefore trivial in those spaces for which Assumption H is 
not true. We shall therefore base our reasoning, in this and the following 
two sections, on Assumption H Q J though most of the theorems are obviously 
true also in case H is false. This is why some of the theorems are labeled as 
dependent on Assumptions A and E, whereas the proofs given involve H Q also. 

The corollary of Theorem 3, Chap. II, when applied to harmonic 
sets yields the following: 

THEOREM 2. The harmbnic conjugate of an element with respect to 
two oilier elements of a one-dimensional primitive form is a unique 
element of the form. (A, E) 

Theorem 1 applied to the special case of harmonic sets gives 

THEOREM 3. Any section or projection of a harmonic set is a 
harmonic set. (A, E) 

COROLLARY. If a set of four elements of any one-dimensional prim- 
itive form is projectile with a harmonic set, it is itself a harmonic set. 
(A,E) 

THEOREM 4. If 1 and 2 are harmonic conjugates with respect to 
3 and 4, 3 and 4 are harmonic conjugates with respect to 1 and 2. 
(A, E, H ) 

Proof. By Theorem 2, Chap. Ill, there exists a projectivity 

1234^-3412. 

But by hypothesis we have H(34, 12). Hence by the corollary of 
Theorem 3 we have H (12, 34). 

By virtue of this theorem the pairs 1, 2 and 3, 4 in the expression 
H (12, 34) play the same r81e and may be interchanged,* 

*The corresponding theorem for the more general expression Q(12S, 456) 
cannot be derived without the use of an additional assumption (cf . Theorem 24, 
Chap. IV). 



82 THE FUNDAMENTAL THEOBEM [CHAP.IV 

THEOREMS. Given two harmonic sets H(12, 34) and H(1'2', 3'4'), 
there exists a projectivity such that 1234^ 1'2'3'4'. (A, E) 

Proof. Any projectivity 123 ^ 1'2'3' (Theorem 1, Chap. Ill) must 
transform 4 into 4' by "virtue of Theorem 3, Cor., and the fact that 
the harmonic conjugate of 3 with respect to 1 and 2 is unique (Theo- 
rem 2). This is the converse of Theorem 3, Cor. 

COROLLABY 1. If H (12, 34) and H (12', 3'4') are two harmonic sets 
of different one-dimensional forms having the element 1 in common, 
we ham 1234 = 12'3'4'. (A, E) 

For under the hypotheses of the corollary the projectivity 123-^ 1'2'3' 
of the preceding proof may be replaced by the perspectivity 123 == 12'3'. 

COEOLLAEY 2. If H (12, 3 4) is a harmonic set, there exists a projec- 
tivity 1234^-1243. (A,E) 

This follows directly from the last theorem and the evident fact 
that if H (12, 34) we have also H (12, 43). The converse of this 
corollary is likewise valid ; the proof, however, is given later in this 
chapter (ef. Theorem .27, Cor. 5). 

We see as a result of the last corollary and Theorem 2, Chap. Ill, 
that if we have H(12, 34), there exist projectivities which will trans- 
form 1234 into any one of the eight permutations 

1234, 1243, 2134, 2143, 3412, 3421, 4312, 4321.* 

In other words, if we have H(12, 34), we have likewise H(12, 43), 
H(21, 34), H(21, 43), H(34, 12), H(34, 21), H(43, 12), H(43, 21). 

THEOREM 6. The two sides of a complete quadrangle which meet in 
a diagonal point are harmonic conjtyates with respect to the two sides 
of the diagonal triangle which meet in this point. (A, E) 

Proof. The four sides of the complete quadrangle which do not 
pass through the diagonal point in question form a quadrilateral 
which defines the set of four lines mentioned as harmonic in the 
way indicated (fig. 36). 

It is sometimes convenient to speak of a pair of elements of a 
form as harmonic with a pair of elements of a form of different ^ 
kind. For example, we may say that two points are harmonic with 
two lines in a plane with the points, if the points determine two 

* These transformations form the so-called eigM-group. 



31] 



HABM02SIC SETS 



83 




lines through the intersection of the given lines which are harmonic 
with the latter; or, what is the same thing, if the line joining the 
points meets the lines in two points 
harmonic with the given points. 
With this understanding we may 
restate the last theorem as follows: 
The sides of a complete quadrangle 
which meet in a diagonal point are 
harmonic with the other two diago- 
nal points. In like manner, we may 
say that two points are harmonic 
with two planes, if the line joining 
the points meets the planes in a 
pair of points harmonic with the 
given points ; and a pair of lines is 
harmonic with a pair of planes, if FIG. 35 

they intersect on the intersection 

of the two planes, and if they determine with this intersection two 
planes harmonic with the given planes. 

EXERCISES 

1. Prove Theorem 4 directly from a figure without using Theorem 2, 
Chap. III. 

2. Prove Theorem 5, Cor. 2, directly from a figure. 

3. Through a given point in a plane construct a line which passes through 
the point of intersection of two given lines in the plane, without making use 

_cjfjhe latter point. 

4. A line meets the sides of a triangle ABC in the points A I9 B v C 1? and 
the harmonic conjugates A%, B 2 , C 3 of these points with respect to the two 
vertices on the same side are determined, so that we have H(^jB, C^C*), 
H(BC,A l A 2 )^nd.H(CA,B l B^ Show that A 19 B v C 2 ; ^C^A^C^A^ 
are collinear; that AA^ BB 2 , CC 2 are concurrent; and that AA 2 , BB V CC X ; 
AA V BB^ CCtf AA^BBv CC 2 are also concurrent. 

5. If each of two sides AB, EC of a triangle ABC meets a pair of opposite 
edges of a tetrahedron in two points which are harmonic conjugates with 
respect to .4, B and B, C respectively, the third side CA will meet the third 
pair of opposite edges in two points which are harmonic conjugates with 
respect to C, A. 

6. A, B, C, D are the vertices of a quadrangle the sides of which meet a 
given transversal I in the six points P 1? P 2 , P s , P 4 , P 5 , P 6 ; the harmonic conju- 
gate of each of these points with respect to the two corresponding vertices of the 



84 THE FUOTDAMElSrTAL THEOEEM [CHAP. IV 

quadrangle is constructed and these six points are denoted by P{, P%, P%, P, 
P$, PQ respectively. The three lines joining the pairs of the latter points 
which lie on opposite sides of the quadrangle meet in a point P, which is the 
harmonic conjugate of each of the points in -which these three lines meet I 
with respect to the pairs of points P' denning the lines. 

7. Denning the polar line of a point with respect to a pair of lines as the 
harmonic conjugate line of the point with regard to the pair of lines, prove 
that the three polar lines of a point as to the pairs of lines of a triangle form 
a triangle (called the cogredient triangle) perspective to the given triangle. 

8. Show that the polar line denned in Ex. 7 is the same as the polar line 
denned in Ex. 8, p. 52. 

9. Show that any line through a point and meeting two intersecting 
lines Z, V meets the polar of with respect to I, I' in a point which is the 
harmonic conjugate of with respect to the points in which the line through 
meets Z, V. 

10. The axis of perspectivity of a triangle and its cogredient triangle is the 
polar line (cf . p. 4C) of the triangle as to the given pointt 

11. If two triangles are perspective, the two polar lines of a point on their 
axis of perspectivity meet on the axis of perspectivity. 

12. If the lines joining corresponding vertices of two n-lines meet in a point, 
the points of intersection of corresponding sides meet on a line. 

13. (Generalization of Exs. 7, 10.) The n polar lines of a point P as to the n 
(n l)-lines of an n-line in a plane form an ra-line (the cogredient w-line) 
whose sides meet the corresponding sides of the given n-line in the points of 
a line p. The line p is called the polar of P as to the ra-line.* 

14. (Generalization of Ex. 11.) If two si-lines are perspective, the two 
polar lines of a point on their axis of perspectivity meet on this axis. 

15. Obtain the plane duals of the last- two problems. Generalize them to 
three- and 7i-dimensional space. These theorems are fundamental for the con- 
struction of polars of algebraic curves and surfaces of the n-th degree* 

32. Kets of rationality on a line, DEFINITION. A point P of a line 
is said to be "harmonically related to three given distinct points A, B t G 
of the line, provided P is one of a sequence of points A, B, C, IT 1} j6T 2 , _H" 3 , 
- of the line, finite in number, such that ^ is the harmonic conju- 
gate of one of the points A, B, C with respect to the other two, and 
such that every other point H t is harmonic with three of the set A, B, C, 
H v HI, - - 3 j5" _ r The class of all points harmonically related to three 
distinct points A, B, C on a line is called the one-dimensional net of 
rationality defined by A, B, (7; it is denoted by R(A0). A net of 
rationality on a line is also called a linear net 

* This is a definition "by induction of the polar line of a point with respect to an 
n-line. 



32] NETS OF BATIOXALITY 85 

THEOREM 7. If A, 3, C, D and A 1 , B f , C f , I) 1 are respectively points 
of two lines stick that ABCD-^A r B f C r Z> f , and if D is harmonically 
related to A, 3, C, then D r is harmonically related to A r ,B r , C f . (A, E) 

This follows directly from the fact that the projectivity of the theo- 
rem makes the set of points Hj which defines D as harmonically related 
to A,B, C pro jective with a set of points H( such that every harmonic set 
of points of the sequence A 9 B, C, H v H^ - -, D is homologous with a 
harmonic set of the sequence^ 7 , B f , C f , JBJ, H 9 ^ '-,D f (Theorem 3, Cor.). 

COROLLARY. If a class of points on a line is projectile with a net 
of rationality on a line, it is itself a net of rationality. 

THEOREM 8. If K, L, JJ are three distinct points of R (ABC), A, B, C 
are points of R (KLM). (A, E) 

Proof. From the projectivity ABCKj^ BAKC follows, by Theorem?, 
that C is a point of R (ABIC). Hence all points harmonically related 
to A, B, C are, by definition, harmonically related to A, B, K. Since K 
is, by hypothesis, in the net R (ABC), the definition also requires that 
all points of R(ABK] shall be points of R(ABC). Hence the nets 
R(ABC) and R(ABK) are identical; and so R(ABC) = R(ABK) 



COROLLARY. A net of rationality -on a line is determined ly any 
distinct three of its points. 

THEOREM 9. If all but one of the six (or five, or four) points of a 
quadrangular set are points of the same net of rationality R, this 
one point is also a point of R. (A, E) 

Proof. Let the sides of the quadrangle PQRS (fig. 37) meet the 
line I as indicated in the points A, A^ B,B^ C } C v so that B = B^\ 
and suppose that the first five of these are points of a net of rationality 

R =: 



We must prove that G : is a point of R. Let the pair of lines ES and 
PQ meet in B f . We then have 

S M 

BCB^A = BQB f P = BAjBiCr 

Since A is in. R(BCBJ, it follows from this projectivity, in view of 
Theorem 7, that C l is in R (BA&) == R. 

DEFINITION. A point P of a line is said to be guadrangularly 
related to three gwen distinct points A, B, C of the line, provided 



86 



THE FUNDAMENTAL THEOEEM 



[CHAP. IV 



P is one of a sequence of points A, B, C, H v H^ H# of the line, 
finite in number, such that H^ is the harmonic conjugate of one of 
the points A, B, with respect to the other two, and such that every 
other point H t is one of a quadrangular set of which the other five 
belong to the set A, B, C, & v H v - - -, ^_ r 




COROLLAET. The class of all points Quadrangularly related to three 
distinct collinear points A, B, C is R(ABC). (A, E) 

From the last corollary it is plain that R (ABO) consists of all points that 
can be constructed from A , .5, C by means of points and lines alone ; that is 
to say, all points -whose existence can be inferred from Assumptions A, E, H . 
The existence or nonexistence of further points on the line ABC is unde- 
termined as yet. The analogous class of points in a plane is the system of all 
points constructive, by means of points and lines, out of four points A,B,C, D, 
no three of which are collinear. This class of points is studied by an indirect 
method in the next section. 

33. Nets of rationality in the plane. DEFINITION. A point is said 
to be rationally related to two noncollinear nets of rationality R v R 2 
having a point in common, provided it is the intersection of two lines 
each of which joins a point of R l to a distinct point of R 2 . A line is 
said to be rationally related to R and R 2 , provided it joins two points 
that are rationally related to them. The set of all points and lines 
rationally related to R v R 2 is called the net of rationality in a plane 
(or of two dimensions) determined by R v R 2 ; it is also called the 
planar net defined by R x> R 2 . 

From this definition it follows directly that all the points of R x 
and R 2 are points of the planar net defined by R v R 2 . 



133] NETS OF EATIONALITY 87 

THEOKEM 10. Any line of the planar net R 2 defined "by R v R 2 meets 



Proof. We prove first that if a line of the planar net R 2 meets R v 
it meets R 2 . Suppose a line I meets R x in A^ it then contains a second 
point P of R 2 . By definition, through P pass two lines, each of which 
joins a point of R 1 to a distinct point of R 2 . If I is one of these lines, 
the proposition is proved ; if these lines are distinct from Z, let them 
meet R x and R 2 respectively in the points B v B 2 and JJ, J% (fig. 38). 
If is the common point of R v R 2 , we then have 



,, 

where -4 2 is the point in which I meets the line of R 2 . Hence A* is a 
point of R 2 (Theorem 7). 

Now let I be any line of the net R 3 , and let P, Q be two points 
of the net and on I (def.). If one of these points is a point of R x or 
R 2 , the theorem is proved by the case just considered. If not, two 
lines, each joining a point of R l to a distinct point of R s , pass through 
P; let them meet R l in A v B It and R 2 in A z , B^ respectively (fig. 38). 
Let the lines QA^ and QB l meet R 2 in A[ and B^ respectively (first case). 




Then if I meets the lines of R l and R 2 in Jj* and P% respectively, the 
quadrangle PQA I B 1 gives rise to the quadrangular set Q(I^A^B Z9 
OB^) of which five points are points of R 2 ; hence P z is a point of R 2 
(Theorem 9). J? is then a point of R x by the first case of this proof. 

THEOREM 11. The intersection of any two lines of a planar net is 
a point of the planar net (A, E) 



88 



THE FUNDAMENTAL THEOEEM 



[CHAP. IV 



Proof. This follows directly from the definition and the last theo- 
rem, except when one of the lines passes through 0, the point common 
to the two linear nets R 1? R 2 defining the planar net. In the latter 
case let the two lines of the planar net be l v 1 2 and suppose 1 2 passes 
through 0, while ^ meets R 1? R 2 in A v A z respectively (fig. 39). If the 
point of intersection P of IJ 2 were not a point of the planar net, 1 2 

would, by definition, 
contain a point Q of 
the planar net, dis- 
tinct from and P. 
The lines QA l and 
QA a would meet R 2 
and R 1 in two points 
2? 2 and 2^ respec- 
tively. The point <7 2 
in which the line 
PB^ met the line of 
R 2 would then be the 
harmonic conjugate 




FIG. 39 



of JB Q with respect to and A 2 (through the quadrangle 



<7 2 would therefore be a point of 



R 2 , and hence 



P would be a 



point of the planar net, being the intersection of the lines A^A 2 
and B^dy 

THEOREM 12. TJie points of a planar net R 2 on a line of the planar 
net form a linear net. (A, E) 

Proof. Let the planar net be defined by the linear nets R x , R 2 and 
let I be any line of the planar net. Let P be any point of the planar 
net not on I or R l or R 2 . The lines joining P to the points of R 2 on I 
meet R x and R 2 by Theorems 10 and 11. Hence P is the center of 
a perspectivity which makes the points of R 2 on I perspective with 
points of R 1 or R 2 . Hence the points of I belonging to the planar net 
form a linear net. (Theorem 7, Cor.) 

COROLLARY. The planar net R* defined ly two linear nets R v R 2 is 
identical with the planar net R| defined "by two linear nets R 8 , R 4 , pro- 
vided R s , R 4 are linear nets in R^. (A, E) 

For every point of R 2 is a point of R| by the above theorem, and 
every point of R^ is a point of Rf by Theorem 10. 



33,34] NETS OF KATIONALITY 89 

EXERCISE 

If A, B 9 C, D are the vertices of a complete quadrangle, there is one and 
only one planar net of rationality containing them ; and a point P belongs to 
this net if and only if P is one of a sequence of points ABCDD^D^ , finite 
in number, such that D l is the intersection of two sides of the original quad- 
rangle and such that every other point D t is the intersection of two lines join- 
ing pairs of points of the set ABCDD^ A-i- 

34. Nets of rationality in space. DEFINITION. A point is said to 
be rationally related to two planar nets R 2 , R| in different planes but 
having a linear net in common, provided it is the intersection of two 
lines each, of which joins a point of R 2 to a distinct point of R 2 2 . 
A line is said to be rationally related to R 2 , R|, if it joins two, a plane 
if it joins three, points which are rationally related to them. The set 
of all points, lines, and planes rationally related to R 2 , R| is called the 
net of rationality in space (or of three dimensions) determined by 
RI > R% 5 ft is a l so called the spatial net defined by R 2 , R 2 2 . 

Theorems analogous to those derived for planar nets may now be 
derived for nets of rationality in space. We note first that every point 
of R 2 and of R * is a point of the spatial net R 3 defined by R x 2 , R * (the 
definition applies equally well to the points of the linear net common 
to R 2 , R|) ; and that no other points of the planes of these planar nets 
are points of R 3 . The proofs of the fundamental theorems of align- 
ment, etc., for spatial nets can, for the most part, be readily reduced 
to theorems concerning planar nets. We note first : 

LEMMA. Any line joining a point A^ of R? to a distinct point P of 
R 3 meets R 2 . (A, E) 

Proof. By hypothesis, through P pass two lines, each of which 
joins a point of R 2 to a distinct point of R 2 2 . We may assume these 
lines distinct from the line PA V since otherwise the lemma is proved. 
Let the two lines through P meet R a 2 , R| in B v B 2 and C v C z respec- 
tively (fig. 40). If A v B v C l are not collinear, the planes PA^ and 
PA^ meet R 2 in the lines A^B^ and A^ respectively, which meet 
the linear net common to R 2 , R 2 in two points S, T respectively 
(Theorems 11, 12). The same planes meet the plane of R 2 2 in the lines 
SB 2 and TC^ respectively, which are lines of R 2 , since S 9 T are points 
of R 2 2 . These lines meet in a point A 2 of R 2 2 (Theorem 11), which 
is evidently the point in which the line PA meets the plane of R 2 2 . 
If A v B v C^ are collinear, let J 2 be the intersection of PA^ with the 



90 



THE FUNDAMENTAL THEOEEM 



[CHAP. IV 



plane of R 2 2 , and S the intersection of A l B l with the linear net 
common to R? and R 2 2 . Since A 1 is in R (SBfiJ, the perspectivity 

SC 1 B l A = SC^Ao implies that A 2 is in R(SB 2 C 2 ) and hence in R 3 2 . 

m P 




Fio. 40 



THEOREM 13. Any line of the spatial net R 3 defined by R*, R| meets 
R*andR*. (A, E) 




FIG. 41 



Proof. By definition the given line I contains two points A and B 
of the net R 8 (fig. 41). If A or B is on Rf or R|, the theorem reduces 
to the lemma. If not, let ^ be a point of R.?, and A z and B 2 the points 
in which, by the lemma, %A and P^B meet R 2 2 ; also let JJ' be any 



84] STETS OF RATIONALITY 91 

point of R x 2 not in the plane P^AB, and let P^A and P^B meet R 3 2 in A 
and .#2". The lines J 2 2 and -.4^| meet in a point of R 2 2 (Theorem 11), 
and this point is the point of intersection of I with the plane of R|. 
The argument is now reduced to the case considered in the lemma. 

THEOREM 14 The points of a spatial net lying on a line of the 
spatial net form a linear net. (A, E) 

Proof. Let I be the given line, R 2 and R 2 2 the planar nets defining 
the spatial net R 3 , and L : and L 2 the points in which (Theorem 13) 
I meets Rf and R 2 2 (L t and L 2 may coincide). Let A^ be any point of 
R 2 not on I or on R 2 , and S the point in which A^L^ meets the linear 
net common to R 2 and R 2 2 (fig. 42). If L^ and L 2 are distinct, the lines 





PIG. 42 FIG. 43 

SL t and SL 2 meet R 2 and R 2 2 in linear nets (Theorem 12); and, by 
Theorem 13, a line joining any point P of R 3 on I to A l meets each 
of these linear nets. Hence all points of R 3 on I are in the planar 
net determined by these two linear nets. Moreover, by the definition 
of R 8 , all the points of the projection from A of the linear net on SL 2 
upon I are points of R 3 . Hence the points of R 3 on I are a linear net. 
If x = ,= S, then, by definition, there is on I a point A of R*, and 
the line AA^ meets R| in a point A z (fig. 43). The lines SA^ and SA 2 
meet R? and R| in linear nets R x and R 2 by Theorem 12. If JB X is 
any point of R x other than Jt p the line .A^ meets R 2 2 in a point J? 2 by 
Theorem 13. By Theorem 12 all points of I in the planar net deter- 
mined by Rj and R 2 form a linear net, and they obviously belong to R 3 . 
Moreover, any point of R 3 on I, when joined to A v meets R 2 2 by Theo- 
rem 13, and hence belongs to the planar net determined by R x and R 2 . 
Hence, in this case also, the points of R 8 on I constitute a linear net. 



92 THE FUNDAMENTAL THEOEEM [CHAP, iv 

THEOREM 15. The points and lines of a spatial net R 3 which lie on 
a plane a of the net form a planar net. (A, E) 

Proof. By definition a contains three noncollinear points A, B, C of 
R 8 , and the three lines AB, BC, CA meet the planar nets R* and R 2 2 , 
which determine R 3 , in points of two linear nets R x and R 2 , consisting 
entirely of points of R 3 . These linear nets, if distinct, determine a 
planar net R 2 in a, which, by Theorem 10, consists entirely of points 
and lines of R 3 . Moreover, any line joining a point of R 3 in a to A 
or B or C must, by Theorem 13, meet R x and R 2 and hence be in R\ 
Hence all points and lines of R 3 on a are points and lines of R 2 . This 
completes the proof except in case R t = R a , which case is left as an 
exercise. 

COROLLARY 1. A net of rationality in space is a space satisfying 
Assumptions A and E, if "line" le interpreted as "linear net" and 
"plane" as "planar net." (A, E) 

For all assumptions A and E, except A 3, are evidently satisfied ; 
and A 3 is satisfied because there is a planar net of points through 
any three points of a spatial net R 3 , and any two linear nets of this 
planar net have a point in common. 

This corollary establishes at once all the theorems of alignment in 
a net of rationality in space, which are proved in Chap. I, as also the 
principle of duality. "We conclude then, for example, that two planes 
of a spatial net meet in a line of the net, and that three planes of a 
spatial net meet in a point of the net (if they do not meet in a line), 
etc. Moreover, we have at once the following corollary : 

COROLLARY 2. A spatial net is determined ly any two of its planar 

nets. (A, E) 

EXERCISES 

1. If A,B, C,Z>, E are the vertices of a complete space five-point, there is 
one and only one net of rationality containing them all. A point P belongs to this 
net if and only if P is one of a sequence of points ABCDEI^I^ , finite 'in 
number, such that I I is the point of intersection of three faces of the original 
five-point and every other point J t - is the intersection of three distinct planes 
through triples of points of the set ABCDEI^ - - - 1^^. 

2. Show that a planar net is determined if three noncollinear points and a 
line not passing through any of these points is given. 

3. Under what condition is a planar net determined by a linear net and two 
points not in this net ? Show that two distinct planar nets in the same plane 
can have at most a linear net and one other point in common. 



34, 35] . THE FUNDAMENTAL THEOREM: 93 

4. Show that a set of points in a plane -which is protective with L, planar 
net is a planar net. 

5. A line joining a point P of a planar net to any point not in the net, but 
on a line of the net not containing P : has no other point than P in common 
with the net. 

6. Two points and two lines in the same plane do not in general belong to 
the same planar net. 

7. Discuss the determination of spatial nets by points and planes, similarly 
to Exs. 2, 3, and 6. 

8. Any class of points projective with a spatial net is itself a spatial net. 

9. If a perspective collineation (homology or elation) in a plane with 
center A and axis I leaves a net of rationality in the plane invariant, the 
net contains A and /. 

10. Prove the corresponding proposition for a net of rationality in space 
invariant tinder a perspective transformation. 

11. Show that two linear nets on skew lines always belong to some spatial 
net ; in fact, that the number of spatial nets containing two given linear 
nets on skew lines is the same as the number of linear nets through two given 
points. 

12. Three mutually skew lines and three distinct points on one of them 
determine one and only one spatial net in which they lie. 

13. Give further examples of the determination of spatial nets by lines, 

35. The fundamental theorem of projectivity. It has been shown 
(Chap. Ill) that any three distinct elements of a one-dimensional 
form may be made to correspond to any three distinct points of a 
line by a projective transformation. Likewise any four elements of 
a two-dimensional form, no three of which belong to the same one- 
dimensional form, may be made to correspond to the vertices of a 
complete planar quadrangle by a projective transformation ; and any 
five elements of a three-dimensional form, no four of which belong 
to the same two-dimensional form, may be made "to correspond to 
the five vertices of a complete spatial five-point by a projeetive 
transformation. 

These transformations are of the utmost importance. Indeed, it is 
the principal object of projective geometry to discover those prop- 
erties of figures which remain invariant when the figures are sub- 
jected to projective transformations. The question now naturally 
arises, Is it possible to transform any four elements of a one- 
dimensional form into any four elements of another one-dimensional 
form ? This question must be answered in the negative, since a har- 
monic set must always correspond to a harmonic set. The question 



94 THE FUNDAMENTAL THEOEEM [CHAP.IV 

tlien arises whether or not a projective correspondence between one- 
dimensional forms is completely determined when three pairs of 
homologous elements are given. A partial answer to this funda- 
mental question is given in the next theorem. 

LEMMA 1. If a projectivity leaves three distinct points of a line fixed, 
it leaves fixed every point of the linear net defined ~by these points. 

This follows at once from the fact that if three points are left 
invariant by a projectivity, the harmonic conjugate of any one of 
these points with respect to the other two must also be left inva- 
riant by the projectivity (Theorems 2 and 3, Cor.). The projectivity 
in question must therefore leave invariant every point harmonically 
related to the three given points. 

THEOREM 16. THE FUNDAMENTAL THEOREM OF PROJECTIVITY FOR A 
NET OF RATIONALITY ON A LINE. If A, B, C, D are distinct points of 
a linear net of rationality, and A r , J3 f , C f are any three distinct points 
of another or the same linear net, then for any proj'ectivities giving 
ABCD-^AWC'D* and ABCD^A r r O r D[ 9 we have D J =D[. (A, E) 

Proof. If TT, TT X are respectively the two pro jectivities of the theorem, 
the projectivity Tr^" 1 leaves A'JB f C f fixed and transforms D f into D[. 
Since D f is harmonically related to A f , B r , C r (Theorem 7), the theorem 
follows from the lemma. 

This theorem gives the answer to the question proposed in its 
relation to the transformation of the points of a linear net. The 
corresponding proposition for all the points of a line, ie. the prop- 
osition obtained from the last theorem by replacing " linear net " by 
"line," cannot be proved without the use of one or more additional 
assumptions (cf. 50, Chap. VI). We have seen that it is equiva- 
lent to the proposition: If a projectivity leaves three points of a 
line invariant, it leaves every point of the line invariant. Later, by 
means of a discussion of order and continuity (terms as yet unde- 
fined), we shall prove this proposition. This discussion of order 
and continuity is, however, somewhat tedious and more difficult 
than the rest of our subject ; and, besides, the theorem in question 
is true in spaces,* where order and continuity do not exist. It has 

* Different, of course, from ordinary space; " rational spaces*' (cf. p. 98 and 
the next footnote) are examples in which continuity does not exist; u finite spaces," 
of which examples are given in the introduction ( 2), are spaces in which neither 
order nor continuity exists. 



35] THE FUNDAMENTAL THEOREM 95 

therefore seemed desirable to give some of the results of this 
theorem before giving its proof in terms of order and continuity. 
To this end we introduce here the following provisional assumption 
of projectivity, which will later be proved a consequence of the order 
and continuity assumptions which will replace it. This provisional 
assumption may take any one of several forms. "We choose the fol- 
lowing as leading most directly to the desired theorem : 

AN ASSUMPTION OF PROJECTIVITY : 

P. If a projectivity leaves each of three distinct points of a line- 
invariant, it leaves every point of the line invariant* 

We should note first that the plane and space duals of this assump- 
tion are immediate consequences of the assumption. The principle of 
duality, therefore, is still valid after our set of assumptions has been 
enlarged by the addition of Assumption P. 

"We now have : 

THEOEEM 17. THE FUNDAMENTAL THEOREM OF PROJECTTTE GEOM- 
ETRY, f If 1, @, 3,4- are any four elements of a one-dimensional primitive 
form, and l f , % ? , 3 l are any three elements of another or the same one- 
dimensional primitive form, then for any projectivities giving 134~^ 
1W&4' and 1^34^1 r ^3 f 4[ } we have 4'=4[. (A, E, P) 

Proof. The proof is the same under the principle of duality as that 
of Theorem 16, Assumption P replacing the previous lemma. 

This theorem may also be stated as follows : 

A projectivity 'between one-dimensional primitive, forms is uniquely 
determined when three pairs of homologous elements are given. (A, E, P) 

COROLLARY. If tico pencils of points on different lines are projectile 
and have a self-corresponding point, they are perspective. (A, E, P) 

* We have seen in the lemma of the preceding theorem that the projectivity 
described in this assumption leaves invariant every point of the net of rationality 
defined by the three given points. The assumption simply states that if all the points 
of a linear net remain invariant under a protective transformation, then all the points 
of the line containing this net must also remain invariant. It will be shown later 
that in the ordinary geometry the points of a linear net of rationality on a line corre- 
spond to the points of the line whose coordinates, when represented analytically, are 
rational numbers. This consideration should make the last assumption almost, if 
not quite, as intuitionally acceptable as the previous Assumptions A and E. 

t On this theorem and related questions there is an extensive literature to which 
references can be found in the Encyklopadie articles on Protective Geometry and 
Foundations of Geometry. It is associated with the names of von Staudt, Klein, 
Zeuthen, Liiroth, Darboux, F. Schur, Pieri, Wiener, Hilbert. Cf . also 50, Chap. VI. 



96 THE FUNDAMENTAL THEOREM [CHAP.IV 

Proof. For if is the self-corresponding point, and AA f and BB' 
are any two pairs of homologous points distinct from 0, the perspec- 
tivity whose center is the intersection of the lines A A', BB 1 is a 
projectivity between the two lines which has the three pairs of 
homologous points 00, AA f , BB f , which must be the projectivity of 
the corollary by virtue of the last theorem. 

The corresponding theorems for two- and' three-dimensional forms 
are now readily derived. We note first, as a lemma, the propositions 
in a plane and in space corresponding to Assumption P. 

LEMMA 2. A projectile transformation which leaves invariant each 



* j. * four > j. ,c a plane three , . , , 7 , ^ 
of a set of - points of * no of WIMMI oelonq to me same 
J J five * J space four J y 

line ^ . . . . _, - the plane. , . 

, leaves invariant every point of (A. K P) 

plane J * J space. ^ ' 

Proof. If A, B, C, D are four points of a plane no three of which 
are collinear, a projective transformation leaving each of them in va- 
riant 'must also leave the intersection of the lines AB, CD invariant. 
By Assumption P it then leaves every point of each of the lines AB, 
CD invariant. Any line of the plane which meets the lines AB and 
CD in two distinct points is therefore invariant, as well as the inter- 
section of any two such lines. But any point of the plane may be 
determined as the intersection of two such lines. The proof for the 
case of a projective transformation leaving invariant five points no 
four of which are in the same plane is entirely similar. The existence 
of perspective collineations shows that the condition that no three 
(four) of the points shall be on the same line (plane) is essential. 

THEOREM 18. A projective collineation* "between two planes (or 
within a single plane) is uniquely determined when four pairs of 
homologous points are given, provided no three of either set of four 
points are collinear. (A, E, P) 

Proof. Suppose there were two collineations TT, TT X having the given 
pairs of homologous points. The collineation irjr~ l is then, by the 
lemma, the identical collineation in one of the planes. This gives at 
once TTj = TT, contrary to the hypothesis. 

* We confine the statement to the case of the collineation for the sake of sim- 
plicity of enunciation. Projective transformations which are not collineations will 
be discussed in detail later, at which time attention will be called explicitly to the 
fundamental theorem. 



35] THE FUSTDAMEXTAL THEOBEM 97 

By precisely similar reasoning we have : 

THEOREM 19. A protective collineation in space is uniquely deter- 
mined when five pairs of liomoloyous points are given, provided no 
four of either set of five points are iii the same plane. (A, E, P) 

The fundamental theorem deserves its name not only because so 
large a part of projective geometry is logically connected with it, but 
also because it is used explicitly in so many arguments. It is indeed 
possible to announce a genial course of procedure that appears in 
the solution of most " linear " problems, i.e. problems which depend on 
constructions involving points, lines, and planes only. If it is desired 
to prove that certain three lines l v 1 2 , / 3 pass through a point, find two 
other lines m v m^ such that the four points mj v mj^ mJ ZJ m^n^ may 
be shown to be projective with the four points mj v mj^, mj,^ mjn^ 
respectively. Then, since in this projectivity the point m^n^ is self- 
corresponding, the three lines l v l, 1 B joining corresponding points 
are concurrent (Theorem 17, Cor.). The dual of this method appears 
when three points are to be shown collinear. This method may be 
called the principle of projectivity, and takes its place beside the 
principle of duality as one of the most powerful instruments of pro- 
jective geometry. The theorems of the next section may be regarded 
as illustrations of this principle. They are all propositions from which 
the principle of projectivity could be derived, ie. they are propositions 
which might be chosen to replace Assumption P. 

We have already said that ordinary real (or complex) space is a 
space in which Assumption P is valid. Any such space we call a 
properly projective space. It will appear in Chap. VI that there 
exist spaces in which this assumption is not valid. Such a space, 
ie. a space satisfying Assumptions A and E but not P, we will call 
an improperly projective space. 

From Theorem 15, Cor. 1 and Lemma 1, we then have 

THEOREM 20. A net of rationality in space is a properly projective 
space. (A, E) 

It should here be noted that if we added to our list of Assump- 
tions A and E another assumption of closure, to the effect that all 
points of space belong to the same net of rationality, we should 
obtain a space in which all our previous theorems are valid, in- 
cluding the fundamental theorem (without using Assumption P). 



98 



THE FUNDAMENTAL THEOREM 



[CHAP. IV 



Such a space may be called a rational space. In general, it is clear 
that any complete five-point in any properly or improperly protective 
space determines a subspace which is rational and therefore properly 
projective. 

36. The configuration of Pappus. Mutually inscribed and circum- 
scribed triangles. 

THEOREM 21. If A, B, C are any three distinct points of a line I, 
and A\ B\ C f any three distinct points of another line I' meeting I, 
the three points of intersection of the pairs of lines AB r and A!B> BC 1 
and B'C, CA 1 and C f A 
are collinear. (A, E, P) 




Proof. Let the three points of intersection referred to in the theorem 
be denoted by C r ", A", B" respectively (fig. 44). Let the line "C" 
meet the line B'C in a point D (to be proved identical with A"); 
also let B"C" meet V in A v the line A'B meet AC r in B v the line AB' 
meet A r C in B[. We then have the following perspectivities : 

4 



By the principle of projectivity then, since in the projectivity thus 
established C n is self-corresponding, we conclude that the three lines 
A^A'i B n B v DB meet in the point '. Hence D is identical with A n , 
and A", B ff , C" are collinear. 

It should be noted that the figure of the last theorem is a con- 
figuration of the symbol 




COXFIGUEATICXtf OF PAPPUS 99 

It is known as the concur crtiajL. of Pappus* It should also be noted 
that this configuration may be considered as a simple plane hexagon 
(six-point) inscribed in two intersecting lines. If the sides of such a 
hexagon be denoted in order by 1, 2, 3, 4, 5, 6, and if we call the sides 
1 and 4 opposite, likewise the sides 2 and 5, and the sides 3 and 6 (cf. 
Chap. II, 14), the last theorem may be stated in the following form : 
COROLLARY. If a simple hexagon le inscribed in two intersecting lines, 
the three pairs of opposite sides will intersect in collinear points.^ 

Finally, we may note that the nine points of the configuration of 
Pappus may be arranged in sets of three, the sets forming three 
triangles, 1, 2, 3, such 
that 2 is inscribed in 
1, 3 in 2, and 1 in 3. 
This observation leads 
to another theorem con- 
nected with the Pappus 
configuration. 

THEOREM 22. If 
A^B^C^ be a triangle 

inscribed in a triangle p iG 45 

A^B! V there exists a 

certain set of triangles each of which is inscribed in the former and 
circumscribed about the latter. (A, E, P) 

Proof. Let \a\ be the pencil of lines with center A^ [&] the pencil 
with center B^ and [c] the pencil with center C^ (fig. 45). Consider the 

perspectivities [a] 2 [&] == [c]. In, the projectivity thus estab- 
lished between [a] and [c] the line A l C l is self-corresponding; the 
pencils of lines [a], [c] are therefore perspective (Theorem 17, Cor. 
(dual)). Moreover, the axis of this perspectivity is C 2 A^ for the lines 
Aft* and C^ are clearly homologous, as also the lines A^A Z and O^A^. 
Any three homologous lines of the perspective pencils [a], [&], [c] then 
form a triangle which is circumscribed about A 1 B 1 C 1 and inscribed 
in A^B^C^ 

* Pappus, of Alexandria, lived about 340 A.D. A special case of this theorem may 
be proved without the use of the fundamental theorem (cf. Ex. 3, p. 52). 

t In this form it is a special case of Pascal's theorem on conic sections 
(cf. Theorem 3, Chap. V). 




100 



THE FUNDAMENTAL THEOREM 



[CHAP. IV 



EXERCISES 

1. Given a triangle ABC and two distinct points A', Z>'; determine a point C" 
sucli that the lines A A', BB', CC'are concurrent, and also the lines A I?, BC', CA' 
are concurrent, i.e. such that the two triangles are perspective from two dif- 
ferent points. The two triangles are there said to he doubly perspective. 

2. If two triangles ABC and A'B'C' are doubly perspective in such a way 
that the vertices A, B, C are homologous with A', B', C' respectively in one 
perspectivity and with J3', C', A' respectively in the other, they will also be per- 
spective from a third point in such a way that A, B, C are homologous respec- 
tively with C', A', B'^ i.e. they will be triply perspective. 

3. Show that if A", B", C" are the centers of perspectivity for the triangles 
in Ex. 2, the three triangles ABC, A'B'C', A"B"C" are so related that any two 
are triply perspective, the centers of perspectivity being in each case the vertices 
of the remaining triangle. The nine vertices of the three triangles form the 
points of one configuration of Pappus, and the nine sides form the lines of 
another configuration of Pappus. 

37. Construction of projectivities on one-dimensional forms. 
THEOREM 23. A necessary and sufficient condition for the projectivity 
on a line MNAB -^ MNA'B f ( 31 * JV) is Q(3IAB, NB'A'). (A, E, P) 




Proof. Let n be any line on* N not passing through A (fig. 46). Let O l 
be any point not on n or on MA, and let A l and B t be the intersections 
respectively of O t A and O^B with n. Let 2 be the intersection 



and B'B V 



Then 



0, 



0, 



By Theorem 17 the projectivity so determined on the line AM is the 
same as 



The only possible double points of the projectivity are N and the 
intersection of AN with O a 2 . Hence 0^ passes through M 9 and 
Q(MAB, NB f A f ) is determined by the quadrangle O^A^ 



37] ONE-DIMENSIONAL PEOJECTIYITIES 101 



Conversely, if Q(1LLB, XJB'A') we have a quadrangle O 
and hence 

- 



and by this construction 31 is self-corresponding, so that 



If iu the above construction we have J/"= JV, we obtain a projec- 
tivity with the single double point J/= JV. 

DEFINITION. A projectivity on a one-dimensional primitive form 
with a single double element is called parabolic. If the double ele- 
ment is J/, and AA f 9 BB* are any two homologous pairs, the pro- 
jectivity is completely determined and is conveniently represented 
by MMAB ^ MMA'B'. 

COROLLARY. A necessary and sufficient condition for a parabolic 
projectivity MHAB-^MMA'B* is Q(3IA, MB'A f ). (A, E, P) 

THEOREM 24 If we hare 

Q(ABC, A'3 f C f ), 
we have also Q(A f B'C r , ABC). 

Proof. By the theorem above, 

Q(A3C, A'B'C 1 ) 

implies AA r BC^ AA'C'B', 

which is the inverse of A r AB r C f -^ A A OB 9 

which, by the theorem above, implies 

Q(A'B f C' 9 ABC). 

The notation Q (ABC, A f B f C l ) implies that A, B, C are the traces of a 
point triple of sides of the quadrangle determining the quadrangular set. 
The theorem just proved states the existence of another quadrangle 
for which A\ B r , C 1 are a point triple, and consequently A, B, C are a 
triangle triple. This theorem therefore establishes the existence of 
oppositely placed quadrangles, as stated in 19, p. 50. This result 
can also be propounded as follows: 

THEOREM 25. If two quadrangles I^P 2 P Z P^ and Q^^Q^Q^ are so related 
J? to Q v J% to Q y etc. that five of the sides l$I$(i,j =1,2,8,4; 
i =/) meet the five sides of the second which are opposite to Q t Qj in points 
of a line I, the remaining sides of the two quadrangles meet on L (A,E,P) 



102 THE FUNDAMENTAL THEOEE3M [CHAP. IV 

Proof. The sides of the first quadrangle meet I in a quadrangular 
set Q (P 12 P 13 P^ P*Pu&) ; hence Q (P 2 PA P*PM* But, by hypoth- 
esis, five of the sides of the second quadrangle pass through these 
points as follows : Q^ through J^, Q^Q 9 through P^ 9 Q& through P^> 
Q z Qt through 2J a , QQ 2 through P is , Q 2 Q Z through P u . As five of these 
conditions are satisfied, by Theorem 3, Chap. II, they must all be 
satisfied. 

EXERCISES 

1. Given one double point of a projectivity on a line and two pairs of 
homologous points, construct the other double point. 

2. If a, b, c are three nonconcurrent lines and A', B', C' are three collinear 
points, give a construction for a triangle whose vertices A. , B, C are respectively 
on the given lines and whose sides BC, CJL, AB pass respectively through the 
given points. What happens when the three lines a , I, c are concurrent ? Dualize. 

38. Involutions, DEFINITION. If a projectivity in a one-dimensional 
form is of period two, it is called an involution. Any pair of homol- 
ogous points of an involution is called a conjugate pair of the involution 
or a pair of conjugates. 

It is clear that if an involution transforms a point A into a point A r , 
then it also transforms A r into A] this is expressed by the phrase that 
the points A, A 1 correspond to each other doully. The effect of an invo- 
lution is then simply a pairing of the elements of a one-dimensional 
form such that each element of a pair corresponds to the other ele- 
ment of the pair. This justifies the expression "a conjugate pair" 
applied to an involution. 

THEOREM 26. If for a single, point A of a line which is not a double 
point of a projectivity TT on the. line we have the relations 7r(A)t=A ! 
and 7r(A') = A, the projectivity is an involution. (A, E, P) 

Proof. For suppose P is any other point on the line (not a double 
point of ?r), and suppose 7r(P) = P f . There then exists a projectivity 
giving AA'PP 1 ^ A'AP'P 

(Theorem 2, Chap. III). By Theorem 17 this projectivity is TT, since 
it has the three pairs of homologous points A, A*\ A r , A\ P, P r . But 
in this projectivity P r is transformed into P. Thus every pair of 
homologous points corresponds doubly. 

COROLLARY. An involution is completely determined when two pairs 
of conjugate points are given. (A, E, P) 



INVOLUTIONS 103 

THEOREM 27. A necessary and sufficient condition that three pairs 
of points A, A r ; B, B' ; C, C 1 be conjugate pairs of an involution is 
Q(ABC,A f B'C f ). (A, E, P) 
Proof. By hypothesis we have 

AA'BC-^A'AB'C'. 

By Theorem 2, Chap. Ill, we also have 
A'AS'C'-fiA 
which, with the first projectivity, gives 



A necessary and sufficient condition that the latter projectivity hold 
is Q(ABC, AWC r ) (Theorem 23). 

COROLLARY 1. If an involution has doulle points, they are harmonic 
conjugates with respect to every pair of the involution. (A, E, P) 

For the hypothesis A~A f ,B B J gives at once H(AB 9 CC f ) as the 
condition of the theorem. 

COROLLARY 2. An involution is completely determined when two 
double points are given, or when one double point and one pair of 
conjugates are given. (A, E, P) 

COROLLARY 3. If J/, N are distinct doulle points of a projectivity 
on a line, and A,A r ; B,B' are any two pairs of homologous elements, 
the pairs M, N; A, B J ; A 1 , B are conjugate pairs of an involution.* 
(A, E, P) 

COROLLARY 4 If an involution has one double element, it has another 
distinct from the first. (A, E, P) 

COROLLARY 5. The projectivity ABGD-j^ABDC "between four dis- 
tinct points of a line implies the relation H (AB, CD). (A, E, P) 

Eor the projectivity is an involution (Theorem 26) of which A, B 
are double points. The result then follows from Cor. 1. 
39. Axis and center of tiomology. 

THEOREM 28. If \A] and [B] THEOREM 28'. If [I] and \m\ 
are any two projectile pencils are any two projectile pencils of 
of points in the same plane on lines in the same plane on distinct 

* This relation is sometimes expressed by saying, s l The pairs of points are in 
involution." From what precedes it is clear that any two pairs of elements of a 
one-dimensional form are in involution, but in general three pairs are not. 



104 



THE FUNDAMENTAL THEOEEM 



[CHAP. IV 



distinct lines l v 1 2> there exists a 
line I siich that if A v B I and A^ B z 
are any two pairs of homologous 
points of the two pencils, the lines 
A^ and A^ intersect on I 
(A,E,P) 

DEFINITION. The line I is called 
the axis of homology of the two 
pencils of points. 



points S v S Q9 there exists a point S 
such that if a v \ and a 2> & 2 are 
any two pairs of homologous lines 
of the two pencils, the points a^ 
and a 2 6 1 are collinear with S. 
(A, E, P) 

DEFINITION. The point S is 
called the center of homology of 
the pencils of lines. 



Proof. The two theorems being plane duals of each other, we may 
confine ourselves to the proof of the theorem on the left. From the 
projectivity [S] ^ [A] follows A^B^B^A] (fig. 47). But in this pro- 
jectivity the line A^ is self-corresponding, so that (Theorem 17, Cor.) 



A 




A 



& 

47 



the two pencils are perspective. Hence pairs of corresponding lines 
meet on a line I ; e.g. the lines A^ s and B^ z meet on I as well as 
A^B Z and B^Ay To prove our theorem it remains only to show that 
B 2 A S and A 2 B^ also meet on L But the latter follows at once from 
Theorem 21, since the figure before us is the configuration of Pappus. 
COROLLARY. If [A], [B] are not COROLLARY. If [I], [m] are not 
perspective) the axis of homology is perspective, the center of homology 



the line joining the points homol- 
ogous with the point l^ regarded 
first as a point of l^ and then as 
a point of l y 



is the point of intersection of the 
lines homologous with the line S^ 
regarded first as a line of [I] and 
then as a line of [m]. 



For in the perspectivity J^B] = B l [A'] the line ^ corresponds to 






) > and hence the point l^ corresponds to K t in the projectivity 
[#] 7^ [A], Similarly, ll z corresponds to l^ y 



39] OEKTEE AXD AXIS OF HOROLOGY 105 

EXERCISES 

1. There is one and only one projectivity of a one-dimensional form leaving 
invariant one and only one element 0, and transforming a given other element 
-4 to an element B. 

2. Two protective ranges on skew lines are always perspective. 

3. Prove Cor. 5, Theorem 27, without using the notion of involution. 

4. If MNAB-^MyA'B'* then MXAA'- J/JV.RB'. 

5. If P is any point of the axis of homology of two projective ranges 
M A"^' tliel1 tlie P r J ectiv % ^[-4] 77 **[#] is an involution. Dualize. 

6. Call the faces of one tetrahedron a ls 03, a 3 , a 4 and the opposite vertices 
*i 1? -dU, -4 3 , J 4 respectively, and similarly the faces and vertices of another tetra- 
hedron ft, ft, ft, ft and B 19 B s , B+ If 4i, J a , .4,, ^ lie on ft, ft, ft, ft 
respectively, and B^ lies on a x , B on a 2 , -5 3 on a 3 < then ^ lies on a 4 . Thus 
each of the two tetrahedra related in this fashion is both inscribed and cir- 
cumscribed to the other. 

7. Prove the theorem of Desargues (Chap. II) by the principle of pro- 
jectivity. 

8. Given a triangle ABC and a point A\ show how to construct two points 
B', C' such that the triangles ABC and A'B'C' are perspective from four 
different centers. 

9. If two triangles A^B^ and A^B 2 C Z are perspective, the three points 

(4 A, A^) = C,, (AiC v A&) = B 9 , (B X C 3 , J5 2 CO = A v 

if not collinear, form a triangle perspective with the first two, and the three 
centers of perspectivity are collinear. 

* 10. (a) If it is a projectivity in a pencil of points [J.] on a line a with inva- 
riant points A v AV and if [], \M ] are the pencils of points on two lines /, w 
through A^AQ respectively, show by the methods of Chap. Ill that there exist 
three points S 19 S 3 , S s such that we have 



where v (A ) = A' ; that S v S z , -4 2 are collinear ; and that ^ 2 , g , A ^ are collinear. 
(&) Using the fundamental theorem, show that there exists on the line ^A 2 
a point S such that we have 



(c) Show that (5) could be used as an assumption of projectivity instead of 
Assumption P ; i.e. P could be replaced by : If TT is a projectivity with fixed 
points ^t 1? A%, giving 7r(-4) = A' in a pencil of points 4], and [] is a pencil 
of points on a line I through A 19 there exist two points S v S 2 such that 

S l S, 



106 



THE FUNDAMENTAL THEOREM 



[CHAP. IV 



* 11. Show that Assumption P could be replaced by the corollary of 
Theorem 17. 

* 12. Show that Assumption P could be replaced by the following : If we 
have a projectivity in a pencil of points defined by the perspecthities 

[-X] = [] = [A"], 

and [JIT] is the pencil of points on the line 5^, there exist on the base of [] 
two points Si, S such that we have also 

=[*'] 

40. Types of collineations in the plane. We have seen in the 
proof of Theorem 10, Chap. Ill, that if O^O, is any triangle, there 
exists a collineation II leaving O v 2 , and O s invariant, and trans- 
forming any point not on a side of the triangle into any other such 





III 





FIG. 48 

point. By Theorem 18 there is only one such collineation II. By the 
same theorem it is clear that II is fully determined by the projec- 
tivity it determines on two of the sides of the invariant triangle, say 
2 3 and 1 S . Hence, if H x is a homology with center O l and axis 
2 O st which determines the same projectivity as II on the line O^, 
and if H 2 is a homology with center 3 and axis 1 8 , which deter- 
mines the same projectivity as II on the line 8 , then it is evident 

that 



<! TYPES OF COLLIXEATIOXS 107 

* 

It is also evideut that no point not a vertex of the invariant triangle 
can be fixed unless II reduces to a homology or to the identity. Such 
a transformation II when it is not a hornology is said to be of Type I 9 
and is denoted by Diagram / (fig. 48). 

EXERCISE 

Prove that two honiologies with, the same center and axis are commutative, 
and hence that two projectivities of Type I with the same invariant figure are 
commutative. 

Consider the figure of two points 1? 2 and two lines o v 2 , such 
that 1 and 2 are on o v and o l and 2 are on O r A collineation II 
which is the product of a homology H, leaving 3 and o. 2 invariant, 
and an elation E, leaving O l and o l invariant, evidently leaves this 
figure invariant and also leaves invariant no other point or line- If A 
and B are two points not on the lines of the invariant figure, and we 
require that TL(4\=sB 

this fixes the transformation (with two distinct double lines) among 
the lines at O v and the parabolic transformation among the lines at 2 , 
and thus determines II completely. Clearly if II is not to reduce to a 
homology or an elation, the line AB must not pass through O l or 2 . 
Such a transformation II, when it does not reduce to a homology or 
an elation or the identity, is said to be of Type II and is denoted by 
Diagram II (fig. 48). 

EXERCISE 

Two projective collineations of Type IT, having the same invariant figure, 
are commutative. 

DEFINITION The figure of a point and a line o on is called a 
lineal element Oo. 

A collineation having a lineal element as invariant figure must effect 
a parabolic transformation both on the points of the line and on the 
lines through the point. Suppose Aa and Sb are any two lineal ele- 
ments whose points are not on o or collinear with 0, and whose lines 
are not on or concurrent with o. Let E x be an elation with center 
and axis OA 9 which transforms the point (oa) to the point (ol). Let E 2 
be an elation of center (AB, o) and axis o, which transforms A to B. 
Then II = E 2 E X has evidently no other invariant elements than and o 
and transforms Aa to BL 



108 THE FUNDAMENTAL THEOEEM [CHAP. IV 

Suppose that aaother projeetivity II' would transfer Aa to Bl with 
Oo as only invariant elements. The transformation II 7 would evidently 
have the same effect on the lines of and points of o as II. Hence 
II'II- 1 would he the identity or an elation. But as ITrT 1 ^) = . it 
would be the identity. Hence II is the only projectivity which trans- 
forms Aa to b with Oo as only invariant. 

A transformation having as invariant figure a lineal element and no 
other invariant point or line is said to be of Type III, and is denoted 
by Diagram /// (fig. 48). 

A homology is said to be of Type /Fand is denoted by Diagram IV. 

An elation is said to be of Type V and is denoted by Diagram F. 

It will be shown later that any collineation can be regarded as be- 
longing to one of these five types. The results so far obtained may be 
summarized as follows : 

THEOREM 29. A protective collineation with given invariant figure F, 
if of Type I or II will transform any point P not on a line of F into 
any other such point not on a line joining P to a point of F; if of 
Type III will transform any lineal element Pp such that p is not on 
a point, or P on a line, of F into any other such element Qq; if of 
Type IV or V, will transform any point P into any other point on the 
line joining P to the center of the collineation. 

The r61e of Assumption P is well illustrated by this theorem. In case of 
each of the first three types the existence of the required collineation was proved 
by means of Assumptions A and E, together with the existence of a sufficient 
number of points to effect the construction. But its uniqueness was established 
only by means of Assumption P. In case of Types IV and F, both existence 
and uniqueness follow from Assumptions A and E. 

EXERCISES 

1. State the dual of Theorem 29. 

2. If the number of points on a line is^? + 1, the number of collineations 
with a given invariant figure is as follows : 

Type/, O-2)O~3). 
Type//, o- 2) (>-_!), 
Type m,p(p~ I)*. 
Type/F, p~2. 
TypeF, p-1. 

In accordance with the results of this exercise, when the number of points 
on a line is infinite it is said that there are oo 2 transformations of Type / or //; 
oo 3 of Type ///; and oo 1 of Types /Fand F. 



CHAPTER V* 

CONIC SECTIONS 

41. Definitions. Pascal's and Brianchon's theorems. 

DEFINITION. The set of all points of intersection of homologous 
lines of two protective, nonperspective flat pencils which are on the 
same plane but not on the same point is called a point conic (fig. 49). 
The plane dual of a point conic is called a line conic (fig. 50). The 
space dual of a point conic is called a cone of planes; the space dual 





. 49 



FIG. 50 



of a line conic is called a cone of lines. The point through which 
pass all the lines (or planes) of a cone of lines (or planes) is called 
the vertex of the cone. The point conic, line conic, cone of planes, 
and cone of lines are called one-dimensional forms of the second degree.^ 

The following theorem is an immediate consequence of this defi- 
nition. 

THEOREM 1. The section of a cone of lines ly a plane not 6n the 
vertex of the cone is a point conic. The section of a cone of planes by 
a plane not on the vertex is a line conic. 

Now let A^ and B l be the centers of two flat pencils defining a 
point conic. They are themselves, evidently, points of the conic, for the 
line A^BI regarded as a line of the pencil on A 1 corresponds to some 
other line through B^ (since the pencils are, by hypothesis, protective 

* All the developments of this chapter are on the basis of Assumptions A, E, F, 
and H . 

t A fifth one-dimensional form a self-dual form of lines in space called the 
regulus will be defined in Chap. XL This definition of the first four one-dimen- 
sional forms of the second degree is due to Jacob Steiner (1796-1863). Attention 
will be called to other methods of definition in the sequel. 

109 



110 



CONIC SECTIONS 



[CHAP. V 



but not perspective), and the intersection of these homologous lines 
is B r The conic is clearly determined by any other three of its 
points, say A 2 , B z , <7 2 , because the projectivity of the pencils is then 

determined by 

(AJ*&) 7: B l (AJ}&) 

(Theorem 17, Chap. IV). 

Let us now see how to determine a sixth point of the conic on a 
line through one of the given points, say on a line Z through B y If the 
line I is met by the lines A^, A^C^ B^A^ B^G Z in the points S> T, U, A 




FIG. 51 



-j^ UB^A. The other 



respectively (fig. 5 1), we have, by hypothesis, S 
double point of this projectivity, which we will call C v is given by the 
quadrangular set Q^ST, C^AU) (Theorem 23, Chap. IV). A quad- 
rangle which determines it may be obtained as follows : Let the lines 
A^B^ and A^BZ meet in a point (7, and the lines AC and A^C^ in a 
point B ; then the required quadrangle is A^A^CB^ and C l is determined 
as the intersection of AJB with L 



C l Trill coincide with 2 , if an d Q ^J *f B is on A 2 B^ (fig. 52). This means 
that A C, A^Cty and A^B 2 are concurrent in B. In other words, A must be the 
point of intersection of B X C 2 with the line joining C = (A^B^ (A^) and 
B = (JljCy^o&j)* an< i I must be the line joining B z anil A . This gives, then, 
a construction for a line which meets a given conic in only one point. 

The result of the preceding discussion may be summarized as 
follows : TJie four points A z , B 2 , <7 2 , C l are points of a point conic 



HI] PASCAL'S THEOREM 111 

determined ly two projectile pencils on A^ and B v if and only if the 
three points C' = (.4^,) (A&), B = (A^C.) (A 2 C^, A = (B^) (B.,C\) are 
collinear. The three points in question are clearly the intersections 
of pairs of opposite sides of the simple hexagon A^B^A^B^C y 

Since A v B v C^ may be interchanged with -.4.,, J5 2 , C 2 respectively 
in the above statement, it follows that A v Z? 1? C v C n are points of a 
conic determined by protective pencils on A 2 and B 2 . Thus, if C^ is 
any point of the first conic, it is also a point of the second conic, 
and vice versa. Hence we have established the following theorem : 

THEOREM 2. STEIXER'S THEOREM. If A and B are any two given 
points of a conic, and P is a variable point of this conic, ice have 



In view of this theorem the six points in the discussion may be 
regarded as any six points of a conic, and hence we have 

THEOREM 3. PASCAL'S THEOREM.* The necessary and sufficient con- 
dition that siQTjwwsno tJiree of u'hicJi are collinear, be points of 
the same conic is that the three pairs of opposite sides of a simple 
hexagon of ivhicli they are vertices shall meet in collinear points.^ 

The plane dual of this theorem is ^ 

THEOREM 3'. jteiAXCHoy's THEOREM. Tlie necessary and sufficient 
condition that six lines, no three of ichich are concurrent, be lines of 
a line conic is that the lines joining the three pairs of opposite vertices 
of any simple hexagon of whicli the given lines are sides, shall be 
concurrent.^ 

As corollaries of these theorems we have 

COROLLARY 1. A line in the plane of a point conic cannot have more 
than two points in common with the conic. 

COROLLARY 1'. A point in the plum of a line conic cannot be on 
more than two lines of the conic. 

* Theorem 3 was proved by B. Pascal in 1640 when only sixteen years of age. 
He proved it first for the circle and then obtained it for any conic by projection 
and feection. This is one of the earliest applications of this method. Theorem 8' 
was first given by C. J. Brianchon in 1806 (Journal de 1'ficole Polytechnique, 
Vol. VI, p. 301). 

t The line thus determined by the intersections of the pairs of opposite sides of 
any simple hexagon whose vertices are points of a point conic is called the Pascal 
line of the hexagon. The dual construction gives rise to the Brianchon poM of a 
hexagon whose sides belong to a line conic. 



112 COOTC SECTIONS [OHAP.V 

Also as immediate corollaries of these theorems we have 
THEOREM 4. There is one and only one point conic containing Jim 

given points of a plane no three of which are collinear. 

THEOREM -4'. There is one and only one line conic containing Jive 

given lines of a plane no three of which are concurrent. 

EXERCISES 

1. What are the space duals of the above theorems? 

2. Prove Brianehon's theorem without making use of the principle of 
duality. 

3. A necessary and sufficient condition that six points, no three of "which 
are collinear, be points of a point conic, is that they be the points of inter- 
section (&'), (fa'), (ca'), (a') ( c&/ )> ( a O f tne si des #j &> c an< i a/ ? &'> c/ f two 
perspective triangles, in which a and a', "b and &', c and c' are homologous. 

42. Tangents. Points of contact. DEFINITION. "A line p in the 
plane of a point conic which meets the point conic in one and only 
one point P is called a tangent to the point conic at P. A point P in 
the plane of a line conic through which passes one and only one line 
p of the line conic is called a point of contact of the line conic on'jp. 

THEOREM 5. Through any point of a point conic there is one and 
only one tangent to the point conic. 

Proof. If PQ is the given point of the point conic and P is any 
other point of the point conic, while P is a variable point of this 
conic, we have, by Theorem 2, 



Any line through P Q meets its homologous line of the pencil on 7^ in 
a point distinct from P, except when its homologous line is P^. 
Since a projectivity is a one-to-one correspondence, there is only one 
line on -ZJ which has I(P^ as its homologous line. 

THEOREM 5'. 0% any line of a line conic there is one and only one 
point of contact of the line conic. 

This is the plane dual of the preceding theorem. 

EXERCISE 
Give the space duals of the preceding definitions and theorems. 

Returning now to the construction in the preceding section for the 
points of a point conic containing five given points, we recall that 



42] TA2s T GE]N T TS 113 

the point of intersection Ci of a line I through B was determined by 
the quadrangular set Q(J3 Z ST, C^U). The points and O l can, 
by the preceding theorem, coincide on one and only one of the lines 
through B z * For this particular line Z, A becomes the intersection 




. 52 



of the tangent at B z with I, and the collinearity of the points A, B 9 C 
may be stated as follows : 

THEOREM 6. If the vertices of a simple plane five-point are points 
of a point conic, the tangent to the point conic at one of the vertices 
meets the opposite, side in a point collinear with the points of inter- 
section of the other two pairs of nonadjacent sides. 

This theorem, by its derivation, is a degenerate case of Pascal's 
theorem. It may also be regarded as a degenerate case in its state- 
ment, if the tangent be thought of as taking the place of one side 
of the simple hexagon. 

It should be clearly understood that the theorem has been obtained by 
specializing the figure of Theorem 3, and not by a continuity argument. 
The latter would be clearly impossible, since our assumptions do not require 
the conic to contain more than a finite number of points. 

Theorem 6 may be applied to the construction of a tangent to 
a point conic at any one of five given points P^ P%, P%, P^ jg of the 
point conic (fig. 53). By this theorem the tangent p l at JJ must be 

* As explained in the fine print on page 110, this occurs when I passes through 
the point of intersection of BiOz with the line joining C (A^Bz) (A%Bi) and 



114 



CONIC SECTIONS 



[CHAP. V 



such that the points 2\(P^) = -i (%%) (%**) - B > and (%P 3 ) (P B P^ = C 
are collinear. But B and C are determined by ^, P%, ^, -^, J^, and 
hence p l is the line joining ^ to the intersection of the lines BC 
and P & %. 

Cl 




FIG. 53 ' 

In like manner, if P v P%, P^, P, and jp 1 are given, to construct the 
point P$ on any line I through P of a point conic containing P^ P Z) P SJ P 
and of which p l is the tangent at P 19 we need only determine the points 
A =1\(P Z P^ B = I (JJJ?), and C = (AB) (P 2 P 3 ) ; then P^C meets f in ^ 
(fig. 53). 




FIG. 54 



In case I is the tangent p 4 at P^ P 5 coincides with ^ and the fol- 
lowing points are collinear (fig. 54) : 



42] TAXGEXTS 115 

Hence we have the following theorem : 

THEOREM 7. If the vertices P 13 P 2 , %, P of a simple quadrangle are 
points of a point conic, the tangent at % and the side P S P, the tangent 
at P and the side JJJEJ, and the pair of sides JJJ* and RE. meet in three 
collinear points. 

If %, , P B , P 5 and the tangent p l at P are given, the construction 
determined by Theorem -3 for a point P of the point conic on a line / 
through P 3 is as follows (fig. 53): Determine C-=(P 1 P 5 )(P 2 P B ), A = pj, 
and B = (A C)(P^ ; then P,B meets I in %. 

In case I is the tangent at P z , P coincides with P 3 and we have the 
result that C = (P 1 J)(P 2 P B ), A^p^p,, B = (P 1 K)(P 5 P) are collinear 
points, which gives 




PIG. 5o 

THEOREM 8. If the vertices of a complete quadrangle are points of 
a point conic 9 the tangents at a pair of vertices meet in a poi%t of the 
line joining the diagonal points of the quadrangle which are not on 
the side joining the two vertices (fig. 55). 

The last two theorems lead to the construction for a point conic 
of which there are given three points and the tangents at two of 
them. Eeverting to the notation of Theorem 7 (fig. 54), let the given 
points be P, P P z and the given tangents be p^ p r Let I be any line 
through P y If PZ is the other point in which I meets the point conic, 
the points A^p^(P^ 9 B^p^(P^, and C =(%%)(%%) are collinear. 
Hence, if C^l(P^ and B=p(AC), then F 2 is the intersection of I 
with B%. 

In case I is the tangent p z at JJ, the points JEJ and P B coincide, and 
the points 



116 CONIC SECTIONS [CHAP.V 



are collinear. Hence the two triangles P^P^ and p^sPt are per- 
spective, and we obtain as a last specialization of Pascal's theorem 
(fig. 56) 

THEOREM 9. A trianyle whose vertices are points of a point conic 
is perspective with the triangle formed "by the tangents at these points, 
the tangent at any vertex "being ho7nologous with the side of the first 
triangle which does not contain this vertex. 

COROLLARY. If I(, J^, P are three points of a point conic, the lines 
P^PV P$P are harmonic with the tangent at P z and the line joining P% 
to the intersection of the tangents at jfj and P. 

Proof. This follows from the definition of a harmonic set of lines, 
on considering the quadrilateral P^A, AB> BP^ PJ( (fig. 56). 




FIG. 56 

43. The tangents to a point conic form a line conic. If P v P^ P 8 , P 

are points of a point conic and p v p Z9 p s , p are the tangents to the 
conic at these points respectively, then (by Theorem 8) the line join- 
ing the diagonal points (JJ JJ) (P B I%) and (JfJJIJ) (J%I) contains the inter- 
section of the tangents p v p s and also the intersection of p 2> p^. This 
line is a diagonal line not only of the quadrangle Jf^J^, but also of 
the quadrilateral PiPzPsP^ Theorem 8 may therefore be stated in 
the form: 

THEOREM 10. Tlu complete quadrangle formed ly four points of 
a point conic and the complete quadrilateral of the tangents at these 
points have the same diagonal triangle. 

Looked at from a slightly different point of view, Theorem 8 gives 
also 

THEOREM 11. The tangents to a point conic form a line conic. 



43] TAXGEXTS 117 

Proof. Let JJ, P 2 , P 3 be any three fixed points on a conic, and let P 
be a variable point of this conic. Let p v p*> Pv P be respectively the 
tangents at these points (fig. 57). By the corollary of Theorem 28, 
Chap. IV, JJJE} is the axis of homology of the projectivity between the 
pencils of points on p 1 and p 2 defined by 



But by Theorem 10, if Q=(P 1 P 2 ) (P Z P), the points pj> zi PiP& and Q are 
collinear. For the same reason the points p z p z , pp v Q are collinear. 
It follows, by Theorem 28, Chap. IV, that the homolog of the variable 




PIG. 57 



point Pip is p 2 p> i.e. p is the line joining pairs of homologous points 
on the two lines p v p^ so that the totality of the lines p satisfies the 
definition of a line conic. 

COROLLARY. The center of homology of the projectivity P l [P] -^ P% [P] 
determined "by the points P of a point conic containing P^ 1^ is the 
intersection of the tangents at 1^ P^. The axis of honwlogy of the 
projectimty p l [jp] -^ p% [p] determined by the lines p of a line conic 
containing the lines p lt p 2 is the line joining the points of contact 

f Pv P* 

THEOREM 12. If J^ is a fixed and P a variable point of a point 

conic, and p v p are the tangents at these two points respectively, then 



118 CONIC SECTIONS [CHAP.V 

Proof. Using the notation of the proof of Theorem 11 (fig. 57), 
we have 



where Q is always on JJJJ. But we also have 



and, by Theorem 11, 

Combining these projectivities, we have 



The plane dual of Theorem 11 states that the points of contact of 
a line conic form a point conic. In view of these two theorems and 
their space duals we now make the following 

DEFINITION. A conic section or a conic is the figure formed by a 
point conic and its tangents. A cone is the figure formed by a cone 
of lines and its tangent planes. 

The figure formed by a line conic and its points of contact is then 
likewise a conic as defined above ; i.e. a conic (and also a cone) is a 
self-dual figure. 

The duals of Pascal's theorem and its special cases now give us a 
set of theorems of the same consequence for point conies as for line 
conies. We content ourselves with restating Brianchon's theorem 
(Theorem 3') from this point of view. 

BRIANCHON'S THEOREM. If the sides of a simple hexagon are tan- 
gents to a conic, the lines joining opposite vertices are concurrent; 
and conversely. 

It follows from the preceding discussion that in forming the plane 
duals of theorems concerning conies, the word conic is left unchanged, 
while the words point (of a conic) andjJa^$wL(of a conic) are inter- 
changed. "We shall also, in the future, make use of the phraseTa conic 
passes through a point P, and P is on the conic, when P is a point 
of a conic, etc. 

DEFINITION. If the points of a plane figure are on a conic, the figure 
is said to be inscribed in the conic; if the lines of a plane figure 
are tangent to a conic, the figure is said" to be circumscribed about 
the conic. 



TA]S T GE^TS 119 

EXERCISES 

1. State the plane and space duals of the special cases of Pascal's theorem. 

2. Construct a conic, given (1) five tangents, (2) four tangents and the 
point of contact of one of them, (3) three tangents and the points of contact 
of two of them. 

3. ABX is a triangle whose vertices are on a conic, and a, I. x are the tan- 
gents at A, B, X respectively. If .4, L are given points and X is variable, 
determine the locus of (1) the center of perspectivity of the triangles ABX 
and abx ; (2) the axis of perspectivity. 

JS 4. X, Y, Z are the vertices o a variable triangle, such that X, Y are always 
on two given lines a, I respectively, while the sides XY, ZX, ZY always pass 
through three given points P, A, B respectively. Show that the locus of the 
point Z is a point conic containing A,B,D = (ah), J/ = (AP)l, and JV~ = (BP)a 
(Maclaurin's theorem). Dualize. (The plane dual of this theorem is known 
as the theorem of Braiken ridge.) 

5. If a simple plane fl-point varies in such avray that its sides always pass 
through n given points, while n 1 of its vertices are always on n 1 given 
lines, the nth vertex describes a conic (Poncelet). 

6. If the vertices of two triangles are on a conic, the six sides of these two 
triangles are tangents of a second conic ; and conversely.' Corresponding to 
every point of the first conic there exists a triangle having this point as a 
vertex, whose other two vertices are also on the first conic and whose sides 
are tangents to the second conic. Dualize. 

7. If two triangles in the same plane are perspective, the points in \\hich 
the sides of one triangle meet the nonhomologous sides of the other are on 
the same conic ; and the lines joining the vertices of one triangle to the non- 
homologous vertices of the other are tangents to another conic. 

8. If A , B, C, D be the vertices of a complete quadrangle, whose sides 
AB, AC, AD, BC, BD, CD are cut by a line in the points P, Q, R, S, T, V 
respectively, and if E, F^ff, K, L, JJf are respectively the harmonic conjugates 
of these points with respect to the pairs of vertices of the quadrangle so that 
we have H (AB, PE), H (A C, QF), etc., then the six points E, F, G t A", , AT 
are on a conic which also passes through the diagonal points of the quadrangle 
(Holgate, Annals of Mathematics, Ser. 1, Vol. Til (1893), p. 73). 

9. If a plane a cut the six edges of a tetrahedron in six distinct points, 
and the harmonic conjugates of each of these points with respect to the two 
vertices of the tetrahedron that lie on the same edge are determined, then the 
lines joining the latter six points to any point of the plane a are on a cone, 
on which are also the lines through and meeting a pair of opposite edges of the 
tetrahedron (Holgate, Annals of Mathematics, Ser. 1, Tol. VII (1893), p. 73). 

10. Given four points of a conic and the tangent at one of them, construct 
the tangents at the other three points. Dualize. 

11. A, A', B, B' are the, vertices of a quadrangle, and m, n are two lines 
in the plane of thja quadrangle which meet on AA'. M Is a variable point 



120 



CONIC SECTIONS 



[CHAP. V 



OIL m, the lines BM, B'M meet n in the points N, N' respectively ; the lines 
ANj A'N' meet in a point P. Show that the locus of the lines PJ\1 is a line 
conic, -which contains the lines m, p = P(n, BB'), and also the lines A A', BB', 
A'B'y AB (Amodeo, Lezioni di Geometria Projettiva, Naples (1905), p. 331). 

12. Use the result of Ex. 11 to give a construction of a line conic deter- 
mined by five given lines, and show that by means of this construction it is 
possible to obtain two lines of the conic at the same time (Amodeo, loc. cit.). 

13. If a, 5, c are the sides of a triangle whose vertices are on a conic, and 
m, m' are two lines meeting on the conic which meet a, &, c in the points A, B, C 
and A', B', C' respectively, and which meet the conic again in N 9 N' respec- 
tively, we have ABCNj^A'B'C'N' (cf. Ex. 6). 

14. If A, B, C, D are points on a conic and a, 5, c, d are the tangents to 
the conic at these points, the four diagonals of the simple quadrangle ABCD 
and the simple quadrilateral abed are concurrent. 

44. The polar system of a conic. 

THEOREM 13. If P is a point in THEOREM 13'. If pis a line in the 
the plane of a conic, lut not on the plane of a conic, but not tangent to 



conic, the points of intersection of 
the tangents to the conic at all the 
pairs of points which are collinear 
with P are on a line, which also con*- 
tains the harmonic conjugates of P 
with respect to these pairs of points. 



the conic, the lines joining the points 
of contact of pairs of tangents to the 
conic which meet on p pass through 
a point P, through which pass also 
the harmonic conjugates ofp with 
respect to these pairs of tangents. 




R 



. 58 



Proof. Let P v P% and P^, P be two pairs of points on the conic which 
are collinear with P, and let p v p z be the tangents to the conic at JJ, P 2 
respectively (fig. 58). If 2> v D 2 are the points (%%)(%%) an 



] POLAR SYSTEM 121 

respectively, the line D^D^ passes through the intersection Q of jt\ 3 $ 
(Theorem 8). Moreover, the point P r in which Z^D* meets 2J2* is the 
harmonic conjugate of P with respect to P 13 R (Theorem 6, Chap. IT). 
This shows that the line D^D* QP ! is completely determined by the 
pair of points P^ P,. Hence the same line QP' is obtained by replacing 
P 3> P by any other pair of points on the conic collinear with P, and 
distinct from P v P,. This proves Theorem 13. Theorem 13' is the 
plane dual of Theorem 13. 

DEFINITION. The line thus asso- DEFINITION. The point thus 

ciated with any point P in the associated with any line p in the 

plane of a conic, but not on the plane of a conic, but not tangent 

conic, is called the polar of P to the conic, is called thereof p 

with respect to the conic. If P with respect to the conic. If p is 

is a point on the conic, the polar a tangent to the conic, the pole is 

is defined as the tangent at P. defined as the point of contact of P. 

THEOKEM 14. TJie line joining THEOREM 14'. The point of 

two diagonal points of any com- intersection of two diagonal lines 

plete quadrangle whose vertices of any complete quadrilateral 

are points of a conic is the polar wJiose sides are tangent to a conic 

of the other diagonal point with is the pole of the other diagonal 

respect to the conic. line with respect to the conic. 

Proof. Theorem 14 follows immediately from the proof of Theo- 
rem 13. Theorem 14' is the plane dual of Theorem 14 

THEOREM 15. The polar of a THEOREM 15'. Tlie pole of a 

point P with respect to a conic line p with respect to a conic is 

passes through the points of con- on the tangents to the conic at the 

tact of the tangents to the conic poi/its in which p meets the conic, 

through P, if such tangents exist, if such paints exist 

Proof. Let P l be the point of contact of a tangent through. P, and 
let P%> PS be any pair of distinct points of the conic collinear with P. 
The line through JJ and the intersection of the tangents at jg, P% 
meets the line J%P% in the harmonic conjugate of P with, respect to 
j?fjf (Theorem 9, Cor.). But the line thus determined is the polar of P 
(Theorem 13). This proves Theorem 15. Theorem 15' is its plane dual 

THEOREM 16. If p is the polar of a point P with respect to a conic, 
P is the pole of p with respect to the same conic. 



122 COXIC SECTIONS [CHAP.V 

If P is not on the conic, this follows at once by comparing Theo- 
rem 13 with Theorem 13'. If P is on the conic, it follows immediately 
from the definition. 

THEOREM 17. If the polar of a point P passes through a point Q, 
the polar of Q passes through P. 

Proof. If P or Q are on the conic, the theorem is equivalent to 
Theorem 15. If neither P nor Q is on the conic, let PP^ be a line 

P 




59 



meeting the conic in two points, P^ P 2 . If one of the lines P^Q, P 2 Q 
is a tangent to the conic, the other is also a tangent (Theorem 13); 
the line P^ = PP is then the polar of Q, which proves the theorem 
under this hypothesis. If, on the other hand, the lines I$Q,P 2 Q meet 
the conic again in the points P& P respectively (fig. 59), the point 
(IJJFJ) (P Z P^ is on the polar of Q (Theorem 14). By Theorems 13 and 14 
the polar of (J?-E) (P Z P^) contains the intersection of the tangents at 
P v R and the point Q. By hypothesis, however, and Theorem 13, the 
polar of P contains these points also. Hence we have (P^ (P%P^) = P, 
which proves the theorem. 

COROLLAEY 1. If two vertices of a triangle are the yoles of their 
opposite sides with respect to a conic, the third vertex is the pole of 
its opposite side. 

DEFINITION. Any point on the polar of a point P is said to be 
conjugate to P with regard to the conic; and any line on the pole 



POLAE SYSTEM 123 

of a line p is said to be conjugate to p with regard to the conic. 
The figure obtained from a given figure in the plane of a conic by 
constructing the polar of every point and the pole of every line of 
the given figure with regard to the conic is called the polar or polar 
reciprocal of the given figure with regard to the conic.* A triangle, 
of which each vertex is the pole of the opposite side, is said to be 
self -polar or self-conjugate with regard to the conic. 

COROLLARY 2. TJie diagonal triangle of a complete quadrangle ichose 
vertices are on a conic, or' of a complete quadrilateral whose sides are 
tangent to a conic, is self -polar with regard to the co/iic; and, conversely, 
every self -polar triangle is the diagonal triangle of a complete quad- 
rangle whose points are on the conic, and of a complete quadrilateral 
whose sides are tangent to the conic. Corresponding to a given self-polar 
triangle, one vertex or side of such a quadrangle or quadrilateral may 
be chosen arbitrarily on the conic. 

Theorem 17 may also be stated as follows : If P is a variable point 
on a line q, its polar p is a variable line through the pole Q of q. In the 
special case where q is a tangent to the conic, we have already seen 
(Theorem 12) that we have 

[^AM- 

If Q is not on q, let A (fig. 60) be a fixed point on the conic, a the 
tangent at A, JT the point (distinct from A, if AP is not tangent) in 
which AP meets the conic, and x the tangent at X. "We then have, by 
Theorem 12, 



By Theorem 13, (ax) is on p, and hence p = Q (ax). Hence we have 

[*]*[*] 

If P ! is the point pq, this gives 



But since the polar of P ! also passes through P, this projectivity is 
an involution. The result of this discussion may then be stated as 
follows : 

* It was by considering the polar reciprocal of Pascal's theorem that Brianchon 
derived the theorem named after him. This method was fully developed "by Poncelet 
and Gergonne in the early part of the last century in connection with the principle 
of duality. 



124 C03STC SECTIONS [CHAP.Y 

THEOREM 18. On any line not a tangent to a given conic the pairs 
of conjugate points are pairs of an involution. If the line meets the 
conic in two points, these points are the doulle points of the involution. 

COROLLARY. As a point P varies over a pencil of points, its polar 
with respect to any conic varies over a projectile pencil of lines. 




60 



. DEFINITION. The pairing of the points and lines of a plane brought 
about by associating \vith every point its polar and with every line its 
pole with respect to a given conic in the plane is called a polar system. 

EXERCISES 

1. If in a polar system two points are conjugate to a third point A , the 
line joining them is the polar of A. 

2. State the duals of the last two theorems. 

?*' 3. If a and Z> are two nonconjugate lines in a polar system, every point A 
of a has a conjugate point B on 5. The pencils of points [A] and B] are 
protective ; they are perspective if and only if a and I intersect on the conic 
of the polar system. 

< 4. Let A be a point and 5 a line not the polar of A with respect to a given 
conic, but in the plane of the conic. If on any line I through A we determine 
that point P which is conjugate with the point $, the locus of P is a conic 
passing through A and the pole B of 5, unless the line AB is tangent to the 



] POLAB SYSTEM 125 

conic, In which case the locus of P is a line. If AB is not tangent to the conic, 
the locus of P also passes through the points in which b meets the given conic 
(if such points exist), a-nd also through the points of contact of the tangents to 
the given conic through A (if such tangents exist). Dualize (Reye-Holgate, 
Geometry of Position, p. 106). 

5. If the vertices of a triangle are on a given conic, any line conjugate to 
one side meets the other two sides in a pair of conjugate points. Conversely, 
a line meeting two sides of the triangle in conjugate points passes through 
the pole of the side (von Staudt). 

6. If two lines conjugate with respect to a conic meet the conic in two 
pairs of points, these pairs are projected from any point on the conic by a 
harmonic set of lines, and the tangents at these pairs of points meet any 
tangent in a harmonic set of points. 

7. ^Vith a given point not on a given conic as center and the polar of this 
point as axis, the conic is transformed into itself by a homology of period two. 

8. The Pascal line of any simple hexagon whose vertices are on a conic is 
the polar with respect to the conic of the Brianchon point of the simple hexagon 
whose sides are the tangents to the conic at the vertices of the first hexagon. 

9. If the line joining two points A, B, conjugate with respect to a conic, 
meets the conic in two points, these two points are harmonic with A 9 B. 

10. If in a plane there are given two conies Cf and C|, and the polars oi 
all the points of Cf with respect to C| are determined, these polars are the 
tangents of a third conic. 

11. If the tangents to a given conic meet a second conic in pairs of points, 
the tangents at these pairs of points meet on a third conic. 

12. Given five points of a conic (or four points and the tangent through 
one of them, or any one of the other conditions determining a conic), sho^w 
how to construct the polar of a given point with respect to the conic, 

13. If two pairs of opposite sides of a complete quadrangle are pairs ot 
conjugate lines with respect to a conic, the third pair of opposite sides are 
conjugate -with respect to the conic (Hesse). 

14. If each of two triangles in a plane is the polar of the other with respect 
to a conic, they are perspective, and the axis of perspectivity is the polar of the 
center of perspectivity (Chasles). 

15. Two triangles that are self -polar with respect to the same conic have 
their six vertices on a second conic and their six sides tangent to a third 
conic (Steiner). 

16. Regarding the Desargues configuration as composed of a quadrangle 
and a quadrilateral mutually inscribed (cf. 18, Chap. II), show that the 
diagonal triangle of the quadrangle is perspective with the diagonal triangle 
of the quadrilateral. 

17. Let A , B be any two conjugate points with respect to a conic, and let 
the lin^s AM, BM joining them to an arbitrary point of the conic meet the 
latter again in the points C, D respectively. The lines AD, BC will then meet 
on the conic, and the lines CD and AB are conjugate. Dualize. 



126 CONIC SECTIONS [CHAP.V 

45. Degenerate conies. For a variety of reasons it is desirable to 
regard two coplanar lines or one line (thought of as two coincident 
lines) as degenerate cases of a point conic; and dually to regard 
two points or one point (thought of as two coincident points) as 
degenerate cases of a line conic. This conception makes it possible 
to leave out the restriction as to the plane of section in Theorem 1. 
For the section of a cone of lines by a plane through the vertex of 
the cone consists evidently of two (distinct or coincident) lines, i.e. 
of a degenerate point conic ; and the section of a cone of planes by 
a plane through the vertex of the cone is the figure formed by some 
or all the lines of a flat pencil, i.e. a degenerate line conic. 

EXERCISE 

Dualize in all possible ways the degenerate and nondegenerate cases of 
Theorem 1. 

Historically, the first definition of a conic section was given by the ancient 
Greek geometers (e.g. Mensechmus, about 350 B.C.), who defined them as the 
plane sections of a "right circular cone." In a later chapter we will show 
that in the " geometry of reals " any nondegenerate point conic is protectively 
equivalent to a circle, and thus that for the ordinary geometry the modern 
projective definition given in 41 is equivalent to the old definition. We are 
here using one of the modern definitions because it can be applied before devel- 
oping the Euclidean metric geometry. 

Degenerate conies would be included in our definition (p. 109), if 
we had not imposed the restriction on the generating projective 
pencils that they be nonperspective ; for the locus of the point of 
intersection of pairs of homologous lines in two perspective flat 
pencils in the same plane consists of the axis of perspectivity and 
the line joining the centers of the pencils. 

It will be seen, as we progress, that many theorems regarding non- 
degenerate conies apply also when the conies are degenerate. For 
example, Pascal's theorem (Theorem 3) becomes, for the case of a 
degenerate conic consisting of two distinct lines, the theorem of 
Pappus already proved as Theorem 21, Chap. IV (cf. in particular the 
corollary). The polar of a point with regard to a degenerate conic 
consisting of two lines is the harmonic conjugate of the point with 
respect to the two lines (cf. the definition, p. 84, Ex. 7). Hence the 
polar system of a degenerate conic of two lines (and dually of two 
points) determines an involution at a point (on a line). 



45,40] THEOREM OF DBS ARGUES 127 

EXERCISES 

1. State Brianchon's theorem (Theorem 3') for the case of a degenerate 
line conic consisting of two points. 

2. Examine all the theorems of the preceding sections -with reference to 
their behavior when the conic in question becomes degenerate. 

46. Desargues's theorem on conies. 

THEOREM 19. If the vertices of a complete quadrangle are oil a conic 
which meets a line in two points, the latter are a pair in the invo- 
lution determined on the line ly the pairs of o^iosite sides of the 
quad/ 'anyk >* 

Proof. Reverting to the proof of Theorem 2 (fig. 51), let the line 
meet the conic in the points B v C and let the vertices of the quad- 
rangle be A v A*, B v C 2 . This quadrangle determines on the line an 
involution in which S 9 A and T, U are conjugate pairs. But in the 
proof of Theorem 2 we saw that the quadrangle A^A^BC determines 
Q^ST, C^U). Hence the two quadrangles determine the same 
involution on the line, and therefore B 2 , C l are a pair of the involution 
determined by the quadrangle A^B^Cy 

Since the quadrangles A^AJS^ and A^BC determine the same 
involution on the line when the latter is a tangent to the conic, we 
have as a special case of the above theorem : 

COROLLARY. If the vertices of a complete quadrangle are on a conic, 
the pairs of opposite sides meet the tangent at any other point in pairs 
of an involution of which the point of contact of the tanyent is a double 
point. 

The Desargues theorem leads to a slightly different form of statement for 
the construction of a conic through five given points : On any line through 
one of the points the complete quadrangle of the other four determine an 
involution ; the conjugate in this involution of the given point on the line 
is a sixth point on the conic. 

As the Desargues theorem is related to the theorem of Pascal, so 
are certain degenerate cases of the Desargues theorem related to the 
degenerate cases of the theorem of Pascal (Theorems 6, 7, 8, 9). Thus 
in fig. 53 we see (by Theorem 6) that the quadrangle BCRP^ deter- 
mines on the line P^ an involution in which the points P^ P of the 
conic are one pair, while the points determined by p v P^ and those 

* First given by Desargues in 1639 ; cf. OEuvres, Paris, Vol. I (1864), p. 188. 



128 CONIC SECTIONS [CHAP.V 



determined by % P^ P^ are two other pairs. This gives the following 
special case of the theorem of Desargues : 

THEOBEM 20. If the vertices of a triangle are on a conic, and a line I 
meets the conic in two points, the latter are a pair of the involution 
determined on I by the pair of points in which two sides of the triangle 
meet Z, and the pair in which the third side and the tangent at the 
opposite vertex meet I In case I is a tangent to the conic, the point of 
contact is a double point of this involution. 

In terms of this theorem we may state the construction of a conic through 
four points and tangent to a line through one of them as follows : On any line 
through one of the points which is not on the tangent an involution is deter- 
mined in which the tangent and the line passing through the other two points 
determine one pair, and the lines joining the point of contact to the other two 
points determine another pair. The conjugate of the given point on the line 
in this involution is a point of the conic. 

A further degenerate case is derived either from Theorem 7 or 
Theorem 8. In fig. 54 (Theorem 7) let I be the line P 2 P 3 . The quad- 
rangle ABP^ determines on I an involution in which JJ, P B are one 
pair, in which the tangents at P l} P determine another pair, and in 
which the line P^ determines a double point. Hence we have 

THEOREM 21. If a line I meets a conic in two points and J%, P are 
any other two points on the conic, the points in which I meets the conic 
are a pair of an involution through a double point of which passes the 
line P^P and through a pair of conjugate points of which pass the 
tangents at P lt P^ If I is tangent to the conic, the point of contact is 
the second double point of this involution. 

The construction of the conic corresponding to this theorem may be stated 
as follows : Given two tangents and their points of contact and one other point 
of the conic. On any line I through the latter point is determined an involution 
of which one double point is the intersection with I of the line joining the two 
points of contact, and of which one pair is the pair of intersections with I of 
the two tangents. The conjugate in this involution of the given point of the 
conic on Us a point of the conic. 

EXERCISE 

State the duals of the theorems in this section. 

47. Pencils and ranges of conies. Order of contact. The theorems 
of the last section and their plane duals determine the properties of 
certain systems of conies which we now proceed to discuss briefly. 



47] 



PENCILS AXD RANGES 



129 



DEFINITION. The set of all conies 
through the vertices of a complete 
quadrangle is called a pencil of 
conies of Type I (fig. 61). 

Theorem 19 and its plane dual 
THEOREM 22. Any line (not 
tlirougli a vertex of the deter- 
mining quadrangle) is met ly the 
conies of a pencil of Type I in the 
pairs of an involution* 



DEFINITION. The set of all conks 
tangent to the sides of a complete 
quadrilateral is called a range of 
conies of Type I (fig. 62). 
give at once : 

THEOREM 22'. The tangents 
tlirougli any point (not on a side 
of the determining quadrilateral) 
to the conies of a range of Type I 
are the pairs of an involution. 




FIG. 63 

GOBOLLABY. Through a gen- 
eral^ point in the plane there is 
one and only one, and tangent to 
a general line there are two or no 
conies of a given pencil of Type I. 



EIG. 64 

COBOLLABY. Tangent to a gen- 
eral line in the plane there is one 
and only one, and through a gen- 
eral point there are two or no 
conies of a gimn range of Type I. 



* This form of Desargues's theorem is due to Ch. Sturm, Annales de Mathe*ma- 
tiques, Vol. XVII (1826), p. 180. 

t The vertices of the quadrangle are regarded as exceptional points. 



130 



CONIC SECTIONS 



[CHA.P. V 



DEFINITION. The set of all conies 
through the vertices of a triangle 
and tangent to a fixed line through 
one vertex is called a pencil of 
conies of Type II (fig. 63). 



DEFINITION. The set of all conies 
tangent to the sides of a triangle 
and passing through a fixed point 
on one side is called a range of 
conies of Type II (fig. 64). 



Theorem 20 and its plane dual then give at once : 



THEOREM 23. Any line in the 
plane of a pencil of conies of 
Type II (which does not pass 
through a vertex of the determin- 
ing triangle) is met by the conies 
of the pencil in the pairs of an 
involution* 

COROLLARY. Through a general 
point in the plane there is one and 
only one conic of the pencil; and 
tangent to a general line in the 
plane there are two or no conies 
of the pencil. 



THEOREM 23'. The tangents 
through any point in the plane 
of a range of conies of Type II 
(which is not on a side of the 
determining triangle) to the conies 
of the range are the pairs of an 
involution. 

COROLLARY. Tangent to a gen- 
eral line in the plane there is one 
and only one conic of the range; 
and through a general point in 
the plane there are two or no 
conies of the range. 



DEFINITION. The set of all conies through two given points and 
tangent to two given lines through these points respectively is called 

a pencil or range of conies of Type 
IV* (fig. 65). 

Theorem 21 now gives at once: 

THEOREM 24. Any line in the plane 
of a pencil of conies of Type IV (which 
does not pass through either of the 
points common to all the conies of 
the pencil) is met ly the conies of the 
pencil in the pairs of an involution. 
Through any point in the plane (not 
on either of the lines that are tangent 
to all the conies of the pencil) the 
tangents to the conies of the pencil are the pairs of an involution. Tlie 
line joining the two points common to all the conies of the pencil meets 

* The classification of pencils and ranges of conies into types corresponds to the 
classification of the corresponding plane collineations (cf. Exs. 2, 4, 7, below). 




FIG. Co 



47] PENCILS AND RANGES 131 

any line in a double point of the involution determined on that line. 
And {lie point of intersection of the common tangents is joined to any 
point ly a double line of the involution determined at that point 

COROLLAEY. Throitgli any general point or tanyent to any general 
line in the plane there is one and only one conic of the pencil. 

EXERCISES 

1. What are the degenerate conies of a pencil or range of Type 7? The 
diagonal triangle of the fundamental quadrangle (quadrilateral) of the pencil 
(range) is the only triangle which is self -polar with respect to two conies of 
the pencil (range). 

2. Let A' 2 and B 2 be any two conies of a pencil of Type /, and let P be any 
point in the plane of the pencil. If p is the polar of P with respect to .-I 2 , and 
P' is the pole of p with respect to B' 2 , the correspondence thus established 
between [P] and [P'] is a projective collineation of Type J, whose invariant 
triangle is the diagonal triangle of the fundamental quadrangle. The set of 
all projective collineations thus determined by a pencil of conies of Type I 
form a group. Dualize. 

3. What are the degenerate conies of a pencil or range of Type //? 

4. Let a pencil of conies of Type // be determined by a triangle ABC and 
a tangent a through A. Further, let ' be the harmonic conjugate of a with 
respect to AB and AC, and let A' be the intersection of a and BC. Then 
A, a and A', a' are pole and polar with respect to every conic of the pencil ; and 
no pair of conies of the pencil have the same polars with regard to any other 
points than A and A'. Dualize, and show that all the collineations determined 
as in Ex. 2 are in this case of Type //. 

5. What are the degenerate conies of a pencil or range of Type JT? 

6. Show that any point on the line joining the two points common to all 
the conies of a pencil of Type IV has the same polar with respect to all the 
conies of the pencil, and that these all pass through the point of intersection 
of the two common, tangents. 

7. Show that the collineations determined by a pencil of Type IV by the 
method of Ex. 2 are all homologies (i.e. of Type IV). 

* The pencils and ranges of conies thus far considered have in com- 
mon the properties (1) that the pencil (range) is completely defined 
as soon as two conies of the pencil (range) are given ; (2) the conies 
of the pencil (range) determine an involution on any line (point) in 
the plane (with the exception of the lines (points) on the determining 
points (lines) of the pencil (range)). Three other systems of conies may 
be defined which likewise have these properties. These new systems 

* The remainder of this section may be omitted on a first reading. 



132 



COtflC SECTIONS 



[CHAP. V 



may be regarded as degenerate cases of the pencils and ranges already 
defined. Their existence is established by the theorems given below, 
which, together with their corollaries, may be regarded as degenerate 
cases of the theorem of Desargues. "We shall need the following 

LEMMA. Any conic is transformed lij a protective collineation in 
the plane of the conic into a conic sucli that the tangents at homologous 
points are homologous. 

Proof. This follows almost directly from the definition of a conic. 
Two projective flat pencils are transformed by a projective collineation 
into two projective flat pencils. The intersections of pairs of homologous 
lines of one pencil are therefore transformed into the intersections 
of the corresponding pairs of homologous lines of the transformed 
pencils. If any line meets the first conic in a point P, the transformed 
line will meet the transformed conic in the point homologous with P. 
Therefore a tangent at a point of the first conic must be transformed 
into the tangent at the corresponding point of the second conic. 

THEOREM 25. If a line p Q is a tangent to a conic A 2 at a 'point P OJ 
and Q is any point of A 2 , then through any point on the plane of A 2 

but not on A 2 or p , 
there is one and only 
one conic B* through 
PS and Q, tangent to 
Pv and such that there 
is no point except P^ 
which has the same 
polar with regard to 
loth A* and B\ 

Proof. If P J is any point of the plane not on p Q or 'A 2 , let P be 
the second point in which P Q P ! meets A* (fig. 66). There is one and 
only one elation with center P Q and axis P Q Q changing P into P ; 
(Theorem 9, Chap. III). This elation (by the lemma above) changes 
A* into another conic B* through the points P^ and Q and tangent 
to jp . The lines through J^ are unchanged by the elation, whereas 
their poles (on p ) are subjected to a parabolic projectivity. Hence 
no point on p Q (distinct from P Q ) has the same polar with regard to A* 
as with regard to B*. Since A 2 is transformed into B* by an elation, 
the two conies can have no other points in common than P Q and Q. 




47] 



PEXCILS AXB EAXGES 



133 



That there is only one eonic B' 2 through P r satisfying the con- 
ditions of the theorem is to be seen as follows : Let QP meet p in 3, 
and QP 1 meet p in S r (fig. 66;. The points S and S r must have the 
same polar with regard to A 2 and any conic J5 2 , since this polar 
must be the harmonic conjugate of p^ with regard to P Q Q and P Q P. 
Let p be the tangent to A 2 at P and p f be the tangent to If 2 at P f , 
and let p and p r meet p Q in I and 2 17 respectively. The points 




FIG. 67 



T and T J have the same polar, namely P Q P, with regard to A 2 and 
any conic B*. By the conditions of the theorem the projectivity 



must be parabolic. Hence, by Theorem 23, Cor.,, Chap. IV, 

Q(P ST,P Q T f S r ). 

Hence p and p f must meet on P^Q in a point B so as to form the quad- 
rangle EQPP 1 . This determines the elements P Q , Q, P f > p Q , p f of B*> 
and hence there is only one possible conic J5 2 . 

COROLLARY 1. The conies A 2 and 3 s can have no other points in 
common than P and Q. 

COROLLARY 2. Any line I not on P^ or Q which meets A 2 and B* 
meets them in pairs of an involution in which the points of intersection 
of I with P Q Q <md p Q are conjugate. 

Proof. Let I meet Jf in N and N v J5 2 in L and L v J%Q in M, and 
j? in M 1 (fig. 67). Let K and ^ be the points of A 2 which are trans- 
formed by the elation into L and L^ respectively. By the definition of 
an elation K and S t are colHnear with Jf, while K is on the line JJ 
and J^ on LJ%. Let KN^ meet p in jK, and j^JJ meet JCS^ in & 



134 



COXIC SECTIONS 



[CHAP. V 



Then, since JN", E, X v K are on the conic to which. p Q is tangent at ^, we 
have, by Theorem 6, applied to the degenerate hexagon P^K^EN^N, 
that S 9 L v and E are collinear. Hence the complete quadrilateral 
Sit, JEYy JOTp I has pairs of opposite vertices on P Q 3I and P Q M V P Q N 
and J^Y r JJ and P L r Hence 



ox . The set of all conies 
through a point Q and tangent to 
a line p^ at a point -Z^ and such 
that no point of p except % has 
the same polar with regard to two 
conies of the set, is called a pencil 
of conies of Type III (tig. 68). 



DEFINITION. The set of all conies 
tangent to a line q and tangent to 
a line p Q at a point P Q , and such 
that no line on P except p has 
the same pole with regard to two 
conies of the set, is called a range 
of conies of Type III (fig. 69). 




FIG. 08 



FIG. 09 



Two conies of such a pencil (range) are said to have contact of the 
second order, or to osculate, at JJ. 

Corollary 2 of Theorem 25 now gives at once: 



THEOREM 26. Any line in the 
plane of n peticil of conies of 
Type III, which is not on either of 
the common points of the pencil, is 
met hy the conies of the pencil in the 
pa i/'s of a n iii i*oht t ion. TJi ro ugh 
any point iti the plane except the 
ctnnitwn points there is one and 
only one conic of th e pencil; and 
tanyent to any line not through 
either of the common points there 
are ttro or no w flics of the pencil. 



THEOREM 26'. Tlirougli any point 
in tlie plane of a range of conies of 
Type III, which is not on either of 
the common tangents of the range, 
the tangents to the conies of the pen- 
cil are the pairs of an involution. 
Tangent to any line in the plane ex- 
cept the common tangents there is 
one and only one conic of the range; 
and through any point not on either 
of the common tangents there are 
two or no conies of the range. 



* This argument has implicitly proved that three pairs of points of a conic, as 
Nii P Q> such that the lines joining them meet in a point M, are projected 
from any point of the conic by a quadrangular set of lines (Theorem 16, Chap. VUI). 



47] PENCILS AND BAXGES 135 

The pencil is determined by The range is determined by 

the two common points, the com- the two common tangents, the 

mon tangent, and one conic of the common point, and one conic of 

pencil. the range. 

EXERCISES 

1. What are the degenerate conies of this pencil and range? 

2. Show that the collineation obtained by making correspond to any point P 
the point P' which has the same polar p with regard to one given conic of the 
pencil (range) that P has with regard to another given conic of the pencil (range) 
is of Type 1/7. 

THEOREM 27. If a line p^ is tangent to a conic A 2 at a point P Qt 
there is one and only one conic tangent to p$ at P^ and passing 
through any other point P 1 of the plane of A 2 not on p Q or A 2 
which determines for every point of p Q the same polar line as does A~. 

Proof. Let P be the second point in which P^P* meets A 2 (fig. 70). 
There is one and only one elation of which 2J is center and p Q axis, 
changing P to P 1 . This elation changes A 2 into a conic B* through 

B* 




P r , and is such that if q is any tangent to A 2 at a point Q, then q is 
transformed to a tangent $ f of B* passing through gjp , and Q is trans- 
formed into the point of contact Q r of q r , collinear with Q and JJ. 
Hence there is one conic of the required type through P f . 

That there is only one is evident, because if I is any line through JP', 
any conic B* must pass through the fourth harmonic of P J with regard 
to lp Q and the polar of lp a as to A* (Theorem 13). By considering two 
lines I we thus determine enough points to fix B\ 

COBOLLAEY 1. By duality there is one and only one conic B z tangent 
to any line not passing through P y 



136 CONIC SECTIONS [CHAF.V 

COROLLARY 2. Any line I not on P^ which meets A* and B* meets 
them in pairs of an involution one doitble point of which is lp Q) and 
the other the point of I conjugate to lp Q with respect to A 2 . A dual 
statement holds for any point L not on p Q . 

COROLLARY 3. The conies A 2 and B' 2 can have no other point in 
commoti thati J?J cnid no other tangent iti common than p Q . 

Proof. If they had one other point P in common, they would have 
in common the conjugate of P in the involution determined on any 
line through P according to Corollary 2. 

DEFINITION. The set of all conies tangent to a given line p Q at a 
given point P , and such that each point on p has the same polar 
with regard to all conies of the set, is called a pencil or range of 
conies of Type V. Two conies of such a pencil are said to have 
contact of the third order, or to hyperosculate at JJ. 

Theorem 27 and its first two corollaries now give at once: 

THEOREM 28. Any line I not on the common point of a pencil of 
Type V is met ly the conies of the pencil in pairs of an involution 
one double point of which is the intersection of I with the common 
tangent. Through any point L not on the common tangent the pairs 
of tangents to the conies of the pencil form an involution one double 
line of which is the line joining L to the common point. There is one 
conic of the set through each point of the plane not on the common 
tangent, and one conic tangent to each line not on the common point. 

The pencil or range is determined by the common point, the common 
tangent, and one conic of the set 

EXERCISES 

1. What are the degenerate conies of a pencil of Type F? 

2. Show that the collineation obtained by making correspond to any 
point P the point Q which has the same pole p with regard to one conic of 
a pencil of Type V that P has with regard to another conic of the pencil is 
an elation. 

3. The lines polar to a point A with regard to all the conies of a pencil 
of any of the five types pass through a point A'. The points A and A' are 
double points of the involution determined by the pencil on the line A A'. 
Construct A'. Dualize. Derive a theorem on the complete quadrangle as a 
special case of this one. 

4. Construct the polar line of a point A with regard to a conic C 2 being 
given four points of C* and a conjugate of A with regard to <7 2 . 



47] PENCILS AXD BADGES 137 

5. Given an involution I on a line Z, a pair of points A and A' on I not 
conjugate in I, and any other point B on L construct a point B' such that .4 
and A' and B and 5' are pairs of an involution I' whose double points are a 
pair in I. The involution I' may also be described as one which is commu- 
tative with I, or such that the product of I and T is an involution. 

6. There is one and only one conic through three points and having a 
given point P and line p as pole and polar. 

7. The conies through three points and having a given pair of points as 
conjugate points form a pencil of conies. 

MISCELLANEOUS EXERCISES 

1. If and o are pole and polar with regard to a conic, and *1 and B are 

two points of the conic collinear with 6>, then the conic is generated by the 

two pencils A [P] and B [P'] where P and P' are paired in the involution 

on o of conjugates with regard to the conic. 

/ 2. Given a complete five-point ABCDE. The locus of all points X such 

that X (BCDE) T A (BCDE) 

is a conic. A 

V 3. Given two projective nonperspective pencils, [p] and [<?]. Every line I 



upon which the projectivity Z[^] -r-l^q] is involutorie passes through a fixed 
point 0. The point is the pole of the line joining the centers of the pencils 
with respect to the conic generated by them. 

^ 4. If two complete quadrangles have the same diagonal points, their eight 
vertices lie on a conic (Cremona, Projective Geometry (Oxford, 1885), Chap.XX). 

5. If two conies intersect in four points, the eight tangents to them at 
these points are on the same line conic. Dualize and extend to the cases 
where the conies are in pencils of Types II-V. 

6. All conies with respect to which a given triangle is self -conjugate, and 
which pass through a fixed point, also pass through three other fixed points. 
Dualize. 

7. Construct a conic through two given points and with a given self- 
conjugate triangle. Dualize. 

8. If the sides of a triangle are tangent to a conic, the lines joining two 
of its vertices to any point conjugate with regard to the conic to the third 
vertex are conjugate with regard to the conic. Dualize. 

9. If two points P and Q on a conic are joined to two Conjugate points P f 9 Q? 
on a line conjugate to PQ, then PP' and QQf meet on the conic. 

10. If a simple quadrilateral is circumscribed to a conic, and if I is any 
transversal through the intersection of its diagonals, / will meet the conic and 
the pairs of opposite sides in conjugate pairs of an involution. Dualize. 

11. Given a conic and three fixed collinear points A, B, C. There is a fourth 
point D on the line AB such that if three sides of a simple quadrangle in- 
scribed in the conic pass through A, B, and C respectively, the fourth passes, 
through D (Cremona, Chap. XVII). 



138 CONIC SECTIONS [CHAP.V 

12. If a variable simple n-line ( even) is inscribed in a conic in such a way 
that n 1 of its sides pass through n 1 fixed collinear points, then the other 
side passes through another fixed point of the same line. Dualize this theorem. 

13. If two conies intersect in two points A. B (or are tangent at a point A) 
and two lines through A and B respectively (or through the point of contact 
.4) meet the conies again in 0, (/ and L, Z/, then the lines OL and O'L' meet 
on the line joining the remaining points of intersection (if existent) of the 
two conies. 

14. If a conic C 2 passes through the vertices of a triangle which is self- 
polar with respect to another conic K 2 , there is a triangle inscribed in C 2 and 
self -polar with regard to J 2 , and having one vertex at any point of C 2 . The 
lines which cut C 3 and K* in two pairs of points which are harmonically con- 
jugate to one another constitute a line conic C|, which is the polar reciprocal 
of C 2 with regard to K* (Cremona, Chap. XXII). 

15- If a variable triangle is such that two of its sides pass respectively 
through two fixed points 0' and lying on a given conic, and the vertices oppo- 
site them lie respectively on two fixed lines u and u', while the third vertex 
lies always on the given conic, then the third side touches a fixed conic, which 
touches the lines u and u'. Dualize (Cremona, Chap. XXII). 

16. If P is a variable point on a conic containing A, B, C, and Ms a vari- 
able line through P such that all throws T (PA, PB: PC, /) are projective, 
then all lines I meet in a point of the conic (Schroter, Journal fur die reine und 
angewandte Mathematik, Tol. LXII, p. 222). 

17. Given a fixed conic and a fixed line, and three fixed points A, B, C on 
the conic, let P be a variable point on the conic and let PA, PB, PC meet 
the fixed line in A', B'* C'. If is a fixed point of the plane and (OA', PB'} = K 
and (j&TC") = /, then K describes a conic and I a pencil of lines whose center is 
on the conic described by K (Schroter, loc. cit.). 

18. Two triangles ABC and PQR are perspective in four ways. Show that 
if ABC and the point P are fixed and Q, R are variable, the locus of each of 
the latter points is a conic (cf. Ex. 8, p. 105, and Schroter, Mathematische 
Annalen, Vol. II (1870), p. 553). 

19. Given six points on a conic. By taking these in all possible orders 
60 different simple hexagons inscribed in. the conic are obtained. Each of 
these simple hexagons gives rise to a Pascal line. The figure thus associated 
with any six points of a conic is called the hexagramma mysticum.* Prove the 
following properties of the hexagramma niysticum : 

i. The Pascal lines of the three hexagons PiPoPgP^Pg, P^P^P Z P^P 5 P^ 
and PiPgPsP^PgP^ are concurrent. The point thus associated with such a set 
of three hexagons is called a Steiner point. 
ii. There are in all 20 Steiner points. 

* On the Pascal hexagram cf . Steiner-Schroter, Vorlesungen tiber Synthetische 
Geometrie, Vol. II, 28 ; Salmon, Conic Sections in the Notes ; Christine Ladd, 
American Journal of Mathematics, Vol. IE (1879), p. 1. 



47] EXERCISES 139 

iii. From a given simple hexagon five others are obtained by permuting 
in all possible ways a set of three vertices no two of which are adjacent. The 
Pascal lines of these six hexagons pass through two Steiner points, which are 
called conjugate Steiner points. The 20 Steiner points fall into ten pairs of 
conjugates. 

iv. The 20 Steiner points lie by fours on 15 lines called Steiner lines. 
v. AVhat is the symbol of the configuration composed of the 20 Steiner 
points and the 15 Steiner lines ? 

20. Discuss the problem corresponding to that of Ex. 19 for all the special 
cases of Pascal's theorem. 

21. State the duals of the last two exercises. 

22. If in a plane there are given two conies, any point A has a polar with 
respect to each of them. If these polars intersect in A', the points A , A' are 
conjugate with respect to both conies. The polars of A / likewise meet in A. 
In this way every point in the plane is paired with a uniqxte other point. By 
the dual process every line in the plane is paired with a unique line to which 
it is conjugate with respect to both conies. Show that in this correspondence 
the points of a line correspond in general to the points of a conic. All such 
conies which correspond to lines of the plane have in common a set of at most 
three points. The polars of every such common point coincide, so that to each 
of them is made to correspond all the points of a line. They form the excep- 
tional elements of the correspondence. Dualize (Reye-Holgate, p. 110).* 

23. If in the last exercise the two given conies pass through the vertices of 
the same quadrangle, the diagonal points of this quadrangle are the " common 
points " mentioned in the preceding exercise (Reye-Holgate, p. 110). 

24. Given a cone of lines with vertex and a line u through 0. Then a 
one-to-one correspondence may be established among the lines through by 
associating with every such line a its conjugate a' with respect to the cone 
lying in the plane au. If, then, a describes a plane ?r, of will describe a cone of 
lines passing through u and through the polar line of ?r, and which has in 
common with the given cone any lines common to it and to the given cone 
and the polar plane of u (Reye-Holgate, p. 111).* 

25. Two conies are determined by the two sets of five points A, B, C,D,E 
and A,B,C 9 H, K. Construct the fourth point of intersection of the two conies 
(Castelnuovo, Lezioni di Geometria, p. 391). 

26. Apply the result of the preceding Exercise to construct the point P such 
that the set of lines P(A, JB, C, D, E) joining P to the vertices of any given 
complete plane five-point be projective with any given set of five points on a 
line (Castelnuovo, loc. cit.). 

27. Given any plane quadrilateral, construct a line which meets the sides 
of the quadrilateral in a set of four points projective with any given set of 
four collinear points. 

* The correspondences defined in Exs. 22 and 24 are examples of so-called 
quadratic correspondences. 



140 COKIC SECTIONS [CHAP.V 

28. Two sets of five points .1, B, C, Z>, E and .4, B, H, K 9 L determine 
two conies which, intersect again in two points A", I". Construct the line XT 
and show that the points JY", F are the double points of a certain involution 
(Castelnuovo, loc. cit.). 

29. If three conies pass through two given points A, B and the three pairs 
of conies cut again in three pairs of points, show that the three lines joining 
these pairs of points are concurrent (Castelnuovo, loc. cit.). 

30. Prove the converse of the second theorem of Desargues : The conies 
passing through three fixed points and meeting a given line in the pairs of 
an involution pass through a fourth fixed point. This theorem may be used 
to construct a conic, given three of its points and a pair of points conjugate 
with respect to the conic. Dualize (Castelnuovo, loc. cit.). 

31. The poles of a line with respect to all the conies of a pencil of conies 
of Type 1 are on a conic which passes through the diagonal points of the 
quadrangle defining the pencil. This conic cuts the given line in the points 
in which the latter is tangent to conies of the pencil. Dualize. 

32. Let p be the polar of a point P with regard to a triangle ABC. If P 
varies on a conic which passes through ^i , B, C* then p passes through a fixed 
point Q (Cayley, Collected Works, Vol. I, p. 301). 

33. If two conies are inscribed in a triangle, the six points of contact are 
on a third conic. 



CHAPTER VI 

ALGEBRA OF POINTS AND OUTE -DIMENSIONAL COORDINATE 

SYSTEMS 

48, Addition of points. That analytic methods may be introduced 
into geometry on a strictly protective basis was first shown by von 
Staudt.* The point algebra on a line which is defined in this chapter 
without the use of any further assumptions than A, E, P is essentially 
equivalent to von Staudt's algebra of throws (p. 60), a brief account 
of which will be found in 55. The original method of von Staudt 
has, however, been considerably clarified and simplified by modern 
researches on the foundations of geometry.f All the definitions and 
theorems of this chapter before Theorem 6 are independent of As- 
sumption P. Indeed, if desired, this part of the chapter may be read 
before taking up Chap. IV. 

Given a line I, and on I three distinct (arbitrary) fixed points which 
for convenience and suggestiveness we denote by P , P^ P n , we define 
two one-valued operations $ on pairs of points of I with reference to 
the fundamental points P Q) JJ, P^. The fundamental points are said 
to determine a scale on I. 

DEFINITION. In any plane through I let L and IL be any two lines 
through P^, and let Z be any line through 1% meeting / and II in 
points A and A f respectively (fig. 71). Let P x and P y be any two points 
of I, and let the lines P X A and P y A* meet II and Z in the points X 
and Y respectively. The point P x+y , in which the line XT meets Z, is 
called the sum of the points P x and P v (in symbols J^ + ^ = -^ +y ) in 



* K G. C. von Staudt (1798-1867), BeitrSge zur Geometric der Lage, Heft 2 (1857), 
pp. 166 et seq. This "book is concerned also with the related question of the inter- 
pretation of imaginary elements in geometry. 

t Of., for example, G. Hessenberg, Ueber einen Geometrischen Calcul, Acta 
Mathematiea, Vol. XXIX, p. 1. 

J By a one-valued operation o on a pair of points A, B is meant any process 
whereby with every pair A^ B is associated a point O, which is unique provided 
the order of A, B is given; in symbols AoBC. Here "order" has no geo- 
metrical significance, but implies merely the formal difference of AoB and Bo A. 
If AoB BoA, the operation is commutative; if (AoB)oC=Ao(BoC), the opera- 
tion is associative. 

141 



142 



ALGEBEA OF POINTS 



[CHAP. VI 



the scale ^, P v P n . The operation of obtaining the sum of two points 
is called addition* 




THEOREM 1. If P x and P y are distinct from P Q and P ee) Q(P a ,P x P Q9 
y ) is a necessary and sufficient condition for the equality 



This follows immediately from the definition, AXA!Y being a 
quadrangle which determines the given quadrangular set. 

COROLLARY 1. If P x is any point of I, we have P x +P Q = P + P x =iI x> 

This is also an immediate consequence of the definition. 

COROLLARY 2. The operation of addition is one-valued for every 
pair of points P& P y of I, except for the pair P mi Jf^. (A, E) 

This follows from the theo- 
rem above and the corollary of 

* The historical origin of this con- 
struction will he evident on inspection 
of the attached figure. This is the 
figure which results, if we choose for 
& the "line at infinity" in the plane 
in the sense of ordinary Euclidean 
geometry (cf. p. 8). The construction 
is clearly equivalent to a translation 
of the vector P Q P y along the line Z, 
which brings its initial point into coincidence with the terminal point of the vector 
PoP x , which is the ordinary construction for the sum of two vectors on a line. 




48] ADDITION 143 

Theorem 3, Chap. II, in case P x and P y are distinct from JfJ and P um 
If one of the points P x , P y coincides with P Q or P 9 , it follows from 
Corollary 1. 

COROLLARY 3. The operation of addition is associative; i.e. 



for any three points P^ P v , P s for which the above expressions are 
defined. (A, E) 

Proof (fig. 73). Let J^-f J^ be determined as in the definition by- 
means of three lines ?, 7, 7 and the line XF. Let the line ^Y be 
denoted by Z ', and by means of l*JL 9 7 ' construct the point (J + P y ) 4- ^, 




. 73 



which is determined by the line XZ, say. If now the point P y -}~P z 
be constructed by means of the lines l m , l, 7 ; , and then the point 
P x + (P y + %) be constructed by means of the lines l m , l! l^ it will be 
seen that the latter point is determined by the same line XZ. 

COROLLARY 4 The operation of addition is commutative; i.e. 



for every pair of points P x , P y for which the operation is defined. (A, E) 

Proof. By reference to the complete quadrangle AXA ! Y (fig. 71) 
there appears the quadrangular set Q(P m P y P , 1J, jyj +y ), which by the 
theorem implies that P y +I^^P x ^ jr But, by definition, P x +P y = P x + y . 
Hence 



144 ALGEBBA. OF PODTTS [CHAP.TI 

THEOREM 2. Any three points P x , 1^, P a (P a = R) satisfy the relation 
P ppp p p p p 

J -<x> J - J - J -i - L 'o- t - c ' t 



i.e. the correspondence established by making each point P x of I corre- 
spond to P x r = P x +P ai where P a (= P*) is any fixed point of I, is protective. 
(A, E) 

Proof. The definition of addition (fig. 71) gives this projectivity as 
the result of two perspectivities:* 



The set of all projectivities determined by all possible choices of P a in the 
formula Pi-P^P^ is the group described in Example 2, p. 70. The sum of 
two points P a and P b might indeed have been defined as the point into which 
P b is transformed when P is transformed into P a by a projectivity of this 
group. The associative law for addition would thus appear as a special case 
of the associative law which holds for the composition of correspondences in 
general ; and the commutative law for addition would be a consequence of the 
commutativity of this particular group. 




FIG. 74 

49. MtdtipUcation of points. DEFIXITIOX. In any plane through I 
let ? , l v k be any three lines through P Q9 P v P^ respectively, and let ^ 
meet 4" and I* in points A and B respectively (fig. 74). Let P x) P y be any 
two points of 1 9 and let the lines P X A and P y B meet 1 9 and Z in the points 
X and T respectively. The point 2^ in which the line XY meets I is 

* To make fig. 71 correspond to the notation of this theorem, P must be 
identified with P a . 



49] MULTIPLICATION 145 

called the product ofP x by P y (in symbols P x -P y =P^)m the scale P^ P^ P x 
on /. The operation of obtaining the product of two points is called 
multiplication* Each of the points P x , P u is called a factor ' of the 
product P x - P r 

THEOREM 3. If P x and P v are any tv:o i^ints of I distinct from 
P Q , P l} P X) Q(P Q P x P ly P*P y P xlf ) is necessary and sufficient for the equality 
P X -^P^. (A,E) 

This follows at once from the definition, AXJBY being the defining 
quadrangle. 

COROLLARY 1. For any point P, (=F P^) on I ice Jtrn-e the relations 
P^P = P X P, = P X ; P^P^P^P^P,; P X P^P^P^P^ (A,E) 

This follows at once from the definition. 

COROLLARY 2. TJie operation of multiplication is one-valued for 
every pair of points P x , P y of I, except P^ - P^ and P^ P^. (A, E) 

This follows from Corollary 1, if one of the points P 3) P y coincides 
with %, P v or P n . Otherwise, it follows from the corollary, p. 50, in 
connection with the above theorem. 




Po 



* The origin of this construction may also "be seen in a simple construction of 
metric Euclidean geometry, which results from the construction of the definition 
"by letting the line Z he the "line at infinity" (cf. p. 8). In the attached figure 
which gives this metric construction we have readily, from similar triangles, the 
proportions: j^ = y^ = JiP , ^ 

PO JPy PO y Po-Pary 

which, on taking the segment P Pi= 1, gives the desired result PoPy= PoP 



146 



ALGEBBA OF POINTS 



[CHAP. VI 



COBOLLARY 3. The operation of multiplication is associative; i.e. we 
ha re (^ - P y ) P z = P A - (P y P z ) for every three points P x , JJ, P z for which 
these products are defined. (A, E) 

Proof (fig. 76). The proof is entirely analogous to the proof for 
the associative law for addition. Let the point P x P y be constructed 



lo 




FIG. 70 

as in the definition by means of three fundamental lines 1 Q) 1 13 
the point P xy being determined by the line XT. Denote the line . 
by l[, and construct the point P^ P z = (P x P y ) *P S) using the lines Z , l[, / 
as fundamental. Further, let the point P y P s = ^ be constructed by 
means of the lines Z , V v ?, and then let P x *I> yz = P x * ( 1J) be con- 
structed by means of Z , I v I*. It is then seen that the points P x P us 
and P^ JP are determined by the same line. 

By analogy with Theorem 1, Cor. 4, we should now prove that mul- 
tiplication is also commutative. It will, however, appear presently 
that the commutativity of multiplication cannot be proved without 
the use of Assumption P (or its equivalent). It must indeed be clearly 
noted at this point that the definition of multiplication requires the 
first factor P x in a product to form with JjJ and P l a point triple of 
the quadrangular set on I (cf. p. 49); the construction of the product 
is therefore not independent of the order of the factors. Moreover, 
the fact that in Theorem 3, Chap. II, the quadrangles giving the points 
of the set are similarly placed, was essential in the proof of that 



49] MULTIPLICATION 147 

theorem. "We cannot therefore use this theorem to prove the com- 
mutative law for multiplication as in the case of addition. 

An important theorem analogous to Theorem 2 is, however, inde- 
pendent of Assumption P. It is as follows : 

THEOREM 4 If the relation P x -P >f = P xtJ holds Ittvecn any three 
'points P^ P v , P jy on I distinct from f^ we Jut re P^P^P^ ^P^P^ 
and also .SJJJfJJJ P^P^P^; i.e. the correspondence established "by 
makintj each point P x of I correspond to P^P x *P a (or to P x '~P a -Pj}, 
where P a is a ny fixed point of I distitict from zero, is projective. (A, E) 

Proof. The definition of multiplication gives the first of the above 
projectivities as the result of two perspectivities (fig. 76) : 



The second one is obtained similarly. In fig. 76 we have 



The set of all projectivities determined by all choices of P a in the for- 
mula Px = P x -P a is the group described in Example 1, p. 69. The proper- 
ties of multiplication may be regarded as properties of that group in the same 
svay that the properties of addition arise from the group described in Example 
2, p. 70. In particular, this furnishes a second proof of the associative law 
for multiplication. 

THEOREM 5. Multiplication is distributive with respect to addition; 
i.e. if P x) P y , P z are any three points on I (for which the operations 
"below are defined), we have 

P z .(P x +P y )=P s .P I +P,.P y , a nd(Z+P l ).P z =P I .P s + P v .P,. (A,E) 
Proof. Place 

P x +P y = P x + y , P Z -P^P^ P s -P y = P, y , and P z -P^ = P^+ yy 
By Theorem 4 we then have 

5 -? ^l-VV^c -i- y "A * ^0 -? *zx *zy2z(x + y)' 

But by Theorem 1 we also have (&%%> %>%%+) Hence, by 
Theorem 1, Cor., Chap. IV, we have Q^J^JJ, P^P^^ which, 
by Theorem 1, implies -^ B -r-^=-^ (a:+ y ) . The relation 



is proved similarly. 



148 ALQEBEA OF POINTS [CHAP.VI 

50. Tlie commutative law for multiplication. With the aid of 
Assumption P we will now derive finally the commutative law for 
multiplication : 

THEOREM 6. TJie operation of multiplication is commutative; i.e. 
we hare P x JJ = P y - P for every pair of points P x , P y of I for which 
these two products are defined. (A., E, P) 

Proof. Let us place as "before P x - P y = P xy , and J - P x = P yc . Then, by 
the first relation of Theorem 4, and interchanging the points P^ P^ 
we have 



and from the second relation of the same theorem we have 



By Theorem 17, Chap. IV, this requires P^ P^. 

In view of the fact already noted, that the fundamental theorem 
of protective geometry (Theorem 17, Chap. IT) is equivalent to 
Assumption P, the proof just given shows further: 

THEOKEM 7. Assumption P is necessary and sufficient for the com- 
mutative law for multiplication* (A, E) 

51. The inverse operations. DEFINITION. Given two points P a , J% 
on Z, the operation determining a point JP satisfying the relation 
P^ + P x =iI is called subtraction; in symbols IP a =zP x . The point 
P x is called the difference of J? from P a . Subtraction is the inverse of 
addition. 

The construction for addition may readily be reversed to give a con- 
struction for subtraction. The preceding theorems on addition then give : 

THEOEEM 8. Subtraction is a one-valued operation for every pair 
of points P a > P b on I, except the pair P^ j. (A, E) 

COROLLAEY. We have in particular P a P a P Q for every point 
P a (^PJonl (A, E) 

* The existence of algebras in which multiplication is not commutative is then 
sufficient to establish the fact that Assumption P is independent of the previous 
Assumptions A and E. For in order to construct a system (cf. p. 6) which satisfies 
Assumptions A and E without satisfying Assumption P, we need only construct an 
analytic geometry of three dimensions (as described in a later chapter) and use as a 
basis a noncommutative number system, e.g. the system of quaternions. That the 
fundamental theorem of projective geometry is equivalent to the commutative 
law for multiplication was first established by Hilbert, who, in his Foundations of 
Geometry, showed that the commutative law is equivalent to the theorem of Pappus 
(Theorem 21, Chap. IV). The latter is easily seen to be equivalent to the funda- 
mental theorem. 



60,31,32] ABSTRACT XU^IBEE SYSTEM 149 

DEFINITION. Given two points P a , P b on / ; the point P x determined by 
the relation J^-^ = ^ is called the quotient of J% by P a (also the rutio 
of P b to P a ) ; in symbols JIJ/JJ = Jf, or J : P a = JJ. The operation deter- 
mining j/2 is called diinsio/i; it is the inverse of multiplication.* 

The construction for multiplication may also be reversed to give a 
construction for division. The preceding theorems on multiplication 
then give readily : 

THEOREM 9. Division is a one-valued operation for every pair of 
points P a , P b on I except the pairs P Q , P Q and .%, -%. (A, E) 

COKOLLAHY. We hare in particular PJ% = P^ ./_ = J?, ^/J = , 
etc., for every poi lit P a o>i I distinct from JjJ and 1^. (A, E) 

Addition, subtraction, multiplication, and division are known as 
the four rational operations. 

52. The abstract concept of a number system. Isomorphism. The 
relation of the foregoing discussion of the algebra of points on a line 
to the foundations of analysis must now -be briefly considered. "With 
the aid of the notion of a group (cf. Chap. Ill, p. 66), the general con- 
cept of a number system is described simply as follows : 

DEFINITION. A set N of elements is said to form a number system, 
provided two distinct operations, which we will denote by and 
respectively, exist and operate on pairs of elements of N under the 
following conditions: 

1. The set N forms a group with respect to e. 

2. The set N forms a group with respect to o, except that if i + is 
the identity element of N with respect to e, no inverse with respect 
to o esists for i + .f If a is any element of N, a i+ = ^0 a = i+. 

3. Any three elements a, &, c of N satisfy the relations a (b c) 
= (5)(ac) and (l@c)ea=(bQa)e(cd). 

The elements of a number system are called numbers; the two oper- 
ations and are called addition and mult implication respectively. 
If a number system forms commutative groups with respect to both 
addition and multiplication, the numbers are said to form 



* What we have defined is more precisely right-handed division. The left-handed 
quotient is defined similarly as the point P x determined by the relation P x P a P&, 
In a commutative algebra they are of course equivalent. 

t The identity element i+ in a number system is usually denoted by (zero). 

$ The class of all ordinary rational numbers forms a field; also the class of real 
numbers; and the class of all integers reduced modulo p (p a prime), etc. 



150 ALGEBRA OF POINTS [CHAP.YI 

On the basis of this definition may be developed all the theory 
relating to the rational operations i.e. addition, multiplication, sub- 
traction, and division in a number system. The ordinary algebra 
of the rational operations applying to the set of ordinary rational or 
ordinary real or complex numbers is a special case of such a theory. 
TJie whole terminology of tins algebra, in so far as it is definable in 
terms of the four rational operations, will in the future be assumed 
as defined. We shall not, therefore, stop to define such terms as 
reciprocal of a number, exponent, equation, satisfy, solution, root, etc. 
The element of a number system represented by a letter as a will be 
spoken of as the value of a. A letter which represents any one of a 
set of numbers is called a variable; variables will usually be denoted 
by the last letters of the alphabet. 

Before applying the general definition above to our algebra of 
points on a line, it is desirable to introduce the notion of the 
abstract equivalence or isomorphism between two number systems. 
ox^ If two number systems are such that a one-to-one 



reciprocal correspondence exists between the numbers of the two 
systems, such that to the sum of any two numbers of one system 
there corresponds the sum of the two corresponding numbers of the 
other system; and that to the product of any two numbers of one 
there corresponds the product of the corresponding numbers of the 
other, the two systems are said to be abstractly equivalent or (simply) 



When two number systems are isomorphic, if any series of oper- 
ations is performed on numbers of one system and the same series 
of operations is performed on the corresponding numbers of the 
other, the resulting numbers will correspond. 

53. Noiihomogeneous coordinates. By comparing the corollaries 
of Theorem 1 with the definition of group (p. 66), it is at once 
seen that the set of points of a line on which a scale has been estab- 
lished, forms a group with respect to addition, provided the point jg 
be excluded from the set. In this group J?J is the identity element, 
and the existence of an inverse for every element follows from 
Theorem 8. In the same way it is seen that the set of points on 
a line on which a scale has been established, and from which the 

* For the general idea of the isomorphism between groups, see Burnside's Theory 
of Groups, p. 22. 



53] COOEDIXATES 151 

point .g has been excluded, forms a group with respect to multipli- 
cation, except that no inverse with respect to multiplication exists 
for j?J; P^ is the identity element in this group, and Theorem 9 insures 
the existence of an inverse for every point except J. These con- 
siderations show that the first two conditions in the definition of a 
number system are satisfied by the points of a line, if the operations 9 
and o are identified with addition and multiplication as defined in. 
48 and 49. The third condition in the definition of a number 
system is also satisfied in view of Theorem 5. Finally, in view of 
Theorem 1, Cor. 4, and Theorem 6, this number system of points on 
a line is commutative with respect to both addition and multipli- 
cation. This gives then: 

THEOREM 10. TJie set of all points on o, line on which a scale has 
Tjeeii established, and from -ich ieli the point J2 is excluded, forms a field 
with res'pect to the operations of addition and multiplication preciously 
defined. (A, E, P) 

This provides a new way of regarding a point, viz.,, that of regarding 
a point as a nwnber of a number system. This conception of a point 
will apply to any point of a line except the one chosen as J. It is 
desirable, however, both on account of the presence of such an excep- 
tional point and also for other reasons, to keep the notion of point 
distinct from the notion of number, at least nominally. This we do 
by introducing a field of numbers a, I, c,- ,l,k,- , %,?/,%, which 
is isomorphie witli the field of points on a line. The numbers of the 
number field may, as we have seen, be the points of the line, or they 
may be mere symbols which combine according to the conditions 
specified in the definition of a number system ; or they may be ele- 
ments defined in some way in terms of points, lines, etc.* 

In any number system the identity element with respect to addi- 
tion is called zero and denoted by 0, and the identity element with 
respect to multiplication is called one or unity, and is denoted by 1. 
We shall, moreover, denote the numbers 1+1, 1+1+1, - - -, a, - 
by the usual symbols 2, 3, - - -, a, -.f In the isomorphism of our 
system of numbers with the set of points on a line, the point JJ must 
correspond to 0, the point JJ to the number 1 ; and, in general, to every 

* See, for example, 55, on von Staudt's algebra of throws, where the numbers 
are thought of as sets of four points. 

t Cf., however, in this connection 57 below. 



152 ALQEBEA OF PODTTS [CHAP. VI 

point will correspond a number (except to j), and to every number 
of the field will correspond a point. In this way every point of the 
line (except .) is labeled by a number. This number is called the 
(nonhomogeneous) coordinate of the point, to which it corresponds. 
This enables us to express relations between points by means of 
equations between their coordinates. The coordinates of points, or 
the points themselves when we think of them as numbers of a 
number system, we will denote by the small letters of the alphabet 
(or by numerals), and we shall frequently use the phrase "the point a?" 
in place of the longer phrase " the point whose coordinate is x" It 
should be noted that this representation of the points of a line by 
numbers of a number system -is not in any way dependent on the 
commutativity of multiplication; i.e. it holds in the general geom- 
etries for which Assumption P is not assumed. 

Before leaving the present discussion it seems desirable to point 
out that the algebra of points on a line is merely representative, 
under the principle of duality, of the algebra of the elements of any 
one-dimensional primitive form. Thus three lines Z , l v ! of a flat 
pencil determine a scale in the pencil of lines; and three planes 
# , a v a:,, of an axial pencil determine a scale in this pencil of planes ; 
to each corresponds the same algebra. 

54. The analytic expression for a projectivity in a one-dimensional 
primitive form. Let a scale be established on a line I by choosing 
three arbitrary points for P Q9 J%, , ; and let the resulting field of points 
on a line be made isomorphic with a field of numbers 0, 1, a, - , so 
that JJ corresponds to 0, ^ to 1, and, in general, P a to a. For the 
exceptional point -g, let us introduce a special symbol co with excep- 
tional properties, which will be assigned to it as the need arises. 
It should be noted here, however, that this new symbol oo does not 
represent a number of a field as defined on p. 149. 

We may now derive the analytic relation between the coordinates of 
the points on I, which expresses a projective correspondence between 
these points. Let x be the coordinate of any point of I. We have seen 
that if the point whose coordinate is x is made to correspond to either 
of the points 



(I) 3/=v + a, (a 3= oo) 

or (II) #'=a#, (a = 0) 



54] 



LINEAR FRACTIONAL TRANSFORMATION 



153 



where a is the coordinate of any given point on /, each of the result- 
ing correspondences is projective (Theorem 2 and Theorem 4). It is 
readily seen, moreover, that if x is made to correspond to 

(in) *'=*, 

&,* 

the resulting correspondence is likewise projective. For we clearly 
have the following construction for the point l/,c (fig. 77): With the 
same notation as before for the construction of the product of two 




FIG. 77 



numbers, let the line xA meet 7. in X. If T is determined as the 
intersection of IX with 7 , the line BY determines on Z a point x 1 , 
such that text =1, by definition. "We now have 



The three projectivities (I), (II), and (III) are of fundamental 
importance, as the next theorem will show. It is therefore desirable 
to consider their properties briefly ; we will thus be led to define the 
behavior of the exceptional symbol oo with respect to the operations 
of addition, subtraction, multiplication, and division. 

The projectivity a/= a? + a, from its definition, leaves the point %>, 
which we associated with co, invariant. We therefore place oo 4- a = oo 
for all values of a (a = oo). This projectivity, moreover, can have no 
other invariant point unless it leaves every point invariant ; for the 
equation x = x + a gives at once a = 0, if x 3= oo. Further, by prop- 
erly choosing a, any point x can be made to correspond to any point a/; 



154 ALGEBKA OF POINTS [CHAP. YI 

but when one such pair of homologous points is assigned in addition 
to the double point cc, the projectivity is completely determined. 
The resultant or product of any two projectivities % f =x + a and 
#'= + & is clearly a/= x + (a -f &). Two such projectivities are 
therefore commutative. 

The pro jeetivity d = ax, from its definition, leaves the points and cc 
invariant, and by the fundamental theorem (Theorem 17, Chap. IV) 
cannot leave any other point invariant without reducing to the iden- 
tical projectivity. As another property of the symbol co we have 
therefore cc = a cc (a =r= 0). Here, also, by properly choosing a, any 
point x can be made to correspond to any point a/, biit then the pro- 
jectivity is completely determined. The fundamental theorem in this 
case shows, moreover, that any projectivity with the double points 0, co 
can be represented by this equation. The product of two projectivities 
x r = a:c and xf = 6.r is clearly ic r = (at) x, so that any two projectivities of 
this type are also commutative (Theorem 6). 

Finally, the projectivity x f =l/x, by its definition, makes the 
point co correspond to and the point to co. "We are therefore led 
to assign to the symbol co the following further properties : 1/co = 0, 
and 1/0 = x. This projectivity leaves 1 and 1 (defined as 1) 
invariant. Moreover, it is an involution because the resultant of two 
applications of this projectivity is clearly the identity; ie. if the 
projectivity is denoted by TT, it satisfies the relation 7r 2 = 1. 

THEOREM 11. Any projectivity o/i a line is the product of projec- 
timties of the three types (I), (JJ), and (III), and may be expressed 
in the form 

n\ rf 

(1) ' 

Conversely, every equation of this form represents a projectimty, if 
ad-lc^ 0. (A, E, P) 

Proof. We will prove the latter part of the theorem first. If we 
suppose first that c 3= 0, we may write the equation of the given 

transformation in the form 

, __ ad 

(2) ^.2 + 4. 

c cx + d 

This shows first that the determinant ad "be must be different from 
; otherwise the second term on the right of (2) would vanish, which 



54] LIXEAK FEACTIOXAL TEAXSFORMATIOX 155 

would make every x correspond to the same point tt/c, while a pro- 
jectivity is a one-to-one correspondence. Equation (1), moreover, 
shows at once that the correspondence established by it is the result- 
ant of the five : 

,7 1 /7 ''"^V / " 

x l ex, x n = #! + it, x s = > x = ( o -- .f.j, d x -T --- 

<>. 2 \ c f c 

If c = 0, and ad =?= 0, this argument is readily modified to show that 
the transformation of the theorem is the resultant of projectivities of 
the types (I) and (II). Since the resultant of any series of projectiv- 
ities is a projectivity, this proves the last part of the theorem. 

It remains to show that every projectivity can indeed be repre- 

sented bv an equation jsf = - - - To do this simplv, it is desirable 
- " ex + a 

to determine first what point is made to correspond to the point oc by 
this projectivity. If we follow the course of this point through the 
five projectivities into which we have just resolved this transforma- 
tion, it is seen that the first two leave it invariant, the third trans- 
forms it into 0, the fourth leaves invariant, and the fifth transforms 
it into a/c] the point oo is then transformed by (1) into the point 
a/c. This leads us to attribute a further property to the symbol cc, 

viz., , 7 

a , 

- , when x = cc. 

c 

According to the fundamental theorem (Theorem 17, Chap. IV), a pro- 
jectivity is completely determined when any three pairs of homolo- 
gous points are assigned. Suppose that in a given projectivity the 
points 0, 1, oo are transformed into the points p, q> r respectively. 
Then the transformation 



2) 

clearly transforms into p, 1 into gr, and, by virtue of the relation 
just developed for oo, it also transforms oo into r. It is, moreover, of the 
form of (1). The determinant adlc is in this case (j p)(r q}(r~p)> 
which is clearly different from zero, if p, q, r are all distinct. This 
transformation is therefore the given projectivity. 

COEOLLAEY 1. The projectivity sJ = a/x(a, = 0, or oo) transforms 
"into oo and oo into 0. (A, E, P) 



156 ALGEBKA OF POINTS [CH^.TI 

For it is the resultant of the two projectivities, x l = l/x and 
x f = ax v of which the first interchanges and oo, while the second 
leaves them both invariant. We are therefore led to define the symbols 
a/Q and a/cc as equal to co and respectively, when a is neither 
no r oo. 

COROLLARY 2. Any projectivity leaving the point co invariant may 
1e expressed in the form x f = ax + L (A, E, P) 

COROLLARY 3. A/iy projectivity may le expressed analytically "by 
the lilinear equation c%3/+ dx f ax~ 6 = 0; and conversely, any 
~bilinec.tr equation defines a protective correspondence "between its two 
variables. (A, E, P) 

COROLLARY 4. If a projectixity leaves any points invariant, the 
coordinates of these doulle points satisfy the quadratic equation 
car + (da)x-1> = Q. (A, E, P) 

DEFINITION. A system of mn numbers arranged in a rectangular 
array of w rows and n columns is called a matrix. If m = n, it is 
called a square matrix of order n* 

The coefficients \ , ) of the protective transformation (1) form a 

\c aj * J v ' 

square matrix of the second order, which may be conveniently used to 
denote the transformation. Two matrices ( a , j and ( a f ) repre- 

sent the same transformation, if and only if a : a f == b : &'= c : e r = d : d f . 
The product of two projectivities 

, , . ax + 1 , , , v a f x f + V 
' and z" =*7r.(x f )= . . 7 . 



is given by the equation 

x _ (aa 1 + cV) x + la 1 + db f 
- 



This leads at once to the rule for the multiplication of matrices, 
which is similar to that for determinants. 

DEFINITION. The product of two matrices is defined by the equation 

(a 1 V\ (a l\ _ (aa! + cV la} + db f \ 
\c f d f )\c d) \ac f +cd f I 



* For a development of the principal properties of matrices, cf . B6cher, Intro- 
duction to Higher Algebra, pp. 20 ff. 



54,53] THROWS 157 

This gives, in connection with the result just derived, 
THEOREM 12. The product of two projectmties 



is represented ly the product of their matrices; in sytnloh, 



COROLLAEY 1. The determinant of the 'product of tv:o 2)rojectivitie$ 
is equal to the product of their determinants. (A, E,.P) 

COROLLARY 2. The inverse of the projectimty 7r~i a \ is given 
d ~ l where A, B, C, D are the cofactors 



, 



of a, I, c, d respectively in- the determinant , . (A, E, P) 

This follows at once from Corollary 3 of the last theorem by inter- 
changing x 9 x 1 . We may also verify the relation by forming the 

product flrV= [ T 7 * V which transformation is equiva- 
\ aa oc/ 

lent to (~: !j ]. The latter is called the identical matrix. 

/ 7% \ 

COROLLARY 3. Any involution is represented ly ( a ), that is 

\c -"" a I 

ly x f = 2i , with the condition that a?+t>c^ 0. (A, E, P) 
J ex a 

55. Von Staudt's algebra of throws. "We wDl now consider the 
number system of points on a line from a slightly different point of 
view. On p. 60 we defined a throw as consisting of two ordered 
pairs of points on a line ; and defined two throws as equal when they 
are protective. The class of all throws which are protective (ie. equal) 
to a given throw constitutes a class which we shall call a mark. 
Every throw determines one and only one mark, but each mark 
determines a whole class of throws. 

According to the fundamental theorem (Theorem 17, Chap. IV), if 
three elements A 9 B, C of a throw and their pkces in the symbol 
T(A3, CD) are given, the throw is completely determined by the 
mark to which it belongs. A given mark can be denoted by the 
symbol of any onfe of the (projective) throws which define it. We 
shall also denote marks by the small letters of the alphabet. And so, 
since the equa]ity sign (=) indicates that the two symbols between 



158 ALGEBEA OF POINTS [CHAP.YI 

which it stands denote the same thing, we may write T (AB, CD) = 
a = l t if a, I, T(AB, CD) are notations for the same mark. Thus 
T (AB, CD) = T (BA 9 DC) = T (CD, AB) = T(DC, BA) are all symbols 
denoting the same mark (Theorem 2, Chap. III). 

According to the original definition of a throw the four elements 
which compose it must be distinct. The term is now to be extended 
to include the following sets of two ordered pairs, where A, B, C are 
distinct. The set of all throws of the type 7(AB, CA) is called a 
mark and denoted by <x> ; the set of all throws of the type T(AB, CB) 
is called a mark and is denoted by ; the set of all throws of the type 
T(AB, CC) is a mark and is denoted by 1. It is readily seen that 
if %, %> j are any three points of a line, there exists for every point 
P of the line a unique throw T (_? JJ, JJ, P) of the line ; and con- 
versely, for every mark there is a unique point P. The mark co, by 
what precedes, corresponds to the point R> ; the mark to JJ ; and 
the mark 1 to J?. 

DEFINITION. Let T(AB, CDJ be a throw of the mark a, and let 
T(AB, <7D 2 ) be a throw of the mark &; then, if Z> 3 is determined by 
iB, ADJ>^ 9 the mark c of the throw T(AB 9 CD S ) is called the 
of the marks a and &, and is denoted by a + b; in symbols, 
a -f- J = c. Also, the point D[ determined by Q (AD^C y BDJD^ deter- 
mines a mark with the symbol T(AB, CD f B ) = c r (say), which is called 
the product of the marks a> and 6 ; in symbols, ab = c 7 . As to the 
marks and 1, to which these two definitions do not apply, we define 
further : #-f-0 = + # = , ^-0 = 0-^ = 0, and a - 1 = 1 a = a. 

Since any three distinct points A } B, C may be projected into a fixed 
triple j, ^ JJ, it follows that the operation of adding or multiplying 
marks may be performed on their representative throws of the form 
T(%>PQ> JiP)- By reference to Theorems 1 and 3 it is then clear that 
the class of all marks on a line (except oo) forms a number system, with 
respect to the operations of addition and multiplication just defined, 
which is isomorphic with the number system of points previously 
developed. 

This is, in brief, the method used by von Staudt to introduce ana- 
lytic methods into geometry on a purely geometric basis.* "We have 

* Cf . reference on p. 141. Von Staudt used the notion of an involution on a line 
in defining addition and multiplication ; the definition in terms of quadrangular sets 
is, however, essentially the same as his by virtue of Theorem 27, Chap. IV. 



5o,oG] CROSS RATIO 159 

given it here partly on account of its historical importance; partly 
because it gives a concrete example of a number system isoinurpliic 
with the points of a line * ; and partly because it gives a natural 
introduction to the fundamental concept of the cross ratio uf four 
points. This we proceed to derive in the next section. 

56. Ttie cross ratio. We have seen in the preceding section that 
it is possible to associate a number with every throw of four points 
on a line. By duality all the developments of this section apply also 
to the other one-dimensional primitive forms, i.e. the pencil of lines 
and the pencil of planes. TTith every throw of four elements of any 
one-dimensional primitive form there may be associated a definite 
number, which must be the same for every throw protective with the 
first, and is therefore an invariant under any protective transforma- 
tion, i.e. a property of the throw that is not changed when the throw 
is replaced by any protective throw. This number is called the cross 
ratio of the throw. It is also called the doulle ratio or the anhar- 
monic ratio. The reason for these names will appear presently. 

In general, four given points give rise to six different cross ratios. 
For the 24 possible permutations of the letters in the symbol 
T(AB, CD} fall into sets of four which, by virtue of Theorem 2, 
Chap. Ill, have the same cross ratios. In the array below, the per- 
mutations in any line are protective with each other, two permuta- 
tions of different lines being in general not projective : 



AB, CD 


BA, DC 


DC, BA 


CD, AB 


AB, DC 


BA, CD 


CD, BA 


DC, AB 


AC, BD 


CA, DB 


DB, CA 


BD, AC 


AC, DB 


CA, BD 


BD, CA 


DB, AC 


AD, BC 


DA, CB 


.CB, DA 


BC, AD 


AD, CB 


DA, BC 


BC, DA 


CB, AD 



If, however, the four points form a harmonic set H (AB, CD), the 
throws T(AB, CD) and T(AB, DC) are projective (Theorem 5, 
Cor. 2, Chap. IV). In this case the permutations in the first two rows 
of the array just given are all projective and hence have the same cross 
ratio. The four elements of a harmonic set, therefore, give rise to only 
three crostTratios. The values of these cross ratios are readily seen 

* Cf. 53. Here, with every point of a line on which a scale has been estab- 
lished, is associated a mark which Js the coordinate of the point. 



160 ALQEBEA OF POINTS [CHAP.TI 

to be 1, J-, 2 respectively, for the constructions of our number 
system give* at once H (gJJ, JJP 1 ) 9 H (JJ, 1J1J), and H ( jfj, JJJ>). 

We now proceed to develop an analytic expression for the cross 
ratio "Be, (# r r a , # 3 : 4 ) of any four points on a line (or, in general, of any 
four elements of any one-dimensional primitive form) whose coordi- 
nates in a given scale are given. It seems desirable to precede this 
derivation by an explicit definition of this cross ratio, which is inde- 
pendent of von Staudt's algebra of throws. 

DEFINITION. The cross ratio I&.fax^ x 9 x^ of elements x v # a , # 8 , x 
of any one-dimensional form is, if x 19 # a , x z are distinct, the coordi- 
nate X of the element of the form into which # 4 is transformed by 
the projectivity which transforms x v x s , x s into oo, 0, 1 respectively ; 
i.e. the number, X, defined by the projectivity o^oy^-^coOlX. If 
two of the elements x v # 2 , # 8 coincide and x is distinct from all of 
them, we define E (x^, x 3 x) as that one of B (x 2 x 19 x^), B (# 8 # 4 , 
jTjA'j,), Ii (j: 4 r 8 , ayj, for which the first three elements are distinct. 

THEOREM 13. TJie cross ratio B (x^x^ x z x^ of the four dements 
whose coordinates are respectively x v % 2 , x s> x is given ~by the relation 

\ I "~~" 

- 



(A, E, P) 

Proof, The transformation 



is evidently a projectivity, since it is reducible to the form of a 
linear fractional transformation, viz., 



in which the determinant (a^ c s ) (# 2 ^ S )(cc 3 ^ is not zero, pro- 
vided the points x v % Q> x s are distinct. This projectivity transforms 
x v x 2 , X B into oo, 0, 1 respectively. By definition, therefore, this pro- 
jectivity transforms x into the point whose coordinate is the cross 
ratio in question, Le. into the expression given in the theorem. If 
x v x 2 , X B are not all distinct, replace the symbol 1$ (ayj a , x z x^) by one 
of its equal cross ratios B? (x z x v x^), etc. ; one of these must have 
the first three elements of the symbol distinct, since in a cross ratio 
of four points at least three must be distinct (def.). 



56] OEOSS BATIO 161 

COROLLARY 1. We have in particular 

B Uy>: 2 , ,0,'j) = cc, B (.<.y\, V*.' 2 ) ^ an $ ^ ( i W W = 1 
-z/^, c 2 , # 3 are fl/iy rtra? distitid dements of the form. (A, E) 



COROLLARY 2. T&e ?*0s$ /Y^/O ^/ hanno/iic set H(ay: 2 , a.y 4 ) Z6 5 
B (^yj 2 , iy; 4 ) == 1, for we have H (ac 0, 1 1 ). (A, E s P) 

"COROLLARY 3. J/ R (x*^ ^v^) = X, rt^ other five cross ratios of the 
throws composed of the fovr dements x v # 2 , .r s , A' 4 /? 

5 (' Vs ;<: A) = ' E ( V ! * A v a) ^ ~- ' 



,,, 4fl 
(A, E, P) 

The proof is left as an exercise. 

COROLLARY 4. If x v # 2 , C 3 , x form a harmonic set H (j/^, 



(A, E, P) 

The proof is left as an exercise. 

COROLLARY 5. If a, I, c are any three distinct elements of a one- 
dimensional primitive form, and a 1 , b', d are any three other distinct 
elements of the same form, then the correspondence established ~by the 
relation E (ab 9 ex) = B (aV, c ! x r ) is projective. (A, E, P) 

Proof. Analytically this relation gives 

a c b x__a r c f Vx* 
a x 1) c a f x f U c f 

which, when expanded, evidently leads to a bilinear equation in 
the variables x, x f , which defines a projective correspondence by 
Theorem 11, Cor. 3. 
That the cross ratio 



is invariant under any projective transformation may also be verified directly 
by observing that each of the three types (I), (II), (III) of projectivities on 
pp. 152, 153 leaves it invariant. That every projectivity leaves it invariant 
then follows from Theorem 11. 



162 ALGEBRA OF POINTS [CHAP.YI 

57. Coordinates in a net of rationality on a line. We now con- 
sider the numbers associated with the points of a net of rationality 
on a line. The connection between the developments of this chapter 
and the notion of a linear net of rationality is contained in the 
following theorem : 

THEOREM 14. The coordinates of the points of the net of rationality 
R(JJJ.B)/0m a number system, or field, which consists of all numbers 
each of tchich can be obtained by a finite number of rational algebraic 
operations o/i and 1, and only these. (A, E) 

Proof. By Theorem 14, Chap. IV, the linear net is a line of the 
rational space constituted by the points of a three-dimensional net of 
rationality. By Theorem 20, Chap. IV, this three-dimensional net is 
a properly protective space. Hence, by Theorem 10 of the present 
chapter, the numbers associated with R(Oloo) form a field. 

All numbers obtainable from and 1 by the operations of addi- 
tion, subtraction, multiplication, and division are in R(Olco), because 
(Theorem 9, Chap. IV) whenever x and y are in R (Oloo) the quadran- 
gular sets determining x + y, xy, % y, x/y have five out of six 
elements in R(Oloc). On the other hand, every number of R(Oloo) 
can be obtained by a finite number of these operations. This follows 
from the fact that the harmonic conjugate of any point a in R(Oloo) 
with respect to two others, 5, c, can be obtained by a finite number 
of rational operations on a, I, c. This fact is a consequence of Theo- 
rem 13, Cor. 2, which shows that x is connected with a, b, c by the 

relation 

(x b) (a c) -f (x c) (a b) = 0. 

Solving this equation for x, we have 

2 be ab ac 






a number * which is clearly the result of a finite number of rational 
operations on a, 1 9 c. This completes the proof of the theorem. We 
have here the reason for the term net of rationality. 

It is well to recall at this point that our assumptions are not yet sufficient 
to identify the numbers associated with a net of rationality with the system 
of all ordinary rational numbers. We need only recall the example of the 
miniature geometry described in the Introduction, 2, which contained only 

* The expression for x cannot be indeterminate unless & = c. 



57, os] HOMOGENEOUS COORDINATES 163 

three points on a line. If in that triple-system geometry we perform the con- 
struction for the number 1 + 1 on any line in which we have assigned the 
numbers 0, 1, to the three points of the Hue in any way, it will be found 
that this construction yields the point 0. This, is due to the fact previously 
noted that in that geometry the diagonal points of a complete quadrangle 
are collinear. In every geometry to which Assumptions A, E, P apply we 

may construct the points 1 + 1, 1 + 1 + 1 thus forming a sequence of 

points which, with the usual notation for these sums, we may denote by 0, 1, 
2, 3, 4, .--. Two possibilities then present themselves: either the points, 
thus obtained are all distinct, in ^liich case the net R(Olcc) contains all the 
ordinary rational numbers ; or some point of this sequence coincides with one 
of the preceding points of the sequence, in which case the number of points 
in a net of rationality is finite. We shall consider this situation in detail in 
a later chapter, and will then add further assumptions. Here it should be 
emphasized that our results hitherto, and all subsequent results depending only 
on Assumptions A, E, P, are valid not only in the ordinary real or complex 
geometries, but in a much more general class of spaces, which are character- 
ized merely by the fact that the coordinates of the points on a line are the 
numbers of a field, finite or infinite. 

58. Homogeneous coordinates on a line. The exceptional character 
of the point J2, as the coordinate of which we introduced a symbol 
co with exceptional properties, often proves troublesome, and is, more- 
over, contrary to the spirit of protective geometry in which the points 
of a line are all equivalent ; indeed, the choice of the point & was 
entirely arbitrary. It is exceptional only in its relation to the opera- 
tions of addition, multiplication, etc., which we have defined in terms 
of it. In this section we will describe another method of denoting 
points on a line by numbers, whereby it is not necessary to use any 
exceptional symbol. 

As before, let a scale be established on a line by choosing any three 
points to be the points P^ J?, R ; and let each point of the line be 
denoted by its (nonhomogeneous) coordinate in a number system 
isomorphic with the points of the line. We will now associate with 
every point a pair of numbers (x v # 2 ) of this system in a given order, 
such that if x is the (nonhomogeneous) coordinate of any point dis- 
tinct from Jg,, the pair (x v x^f associated with the point is satisfies the 
relation x = xj%y. TVith the point R we associate any pair of the 
form (&, 0), where k is any number (&=0) of the number system 
isomorphic with the line. To every jwint of ilic line corresponds a pair 
of numbers, and to every pair of numbers in the field, except fJie pair 



164 ALGEBPtA OF POINTS [CHAP.TI 

(0, 0), corresponds a unique poiiit of the line. These two numbers are 
called homogeneous coordinates of the point with which they are 
associated, and the pair of numbers is said to represent the point. 
This representation of points on a line by pairs of numbers is not 
unique, since only the ratio of the two coordinates is determined; 
Le. the pairs (x r r 2 ) and (mx v mx 2 ) represent the same point for all 
values of m different from 0. The point P Q is characterized by the 
fact that ^ = ; the point H by the fact that # 2 = ; and the point 
% by the fact that x\ x y 

THEOREM 15. Li homogeneous coordinates a projectwity on a line is 
represented by a linear homogeneous transformation in two variables, 

(1) p*l~**i + lp, (orf-fc^O) 

V pZJ = 0^+. d%, 

where p is an arbitrary factor of proportionality. (A, E, P) 
Proof. By division, this clearly leads to the transformation 



cx + d 

provided x^ and # 2 are both different from 0. If # 3 = 0, the trans- 
formation (1) gives the point (a?/, a?/) = (a t c) ; i.e. the point H = 
(1, 0) is transformed by (1) into the point whose nonhomogeneous 
coordinate is a/c. And if # 2 '=0, we have in (1) (x p % 2 ) = (d, c); 
i.e. (1) transforms the point whose nonhomogeneous coordinate is 
rf/e into the point R. By reference to Theorem 11 the validity 
of the theorem is therefore established. 

As before, the matrix ( a , ) of the coefficients may conveniently 

be used to represent the projectivity. The double points of the pro- 
jectivity, if existent, are obtained in homogeneous coordinates as 
follows : The coordinates of a double point (& v # 2 ) must satisfy the 
equations px ^ = aXl + bz z , 

px z =cx : +dXy 
These equations are compatible only if the determinant of the system 



vanishes. This leads to the equation 

c d *~~ p 



58] 



HOMOGENEOUS COORDINATES 



165 



for the determination of the factor of proportionality p. This equa- 
tion is called the characteristic equation of the matrix representing 
the projeetivity. Every value of p satisfying this equation then leads 
to a double point when substituted in one of the equations (3) ; viz., 
if p 1 be a solution of the characteristic equation, the point 

( t /: r :c z ) = (- 5, a - pj = (d - p v - c) 
is a double point."* 

In homogeneous coordinates the cross ratio TkiAB, CD) of four 
points A = (a v 2 ), B = (l v & 2 ), C = (c v <g, D = (d v rf 2 ) is given by 



(ad) (bd) 

where the expressions (ac) 9 etc., are used as abbreviations for Q-fia l9 
etc. This statement is readily verified by writing down the above 
ratio in terms of the nonhomogeneous coordinates of the four points. 
We will close this section by giving to the two homogeneous coor- 
dinates of a point on a line an explicit geometrical significance. In 
view of the fact that the coordinates of such a point are not uniquely 
determined, a factor of proportionality being entirely arbitrary, there 
may be many such interpretations. On account of the existence of 
this arbitrary factor, we may impose a further condition on the coor- 
dinates (x v # 2 ) of a point, in addition to the defining relation xjx^x, 
where x is the nonhomogeneous coordinate of the point in question, 
We choose the relation x^+ # 2 = 1. If this relation is satisfied, 



= B( 10, oo a?), 



= B( loo, 00). 



Thus homogeneous coordinates subject to the condition o^ 4- # 2 = 1 
can be defined by choosing three points A 9 B, C arbitrarily, and letting 
x l = & (AB, CX} and a? 2 = B (AC, BX). The ordinary homogeneous 
coordinates would then be defined as any two numbers proportional 
to these two cross ratios. 

* This point is indeterminate only if 5 = c = and a = d. The projectiyity is 
then the identity. 



1 -1 




1 


1 




1 


1 -1 




1 


#! X 2 




! SS 2 


I -I 




I 


1 




1 


1 -1 




1 



166 ALGEBRA OE POINTS [CHAP. VI 

59. Projective correspondence between the points of two different 
lines. Hitherto we have confined ourselves, in the development of 
analytic methods, to the points of a single line, or, under duality, to 
the elements of a single one-dimensional primitive form. Suppose 
now that we have two lines I and m with a scale on each, and let 
the nonhomogeneous coordinate of any point of I be represented by 
x, and that of any point of m by y. The question then arises as to 
how a projective correspondence between the point x and the point y 
may be expressed analytically. It is necessary, first of all, to give a 
meaning to the equation y = x. In other words : What is meant by say- 
ing that two points $ on /, and y on m have the same coordinate ? 
The coordinate a? is a number of a field and corresponds to the point 
of which it is the coordinate in an isomorphism of this field with the 
field of points on the line 1. "We may think of this same field of 
numbers as isomorphic with the field of points on the line m. In 
bringing about this isomorphism nothing has been specified except 
that the fundamental points J, Q . determining the scale on m 
must correspond to the numbers 0, 1 and the symbol oo respectively. 
If the correspondence between the points of the line and the numbers 
of the field were entirely determined by the respective correspond- 
ences of the points ^, JJ, R> just mentioned, then we should know 
precisely what points on the two lines I and m have the same coor- 
dinates. It is not true of all fields, however, that this correspondence 
is uniquely determined when the points corresponding to 0, 1, oo are 
assigned.* It is necessary, therefore, to specify more definitely how 
the isomorphism between the points of m and the numbers of the 
field is brought about. One way to bring it about is to make use of 
the projectivity which carries the fundamental points 0, 1, oo of I 
into the fundamental points 0, 1, cc of m, and to assign the coordinate 
& of any point A of I to that point of m into which A is transformed 
by this projectivity. In this projectivity pairs of homologous points 
will then have the same coordinates. That the field of points and the 
field of numbers are indeed made isomorphic by this process follows 
directly from Theorems 1 and 3 in connection with Theorem 1, Cor., 
Chap. IV. TFe may now readily prove the following theorem : 

* This is shown by the fact that the field of all ordinary complex numbers can 
be isomorphic with itself not only by making each number correspond to itself, but 
also by making each number a + ib> correspond to its conjugate a i&. 



59] EXERCISES 167 

THEOREM 16. Any projecthe correspondence between the points [:e] 
and [y] of two distinct lines may le represented analytically by tlu 
relation y x 'by properly choosing the coordinates on the two lines. 
If the coordinates on the two lines are so related that tltc relation 
y = x represents a projectile correspondence, then any projective cor- 
respondence between the points of the tico lines is given by a relation 

CM + T> ,11 

y = _ , f cu ( fa -fc Q\ 

C& + d 

(A, E, P) 

Proof. The first part of the theorem follows at once from the pre- 
ceding discussion, since any projectivity is determined by three pairs 
of homologous points, and any three points of either line may be 
chosen for the fundamental points. In fact, we may represent any 
projectivity between the points of the two lines by the relation y = x, 
by choosing the fundamental points on I arbitrarily; the fundamental 
points on m are then uniquely determined. To prove the second part 
of the theorem, let TT be any given projective transformation of the 
points of the line I into those of m, and let TT O be the projectivity 
y sc, regarded as a transformation from m to /. The resultant 
TTJIT = ^ is a projectivity on ?, and may therefore be represented by 
se/ = (ax + 1) /(ex + d). Since TT = ir~ V a , this gives readily the result 
that TT may be represented by the relation given in the theorem. 

EXERCISES 

1. Give constructions for subtraction and division in the algebra of points 
on a line. 

2. Give constructions for the sum and the product of two lines of a pencil 
of lines in which a scale has been established. 

3. Develop the point algebra on a line by using the properties expressed in 
Theorems 2 and 4 as the definitions of addition and multiplication respec- 
tively. Is it necessary to use Assumption P from the beginning? 

4. Using Cor. 3 of Theorem 9, Chap. Ill, show that addition and multi- 
plication may be defined as follows : As before, choose three points P , P v 
P on a line Z as fundamental points, and let any line through Poo be labeled 
2,0. Then the sum of two numbers P x and P y is the point P x+ y into which P v 
is transformed by the elation with axis I* and center P m which transforms 
P into P.,.; and the product P^-Pj, is the point P^ into which P y is trans- 
formed by the homology with axis Z w and center P which transforms P l into 
P.,,. Develop the point algebra on this basis without using Assumption P, 
except in the proof of the commutativity of multiplication. 



168 ALGEBRA OF POIXTS [CHAP.TI 

5. If the relation ax = ly holds between four points a, 5, x, y of a line, 
show that we have Q(0fca, X#JT). Is Assumption P necessary for this result? 

6. Prove l>v direct computation that the expression -^ ^ : -8 % i s 

fc *>* 7* 7* - ** 

*1 -^ *g ^4 

unchanged in value when the four points x v x 2 , # s? o? 4 are subjected to any- 
linear fractional transformation u/ = 

cx + d 

7. Prove that the transformations 

X' = X, X' = ~, X' = 1-X, X'=-i-r, *' = r^7' *' = ^ 

A I A A "" JL A 

form a group. What are the periods of the various transformations of this 
group? (Cf. Theorem 13, Cor. 3.) 

8. If A, B, C, P v P 2 , , P n are any n -f 3 points of a line, show that 
every cross ratio of any four of these points can be expressed rationally in 
terms of the n cross ratios A, = B (AB< CP ), i = 1, 2, , n. When n = 1 
this reduces to Theorem 13, Cor. 3. Discuss in detail the case n = 2. 

9. If B: (a^o, x^x 4 ) = X, show that 

1-X = 1 X 

^-^"^-^2 *3-*l' 

The relation of Cor. 3 of Theorem 13 is a special case of this relation. 

10. Show that if R (AB, CD) = E (AB< DC), the points form a harmonic 

set H (-43, CD). 

11. If the cross ratio E (.15, CD) = X satisfies the equation X 2 -X + 1 = 0, 

then B (.IB, CD) = B (4 C 5 D) = B (AD, JJC) = X, 

and B (AB.DC) = B (-4 C, BD) = B (4D, CB) = - X 2 . 

12. If A, B, T, F, Z are any five points on a line, show that 

B (-4B, AT) B (4B, rZ) B (-4B, ZZ) = 1. 

13. State the corollaries of Theorem 11 in homogeneous coordinates. 

14. By direct computation show that the two methods of determining the 
double points of a projectivity described in 54 and 58 are equivalent. 

15. If Q(lBC,ATZ),then 

B (AX, FC) + B (BY, ZA) -f B (CZ 9 ZB) = 1. 

16. If M v Jf 2 , J/ 3 are any three points in the plane of a line I but not on 
L the cross ratios of the lines ?, -PJ/p PM^ PM Z are different for any two 
points P on L 

17. If A , B are any two fixed points on a line Z, and X, F are two variable 
points such that B (AB, XY) is constant, the set [Z j is projective with the 
set [F]. 



CHAPTER VII 



COORDINATE SYSTEMS IN TWO- AND THREE-DIMENSIONAL* 

FORMS 

60. Nonhomogeneotts coordinates in a plane. In order to repre- 
sent the points and lines of a plane analytically we proceed as follows : 
Choose any two distinct lines of the plane, which we will call the 
axes of coordinates, and determine on each a scale (48) arbitrarily, 
except that the point of intersection of the lines shall be the 
0-point on each scale (fig. 78). This point we call the origin. Denote 
the fundamental 
points on one of 
the lines, which 
we call the x-axis } 
by O a , I,,., oOp ; and 
on the other line, 
which we will call 
the y-axis, by O y , 
l y , oo y . Let the 

line cc^cc, be de- ^ Tg 

noted by Z. 

Now let P be any point in the plane not on Z. Let the lines Pcc y 
and Poo,,, meet the #-axis and the y-axis in points whose nonhornoge- 
neous coordinates are a and ~b respectively, in the scales just estab- 
lished. The two numbers a, "b uniquely determine and are uniquely 
determined by the point P. Thus every point in the plane not on l m 
is represented by a pair of numbers ; and, conversely, every pair of 
numbers of which one belongs to the scale on the as-axis and the 
other to the scale on the y-axis determines a point in the plane (the 
pair of symbols co^ oo y being excluded). The exceptional character 
of the points on l m mil be removed presently ( 63) by considera- 
tions similar to those used to remove the exceptional character of 

*A11 the developments of this chapter are on the basis of Assumptions 
A,E,P. 

169 




170 COORDINATE SYSTEMS [CHAP, vn 

the point cc in the case of the analytic treatment of the points of a 
line ( 58). The two numbers just described, determining the point 
P, are called the nonhomogencous coordinates of P with reference to 




FIG. 79 

the two scales on the x- and the ^-axes. The point P is then repre- 
sented analytically by the symbol ( , 6). The number a is called the 
^-coordinate or the abscissa of the point, and is always written first 
in the symbol representing the point; the number & is called the 
^/-coordinate or the ordinate of the point, and is always written last 
in this symbol. 

The plane dual of the process just described leads to the corre- 
sponding analytic representation of a line in the plane. For this pur- 
pose, choose any two distinct points in the plane, which we will call 
the centers of coordinates ; and in each of the pencils of lines with 
these centers determine a scale arbitrarily, except that the line o join- 
ing the two points shall be the 0-line in each scale. This line we call 
the origin. Denote the fundamental lines on one of the points, which 
we will call the u-center, by O tt , l a , o> tt ; and on the other point, which 
we will call the v-center, by O p , l r , cc p . Let the point of intersection 
of the lines oo w , oo r be denoted by & (fig. 79). 

Sow let I be any line in the plane not on R. Let the points lco v 
and /oc tt be on the lines of the ^-center and the ^-center, whose non- 
homogeneous coordinates are m and n respectively in the scales just 
established. The two numbers m, n uniquely determine and are 
uniquely determined by the line L Thus every line in the plane not 
on is represented by a pair of numbers ; and, conversely, every pair 
of numbers of which one belongs to the scale on the ^-center and the 
other to the scale on the ^-center determines a line in the plane (the 
pair of symbols OO M , oo p being excluded). The exceptional character 



60, 01] 



COORDINATES IS A PLANE 



171 



of the lines on j will also be removed presently. The two numbers 
just described, determining the line /, are called the nonhomogcueous 
coordinates of I with reference to the two scales on the it- and 
^-centers. The line I is then represented analytically by the symbol 
\m, %]. The number m is called the u-co'drdinate of the line, and is 
always written first in the symbol just given ; the number 11 is called 
the ^-coordinate of the line, and is always written second in this 
symbol. A variable point of the plane will frequently be represented 
by the symbol (x 9 y) ; a variable line by the symbol [u, r]. The coor- 
dinates of a point referred to two axes are called point coordinates ; 
the coordinates of a line referred to two centers are called line coor- 
dinates. The line Z M and the point R are called the singular line and 
the singular point respectively. 

61* Simultaneous point and line coordinates. In developing further 
our analytic methods we must agree upon a convenient relation 
between the axes and centers of the point and line coordinates respec- 
tively. Let us consider any triangle in the plane, say with vertices 




0, Z7, V. let the lines OU and OF be the y- and 0-axes respectively, 
and in establishing the scales on these axes let the points 7, V be 
the points cc y , co a respectively (fig. 80), "Further, let the points J7, 7 
be the ^-center and the ^-center respectively, and in establishing the 



172 COORDINATE SYSTEMS [CHAP, vn 

scales on these centers let the lines UO, TO be the lines co u , cc v 
respectively. The scales are now established except for the choice of 
the 1 points or lines in each scale. Let us choose arbitrarily a point 
I,, on the #-axis and a point l y on the y-axis (distinct, of course, from 
the points 0, U, F). The scales on the axes now being determined, 
we determine the scales on the centers as follows : Let the line on 
V and the point l x on the a-axis be the line l u ; and let the line 
on F and the point l v on the y-axis be the line l p . All the scales 
are now fixed. Let T be the projectivity (59, Chap. VI) between 
the points of the i'-axis and the lines of the w-center in which points 
and lines correspond when their x- and z^-coordinates respectively 
are the same. If TT' is the perspeetivity in which every line on the 
^-center corresponds to the point in which it meets the #-axis, the 
product TT'TT transforms the <?>axis into itself and interchanges and 
GO.,, and I, and l x . Hence TrV is the involution a/ = 1/#. Hence 
it follows that the line on U whose coordinate is u is on the point of 
the x-axis whose coordinate is 1/u; and the point on the x-axis 
whose coordinate is x is oil the line of the u-center whose coordinate 
is I/a*. This is the relation between the scales on the #-axis and 
the 2^-center. 

Similar considerations with reference to the y-axis and the -^-center 
lead to the corresponding result in this case : The line on V 'whose coor- 
dinate is v is on that point of the y-axis whose coordinate is 1/v; 
and the point of the y-axis whose coordinate is y is on that line of the 
v-ceiiter whose coordinate is 1/y. 

62. Condition that a point be on a line. Suppose that, referred to 
a system of point-and-line coordinates described above, a point P has 
coordinates (a, V) and a line / has coordinates \m, n]. The condition 
that P be on I is now readily obtainable. Let us suppose, first, that 
none of the coordinates a, I, m, n are zero. We may proceed in either 
one of two dual ways. Adopting one of these, we know from the 
results of the preceding section that the line [m, n] meets the #-axis 
in a point whose ^-coordinate is 1/m, and meets the 7/-axis in a 
point whose ^-coordinate is 1/n (fig. 81). Also, by definition, the 
line joining P = (a, 1) to V meets the #-axis in a point whose ^-coor- 
dinate is a ; and the line joining P to Z7 meets the y-axis in a point 
whose y-coorclinate is &. If P is on I, we clearly have the following 
perspeetivity : 



62] 



COOEDINATES IN A PLANE 



173 



= 
A 



(1) 
Hence we have 



which, when expanded (Theorem 13, Chap. VI), gives for the desired 
condition 

(3) ma + nl + 1 = 0. 

This condition has been shown to be necessary. It is also sufficient, 
for, if it is satisfied, relation (2) must hold, and hence would follow 
(Theorem 13, Cor. 5, Chap. TI) 

-- Oacc x j- -- Oac A 
m A n v 

But since this projectivity has the self-corresponding element 0, it 
is a perspectivity which leads to relation (1). But this implies that 
P is on L 




FIG. 81 

If now & = (&= 0), we have at once 5 = l/n ; and if & = (a^ 0), 
we have likewise a = 1/m for the condition that P be on But 
each of these relations is equivalent to (3) when a = and J = 
respectively. The combination a = 0, 5 = gives the origin which 
is never on a line [m, ?i] where m^O^n. It follows in the same 
way directly from the definition that relation (3) gives the desired 
condition^ if we have either m = or n = 0. The condition (3) is 
then valid for all cases, and we have 



174 COOEDIXATE SYSTEMS [CHAP.VH 

THEOREM 1. The necessary and sufficient condition that a point 
P = (a, 6) le on a line I = [?n, n] is that the relation ma + ^5 + 1 = 
be satisfied. 

DEFINITION. The equation DEFINITION. The equation 
which is satisfied by the eoordi- which is satisfied by the coordi- 
nates of all the points on a given nates of all the lines on a given 
line and no others is called the point and no others is called the 
point equation of the line. line equation of the point. 

COEOLLARY 1. The point equa- COROLLAEY 1'. The line equa- 
tion of the line [m, n] is tion of the point (a, V) is 
m + ny + 1 = 0. au + bv + 1 = 0. 

EXERCISE 
Derive the condition of Theorem 1 by dualizing the proof given. 

63. Homogeneous coordinates in the plane. In the analytic repre- 
sentation of points and lines developed in the preceding sections the 
points on the line Z7F=0 and the lines on the point were left 
unconsidered. To remove the exceptional character of these points 
and lines, we may recall that in the case of a similar problem in the 
analytic representation of the elements of a one-dimensional form we 
found it convenient to replace the nonhomogeneous coordinate x of 
a point on a line by a pair of numbers x v # 2 whose ratio xjx^ was 
equal to x(x = oo), and such that x z = when x = oo. 

A similar system of homogeneous coordinates can be established for 
the plane. Denote the vertices 0, U 9 V of any triangle, which we will 
call the triangle of reference, by the " coordinates " (0, 0, 1), (0, 1, 0), 
(1, 0, 0) respectively, and an arbitrary point T, not on a side of the 
triangle of reference, by (1, 1, 1). The complete quadrangle OUVT 
is called the frame of reference * of the system of coordinates to be 
established. The three lines UT, VT, OT meet the other sides of the 
triangle of reference in points which we denote by l x = (1, 0, 1), 
1 F =(0, 1, 1), 1,='(1, 1, 0) respectively (fig. 82). 

We will now show how it is possible to denote every point in the 
plane by a set of coordinates (x v # 2 , # 8 ). Observe first that we have 
thus far determined three points on each of the sides of the triangle 

* Frame of reference is a general term that may be applied to the fundamental 
elements of any coordinate system. 



63] HOMOGENEOUS COORDINATES 175 

of reference, viz.: (0, 0, 1), (0, 1, 1), (0, 1, 0) on OU; (0, 0, 1), (1, 0, 1), 
(1, 0, 0) on OF; and (0, 1, 0), (1, 1, 0), (1, 0, 0) on CT. The coordi- 
nates which we have assigned to these points are all of the form 
(x v # 2 , # 3 ). The three points on OU are characterized by the fact that 
x l = 0. Fixing attention on the remaining coordinates, we choose the 
points (0, 0, 1), (0, 1, 1), (67 X 0} as the fundamental points (0, I)/ 
(1, 1), (1, 0) of a system of homogeneous coordinates on the line OU. 
If in this system a point has coordinates (/, 7/1), we denote it in our 
planar system by (0, 1, m). In like manner, to the points of the other 
two sides of the triangle of reference may be assigned coordinates of 
the form (Is, 0, m) and (k, I, 0) respectively. We have thus assigned 
coordinates of the form (x v x^ r s ) to aH the points of the sides of the 
triangle of reference. Moreover, the coordinates of every point on 
these sides satisfy one of the three relations 2^= 0, ^ 2 = 0, # 3 = 0. 

Now let P be any point in the plane not on a side of the triangle 
of reference. P is uniquely determined if the coordinates of its pro- 
jections from any two of the vertices of the triangle of reference on 
the opposite sides are known. Let its projections from U and V on 
the sides OY and OU be (k, 0, n) and (0, 1', n r ) respectively. Since 
under the hypothesis none of the numbers k> %, l f , n f is zero, it is 
clearly possible to choose three numbers (x v % 2 , x z ) such that x l : x 3 
= &:?&, and # a : x z = V : n f . We may then denote P by the coordinates 
(x v x# x s ). To make this system of coordinates effective, however, 
we must show that the same set of three numbers (x v x & x & ) can be 
obtained by projecting P on any other pair of sides of the triangle 
of reference. In other words, we must show that the projection of 
P = (x^ x# x s ) from on the line TJY is the point (x v x 2 , 0). Since 
this is clearly true of the point jP = (l, 1, 1), we assume P distinct 
from 51 Since the numbers # 1} # 2 , % 3 are all different from 0, let us 
place x 1 : x z = a?, and # 2 : # 3 = y, so that x and y are the nonhomoge- 
neous coordinates of (x v 0, # 3 ) and (0, # 2 , x s ) respectively in the scales 
on 07 and OU defined by = O x , l a , F= oo r and = O y , l y , U= ay 
Finally, let OP meet U Y in the point whose nonhomogeneous coor- 
dinate in the scale defined by ?7= O z , l z , F= co z is %} and let OP 
meet the line 1 X U in A. We now have 

V 

* * * yf * z - J y^y y > 



176 COOBDIKATE SYSTEMS [CHAP.VH 

where C is the point in which TA meets OU. This projectivity 
between the lines UV and OU transforms O s into co y , cc s into O y , and 
1 3 into l y . It follows that C has the coordinate 1/2 in the scale on 
Or. We have also 



which gives 

= B 



Substituting x = ^ : # 3 , and ?/ = # 3 : <? 3 , this gives the desired relation 
z = x t : </.' 
follows : 



z = x t : </.' 2 . The results of this discussion may be summarized as 




FIG. 82 

THEOREM 2. DEFINITION. If P is any point not on a side of the 
triangle of reference OUV, there exist three numbers x v x z > x s ( a ^ Dif- 
ferent from 0) such that the projections of P from the vertices 0.17. 
T on the opposite sides have coordinates (x v x^ 0), (x v 0, # 8 ), (0, # 2 , # 8 ) 
respectively. TJiese three numbers are called the homogeneous coordf- 
"nafes^qflP, and P is denoted ly (x v x^ x^ Any set of three numbers 
(not all equarto'Q) determine uniquely a point whose (homogeneous) 
coordinates they are. 

The truth of the last sentence in the above theorem follows from 
the fact that, if one of the coordinates is 0, they determine uniquely 
a point on one of the sides of the triangle of reference ; whereas, if 
none is equal to 0, the lines joining U to (x v 0, # 8 ) and V to (0, # 2 ,j*! 8 ) 
meet in a point whose coordinates by the reasoning above are (x v x y xj). 



63] HOMOGENEOUS COOEDDsATES ITT 

COKOLLAKY. The coordinates (x v 2 2 , z s ) and (l\c v k% 2 , hc 2 ) tktcmiiiu 
the same point, ifk is not 0. 

Homogeneous line coordinates arise by dualizing the above discus- 
sion in the plane. Thus we choose any quadrilateral in the plane as 
frame of reference, denoting the sides by [1, 0, 0], [0, 1, 0], [0, 0, 1], 
[1, 1, 1] respectively. The points of intersection with [1, 1, 1] of the 
lines [1, 0, 0], [0, 1, 0], [0, 0, 1] are joined to the vertices of the tri- 
angle of reference opposite to [1, 0, 0], [0, 1, 0], [0, 0, 1] respectively 
by lines that are denoted by [0, 1, 1], [1, 0, 1], [1, 1, 0]. The three 
lines [1, 0, 0], [1, 1, 0], [0, 1, 0] are then taken as the fundamental 
lines [1, 0], [1, 1], [0, 1] of a homogeneous system of coordinates in 
a flat penciL If in this system a line is denoted by [u v wj, it is 
denoted in the planar system by [w 1? 2 , 0]. In like manner, to the 
lines on the other vertices are assigned coordinates of the forms 
[0, ^ 2 , J and [u v 0, J respectively. As the plane dual of the 
theorem and definition above we then have at once 

THEOREM 2 f . DEFINITION. If 1 is any line not on a, vertex of the 
triangle of reference, there exist three numbers u v u 09 u s all different 
from zero, such that the traces of I on the three sides of the triangle of 
reference are projected from the respective opposite, vertices by the lines 
\u v u z) 0], [u v 0, ttj, [0, 2 , ?/]. FJiese three numbers are called the 
homogeneous coordinates of /, and I is denoted ly [u 19 u^ w s ]. Any 
set of three numbers (not all zero) determine uniquely a line whose 
coordinates they are. 

Homogeneous point and line coordinates may be put into such 
a relation that the condition that a point (x^ % 9 or s ) be on a line 
[!, a , tt s ] is that the relation u^+ ^ 2 # 3 -f u 3 x s = be satisfied. We 
have seen that if (x v x v x 9 ) is a point not on a side of the triangle of 
reference, and we place x = xjx z> and y = # 2 /# 3 , the numbers (x, y) 
are the nonhomogeneous coordinates of the point (x v x s , a? 3 ) referred 
to OF as the re-axis and to OU as the y-asis of a system of nonho- 
mogeneous coordinates in which the point T=(l, 1, 1) is the point 
(1, 1) (0, U, 7 being used in the same significance as in the proof of 
Theorem 2). By duality, if [u v u# -M 3 ] is any line not on any vertex 
of the triangle of reference, and we place u = uju^ and v == uju y 
the numbers [^, v] are the nonhomogeneous coordinates of the line 
[u v u& J referred to two of the vertices of the triangle of reference 



178 



COORDINATE SYSTEMS 



[CHAP. YII 



as CT-eenter and F-center respectively, and in which the line [1, 1, 1] 
is the line [1, 1]. If, now, we superpose these two systems of nonhom- 
ogeneous coordinates in the way described in the preceding section, 
the condition that the point (x, y) be on the line [u 9 v] is that the 
relation uz + vy + l = Q be satisfied (Theorem 1). It is now easy to 
recognize the resulting relation between the systems of homogeneous 
coordinates with which we started. Clearly the point (0, 1, 0) = U is 
the *7-center, (1, 0, 0) = F is the F-center, and (0, 0, 1) = is the third 




FIG. 83 



vertex of the triangle of reference in the homogeneous system of line 
coordinates. Also the line whose points satisfy the relation # x = is 
the line [1, 0, 0], the line for which # s = is the line [0, 1, 0], and 
the line for which # 3 = is the line [0, 0, 1]. Finally, the line 
[1, 1] = [1, 1, 1], whose equation in nonhomogeneous coordinates is 
+ y + 1 = 0, meets the line- x^ = in the point (0, 1, 1), and the 
line # 2 = in the point ( 1, 0, 1). The two coordinate systems are 
then completely determined (fig. 83). 

It now follows at once from the result of the preceding section 
that the condition that (x v x z , # 3 ) be on the line [u v u^ u z ] is 



ufli 4- 



0, if none of the coordinates x lt # 2 , # 8 , u v 



G3] HOMOGENEOUS COORDINATES 179 

is zero. To see that the same condition holds also when one (or more) 
of the coordinates is zero, we note first that the points (0, 1, 1), 
( 1, 0, 1), and ( 1, 1, 0) are collinear. They are, in fact (fig. 83), on 
the axis of perspectivity of the two perspective triangles OTFand 
la-lylc, the center of perspectivity being J, It is now clear that 

the line [1, 0, 0] passes through the point (0, 1, 0;, 
the line [0, 1, 0] passes through the point (1, 0, 0), 
the line [1, 1, 0] passes through the point ( 1, 1, 0). 

There is thus an involution between the points (x v ;<? 2 , 0) of the line 
ac 9 = and the traces (j?/, %.!, 0) of the lines with the same coordinates, 
and this involution is given by the equations 



In other words, the line \u v u, 0] passes through the point ( u z) it v 0). 
Any other point of this line (except (0, 0, 1)) has, by definition, the 
coordinates ( ^ 3 , u v x s ). Hence all points (x^ # 2 , # 3 ) of the line 
[u v u z , 0] satisfy the relation 7^ 4- ?M*2 + Vs 0- The same argu- 
ment applied when any one of the other coordinates is zero estab- 
lishes this condition for all cases. A system of point and a system 
of line coordinates, when placed in the relation described above, will 
be said to form a system of homogeneous point and line coordinates in 
the plane. The result obtained may then be stated as follows : 

THEOREM 3. In a system of homogeneous point and line coordinates 
in a plane the necessary and sufficient condition that a point (% v # 2 , .T 3 ) 
le on a line [u v 2 , ^ 3 ] is that the relation u^ + 11^ 4- u^ = le 
satisfied. 

COROLLAEY. The equation of a line through the origin of a system 
of nonhomogeneous coordinates is of the form mx 4- ny = 0. 

EXERCISES 

1. The line [1, 1, 1] is the polar of the point (1, 1, 1) with regard to the 
triangle of reference (cf. p. 46). 

2. The same point is represented by (a l9 a 2 , a s ) and (& r 5 2 , 5 8 ) if and only 

if the two-rowed determinants of the matrix (J 1 ^ 2 ^ 3 ) are all zero. 

V*l &2 V 

3. Describe nonhomogeneous and homogeneous systems of line and plane 
coordinates in a bundle by dualizing in space the preceding discussion. In 
such a bundle what is the condition that a line be on a plane ? 



180 



COORDINATE SYSTEMS 



[CHAP. Til 



64. The line on two points. The point on two lines. Given two 
points, A = (a v a z , a,) and B = (b v & 2 , & s ), the question now arises as 
to what are the coordinates of the line joining them; and the dual 
of this problem, namely, given two lines, m = [m v m 2 , m s ] and n = 
[n v n n , '7& 8 ], to find the coordinates of the point of intersection of the 
two lines. 

THEOREM 4. The equation of THEOREM 4'. The equation of 
the line joining the points (a^a^a^) the point of intersection of the 
and (b v 5 2> 6 3 ) is lines [m x ,m 2 ,m 8 ] and [n v n^n s ] is 



= 0. 



Proof. "When these determinants are expanded, we get 

3 =0, 



ni 



respectively. The one above is the equation of a line, the one below 
the equation of a point. Moreover, the determinants above both 
evidently vanish when the variable coordinates are replaced by the 
coordinates of the given elements. The expanded form just given 
leads at once to the following: 

COROLLAET V. The coordinates 
of the point of intersection of the 
lines [m v m 2 , m s ], [n v n 2 , w 8 ] are 



COROLLAEY 1. The coordinates 
of the line joining the points 
(a v 2 , a,), (b v 6 2 , b s ) are 



There also follows immediately from this theorem: 
COROLLARY 2. The condition COROLLARY 2'. The condition 
that three points A, B, C be col- that three lines m, n, p be con- 
linear is current is 



= 0. 



. 



-0. 



EXAMPLE. Let us verify the theorem of Desargues (Theorem 1, Chap. II) 
analytically. Choose one of the two perspective triangles as triangle of refer- 
ence, say A' = (0, 0, 1), J5 7 = (0, 1, 0), C" = (1, 0, 0), and let the center of per- 
spectivity be P = (1, 1, 1). If the other triangle is ABC, we may place 



64,65] PEOJECTITE PENCILS 181 

A = (1, 1, rt), B = (1, ft, 1), C = (c, 1, 1) : for the equation of the line PA' 
is x l j: 2 = ; and since J. is, by hypothesis, on this line, its first two coordi- 
nates must be equal, and may therefore be assumed equal to 1 ; the third 
coordinate is arbitrary. Similarly for the other point*. Now, from the above 
theorems and their corollaries we readily obtain in succession the following : 

The coordinates of the line A'& are [1, 0, 0]. 

The coordinates of the line AB are [1 ah, a 1, & 1]. 

Hence the coordinates of their intersection C" are 

C"'=(0. l -&,,i-l). 
Similarly, we find the coordinates of the intersection A" of the lines B'C\ BC 

tobe J"=(l-e,6-l,0): 

and, finally, the coordinates of the intersection B" of the lines C'A'* CA to be 

3"= (e- 1,0,1 -a). 
The points A", B'\ C" are readily seen to satisfy the condition for collinearity. 

EXERCISES 

1. Work through the dual of the example just given, choosing the sides of 
one of the triangles and the axis of perspectivity as the fundamental lines of 
the system of coordinates. Show that the work may be made identical, step 
for step, with that above, except for the interpretation of the symbols. 

2. Show that the system of coordinates may be so chosen that a quadrangle- 
quadrilateral configuration is represented by all the sets of coordinates that 
can be formed from the numbers and 1. Dualize. 

3. Derive the equation of the polar line of any point with regard to the 
triangle of reference. Dualize. 

65. Pencils of points and lines.. Projectivity. A convenient ana- 
lytic representation of the points of a pencil of points or the lines of 
a pencil of lines is given by the following dual theorems : 

THEOREM 5. Any point of a THEOREM 5'. Any line of a 
pencil of points may le repre- pencil of lines may "be represented 
sented ly ly 

P = (X 2 a 1 + \l v X 2 a 3 + \l z , p = [|* a m 1 



where A = (a x & 2 , a s ) and B = where m = [m v m 2 , m g ] and n = 
(&!> 5 2 , 6 S ) are any two distinct [n v n, raj are any two distinct 
points of the pencil. lines of the pencil. 

Proof. We may confine ourselves to the proof of the theorem on 
the left. By Theorem 4, Cor. 2, any point (x v x^ # s ) of the pencil of 
points on the line AB satisfies the relation 



182 COORDINATE SYSTEMS [CHAP.VII 

\ x i x * x s\ 

(1) fl x s ,=0. 

A &s 5 a! 
We may then determine three numbers p, X 2 ', X/, such that we have 

(2) p* 1 = \'a 1 +Xft- (* = 1,2,3) 

The number p cannot be under the hypothesis, for then we should 
have from (2) the proportion x : a a : 3 = 5^ S 3 : 5 8 , which would imply 
that the points A and B coincide. We may therefore divide by p. 
Denoting the ratios \ f /p and \[/p by X 2 and \, we see that every 
point of the pencil may be represented in the manner specified. 
Conversely, every point whose coordinates are of the form specified 
clearly satisfies relation (1) and is therefore a point of the pencil. 

The points A and B in the above representation are called the "base 
points of this so-called parametric representation of the elements of 
a pencil of points. Evidently any two distinct points may be chosen 
as base points in such a representation. The ratio \/\ is called the 
parameter of the point it determines. It is here written in homoge- 
neous form, which gives the point A for the value \=0 and the 
point B for the value X 2 = 0. In many cases, however, it is more 
convenient to write this parameter in nonhomogeneous form, 

P = (a, + 7J> V a 2 + \h, a, + X6 3 ), 

which is obtained from the preceding by dividing by X 2 and replacing 
\/\ by ^- I n this representation the point B corresponds to the 
value X = oc. "We may also speak of any point of the pencil under 
this representation as the point X t : X 2 or the point X when it corre- 
sponds to the value X X /X 2 = X of the parameter. Similar remarks and 
the corresponding terminology apply, of course, to the parametric 
representation of the lines of a flat pencil. It is sometimes convenient, 
moreover, to adopt the notation A -f- XJ? to denote any point "of the 
pencil whose base points are A, B or to denote the pencil itself ; also, 
to use the notation m, + pn to denote the pencil of lines or any line 
of this pencil whose base lines are m, n. 

In order to derive an analytic representation of a projectivity 
between two one-dimensional primitive forms in the plane, we seek 
first the condition that the point X of a pencil of points A + \B be 
on the line p of a pencil of lines m + pn. By Theorem 3 the condition 
that the point X be on the line /t is the relation 



65] PKOJECTIYE PENCILS 183 



When expanded this relation gives 

t=3 i=3 

l a l = 0. 



This is a bilinear equation whose coefficients depend only on the coor- 
dinates of the base points and base lines of the two pencils and not 
on the individual points for which the condition is sought. Placing 



this equation becomes 6>X + Dp. -4X B = 0, 

which may also be written* 
* 



The result may be stated as follows: Any perspective relation "between 
two one-dimensional primitive forms of different "kinds is obtained by 
establishing a protective correspondence between the parameters of the 
two forms. Since any protective correspondence between two one- 
dimensional primitive forms is obtained as the resultant of a sequence 
of such perspectivities, and since the resultant of any two linear frac- 
tional transformations of type (1) is a transformation of the same 
type, we have the following theorem : 

THEOREM 6. Any projectile correspondence "between two one-dimen- 
sional primitive forms in the plane is obtained "by establishing a 
protective relation 



between the parameters p> X of the two forms. 

In particular we have 

COROLLARY 1. Any projectimty in a one-dimensional primitive 
form in the plane is given by a relation of the form 



where X is the parameter of the form, 



^ pi does not vanish because the correspondence between 
A ana /i is (1,1). 



184 COOBDISrATE SYSTEMS [CHAP.TII 

COROLLARY 2. If X 1? X 2 , X 3 , \ are the parameters of four elements 
A v A*, A 3 , A of a one-dimensional primitive form,) the cross ratio 
B (A^A^ A S A 4 ) is given ly 



A projectivity between two different one-dimensional forms may 
be represented in a particularly simple form by a judicious choice of 
the base elements of the parametric representation. To fix ideas, let 
us take the case of two projective pencils of points. Choose any two 
distinct points A 9 B of the first pencil to be the base points, and let 
the homologous points of the second pencil be base points of the 
latter. Then to the values X = and X = oo of the first pencil must 
correspond the values p = and \L = oo respectively of the second. 
In this case the relation of Theorem 6, however, assumes the form 
ju, = &X. Hence, since the same argument applies to any distinct 
forms, we have 

COEOLLARY 3. If tico distinct projective one-dimensional primitive 
forms in the plane are represented parametrically so that the base 
elements form two homologous pairs, the projeetivity is represented by 
a relation of the form p = k\ between the parameters p, X of the two 
forms. 

This relation may be still further simplified. Taking again the case 
discussed above of two projective pencils of points, we have seen that, 
in general, to the point (a^ + l v a 2 + J a , a s + J 8 ), ie. to X = 1, corre- 
sponds the point (0/ + &J/, tf 2 '-f-&& 2 ', #/+&&), ie. the point /-& = . 
Since the point !?'=(&/, J/, 6 g ') is also represented by the set of coordi- 
nates (&/, &/, &/), it follows that if we choose the latter values for the 
coordinates of the base point B 1 , to the value X = 1 will correspond 
the value fi = 1, and hence we have always p. = X. In other words, 
we have 

COROLLARY 4. If two distinct one-dimensional forms are protective, 
the lase elements may le so chosen that the parameters of any two 
homologous elements are equal. 

Before closing this section it seems desirable to call attention 
explicitly to the forms of the equation of any line of a pencil and of 
the equation of any point of a pencil which is implied by Theorem 5 ; 
and Theorem 5 respectively. If we place m = m^ + w 2 # 2 H- m B x s and 



65, 66] EQUATION OF A COXIC 185 

71 = 71^+ nx 2 + }2 3 c>: 3 , it follows from the first theorem mentioned 
that the equation of any line of the pencil whose center is the inter- 
section of the lines m = 0, n = is given by an equation of the form 
ni + fin = 0. Similarly, the equation of any point of the line joining 
A. = a l u l + 2 3 + 3 3 = and B = J^ + 6 a w a -j- Zu? s = is of the 
form A +\B = 0. 

66. The equation of a conic. The results of 65 lead readily to 
the equation of a conic. By this is meant an equation in point (line) 
coordinates which is satisfied by all the points (lines) of a conic, and 
by no others. To derive this equation, let A 9 B Le two distinct points 
on a conic, and let 

tn = m^i^-r wzA' a -t- ?/u,c = 0, 

(1) d = w^ + M tf v 2 + /v' 3 = 0, 



be the equations of the tangent at A, the tangent at J?, and the line 
A3 respectively. The conic is then generated as a point locus by 
two protective pencils of lines at A and B, in which m, p at A are 
homologous with jp, n at B respectively. This projectivity between 
the pencils 

7 m + \2 } - > 

^ ; p + IAH Q 

is given (Theorem 6, Cor. 3) by a relation 

(3) /A = &X 

between the parameters /*, X of the two pencils. To obtain the equa- 
tion which is satisfied by all the points of intersection of pairs of 
homologous lines of these pencils, and by no others, we need simply 
eliminate ^, X between the last three relations. The result of this 
elimination is 

(4) /-#wztt=0, 

which is the equation required. By multiplying the coordinates of 
one of the lines by a constant we may make k = 1. 

Conversely, it is obvious that the points which satisfy any equation 
of type (4) are the points of intersection of homologous lines in the 
pencils (2), provided that /z, = k\. If m, n, p are fixed, the condition 
that the conic (4) shall pass through a point (a v a v a z ) is a linear 
equation in Je. Hence we have 



186 



COOEDIKATE SYSTEMS 



[CHAP. VII 



THEOEEM 7. If m = 0, n = 0, 
p = are the equations of two 
distinct tangents of a conic and 
the line joining (heir points of con- 
tact -respectively, the point equa- 
tion of the conic is of the form 

p~ kmn = 0. 

The coefficient "k is determined by 
any third point on the conic. Con- 
versely, the points which satisfy 
an equation of the above form 
constitute a conic of which m = 
and n = are tangents at points 
on p = 0. 

COROLLAEY. By properly choos- 
ing the triangle of reference, the 
point equation of any conic may 
te put in the form 



THBOEEM 7.' If A = 0, B = 0, 
C = are the equations of two 
distinct points of a conic and the 
intersection of the tangents at these 
points respectively, the line equa- 
tion of the conic is of the form 



where r^= 0, # 3 = are two tan- 
gents, and # 2 = is the line join- 
ing their points of contact. 



The coefficient k is determined ly 
any third line of the conic. Con- 
versely, the lines which satisfy an 
equation of the above form consti- 
tute a conic of which A = and 
B = are points of contact of the 
tangents through C = 0. 

COEOLLAEY. By properly choos- 
ing the triangle of reference, the 
line equation of any conic may 
be put in the form 

w 2 2 2^3 = 0, 

where ^= 0, % 3 = are two points, 
and u 2 = is the intersection of 
the tangents at these points. 



It is clear that if we choose the point (1, 1, 1) on the conic, we have 
k = 1. Supposing the choice to have been thus made, we inquire 
regarding the condition that a line [u v u z , u^ be tangent to the conic 



This condition is equivalent to the condition that the line whose 

equation is 

^a^-f- ^ 2 # 2 + U B % & = 

shall have one and only one point in common with the conic. Elimi- 
nating # 8 between this equation and that of the conic, the points 
common to the line and the conic are determined by the equation 

UjSB? + Ufl^ + 1(, 3 %j = 0. 

The roots of this equation are equal, if and only if we have 

u% 4 u^ = 0. 



tK5,C7] LIXEAE TEAXSFOP^IATIOXS 1ST 

Since this is the line equation of all tangents to the conic, and since 
it is of the form given in Theorem 7', Cor., above, we have here a new 
proof of the fact that the tangents to a point conic form a li/ic conic 
(cf. Theorem 11, Chap. V). 

When the linear expressions for M, n, p are substituted in the equa- 
tion j> 2 ~kmn = of any conic, there results, when multiplied out, a 
homogeneous equation of the second degree in x v # 2 , tt i s , which may 
be written in the form 

(1) a n x{ + a 22 ,<7 + a^c* + 2 a 12 ^ 2 + 2 a^c^ + 2 a^x^ = 0. 

We have seen that the equation of every conic is of this form. We 
have not shown that every equation of this form represents a conic 
(see 85, Chap. IX). 

EXERCISE 
Shcrw that the conic 



degenerates into (distinct or coincident) straight lines, if and only if we have 

Dualize. (A, E, P, H ) 

67. Linear transformations in a plane. We inquire now concern- 
ing the geometric properties of a linear transformation 



(1) 



Such a transformation transforms any point (x v # 2 , # 3 ) of the plane 
into a unique point (#/, # 2 ; , x) of the plane. Eeciprocally, to every 
point x 1 will correspond a unique point x, provided the determinant 
of the transformation 

A 

^31 ^82 a X 

is not 0. For we may then solve equations (1) for the ratios x^ix^n 
in terms of a?/: x : x% as follows : 



(2) />', 

p'x, 



188 COORDINATE SYSTEMS [CHAP.VII 

here the coefficients A tJ are the cofactors of the elements a t) respec- 
tively in the determinant A. 

Further, equations (1) transform every line in the plane into a 
unique line. In fact, the points x satisfying the equation 

UjJL\ + M 2 # 3 + tt a # 8 = 

are, by reference to equations (2), transformed into points x 1 satisfy- 
ing the equation 

(A u ii 1 + A^IS + A^ x[ 4- (A^ + A^u^ + A^u t ) ^ 

+ (A^U! + A^u s + A 5Q u s ) 3 ' = 0, 

which is the equation of a line. If the coordinates of this new line be 
denoted by [/, u^ u\ 9 we clearly have the following relations between 
the coordinates [K V u^ w s ] of any line and the coordinates [/, u^, u\ 
of the line into which it is transformed by (1): 



(3) 



We have seen thus far that (1) represents a collineation in the plane 
in point coordinates. The equations (3) represent the same collineation 
in line coordinates. 

It is readily seen, finally, that this collineation is protective. For 
this purpose it is only necessary to show that it transforms any 
pencil of lines into a projective pencil of lines. But it is clear that if 
m = and n are the equations of any two lines, and if (1) trans- 
forms them respectively into the lines whose equations are m 1 = 
and ?2/=0, any line ??i + Xft=0 is transformed into m f +\n f Q, 
and the correspondence thus established between the lines of the 
pencils has been shown to be projective (Theorem 6). 

Having shown that every transformation (1) represents a projective 
collraeation, we will now show conversely that every projective 
collineation in a plane may be represented by equations of the form 
(1). To this end we recall that every such collineation is completely 
determined as soon as the homologous elements of any complete 
quadrangle are assigned (Theorem 18, Chap. IV). If we can show 
that likewise there is one and only one transformation of the form 
(1) changing a given quadrangle into a given quadrangle, it will 
follow that, since the linear transformation is a projective collineation, 
it is the given projective collineation. 



67] LINEAR TRANSFORiEATIOXS 189 

Given any protective collineation in a plane, let the fundamental 
points (0, 0, 1), (0, 1, 0), (1, 0, 0), and (1, 1, 1) of the plane (which 
form a quadrangle) be transformed respectively into the points 
A - (a v a v aj, B = (b v 6 2 , J s ), C = (c v c z , ), and I) = (d v tl, tl), form- 
ing a quadrangle. Suppose, now, we seek to determine the coefficients 
of a transformation (1) so as to effect the correspondences just indi- 
cated. Clearly, if (0, 0, 1) is to be transformed into (a v a, ,), we 

must have . 

a ls =\a v r/ 23 = X 2 , a 3G =X s , 

X being an arbitrary factor of proportionality, the value (=?= 0) of which 
we may choose at pleasure. Similarly, we obtain 



Since, by hypothesis, the three points A, B, C are not collinear, it 
follows from these equations and the condition of Theorem 4, Cor. 2, 
that the determinant A of a transformation determined in this way 
is not 0. Substituting the values thus obtained in (1), it is seen that 
if the point (1, 1, 1) is to be transformed into (d v d^ d s ), the following 
relations must hold : 

/w7 1 =<r 

pel* = c 



Placing p = 1 and solving this system of equations for v 9 p, \ we 
obtain the coefficients a tj of the transformation. This solution is 
unique, since the determinant of the system is not zero. Moreover, 
none of the values X, p, v will be ; for the supposition that v = 0, 
for example, would imply the vanishing of the determinant 

d l \ 



which in turn would imply that the three points D, B, A are collinear, 
contrary to the hypothesis that the four points A, B> C 9 D form a 
complete quadrangle. 

Collecting the results of this section, we have 

THEOREM 8. Any protective colKneatwn in the plane may be repre- 
sented in point coordinates Tyy equations of form (1) or in line coordi- 
nates ly equations of form (3), and in each case the determinant of 



190 COOBDINATE SYSTEMS [CHAP.VII 

the transformation is different from ; conversely, any transforma- 
tion of one of these forms in which the determinant is different from 
represents a protective collineation in the plane. 

COROLLARY 1. In nonhomogeneous point coordinates the equations of 
a projectile collineation are 



COROLLARY 2. If the singular line of the system of nonhomogeneous 
point coordinates is transformed into itself, these equations can 5e 
written x ! = 



a* 6, 



=0. 



68. Collineations between two different planes. The analytic form 
of a collineation between two different planes is now readily derived. 
Let the two planes be a and /3, and let a system of coordinates be 
established in each, the point coordinates in a being (x v % 2 , # 3 ) and 
the point coordinates in y8 being (y v y z , y s ). Further, let the isomor- 
phism between the number systems in the two planes be established 
in such a way that the correspondence established by the equations 

Vi** *v y =a fc y.^afe 

is protective. It then follows, by an argument (cf. 59, p. 166), 
which need not be repeated here, that any collineation between the 
two planes may be obtained as the resultant of a projectivity in the 
plane a, which transforms a point X, say, into a point X\ and the pro- 
jectivity r = X f between the two planes. The analytic form of any 
protective collineation letween the two planes is therefore : 



with the determinant A of the coefficients 'different from 0. And, con- 
versely, every such transformation in which A = represents a projec- 
tive collineation letween the two planes. 

69. Nonhomogeneous coordinates in space. Point coordinates in 
space are introduced in a way entirely analogous to that used for the 
introduction of point coordinates in the plane. Choose a tetrahedron 
of reference OUVW and label the vertices = 0^= O y = 0,,, Z7= oo a , 



COORDINATES 1ST SPACE 



191 



F= oo y , JT= ex), (fig. 84) ; and on the lines O^oo,, O^oo,, O s co 2 , called 
respectively the x-axis, the y-axis, the 2-axis, establish three scales by 
choosing the points 1 X9 l y , 1 2 . The planes Ooo a co y , Ooo x co g9 Ott y ao a are 
called the xy-plane, xz-plane, yz-plane respectively. The point is 
called the origin. If P is any point not on the plane co x oo y oo z> which 
is called the singular plane of the coordinate system, the plane 
P 00^003 meets the #-axis in a point whose nonhomogeneons coordinate 
in the scale (0^, 1^, co x ) we call a. Similarly, let the plane 
meet the 2/-axis in a point 
whose nonhomogeneous 
coordinate in the scale 
(0,, l y) co y ) is I ; and let 
the plane Poo x co y meet the 
z-axis in a point whose 
nonhomogeneous coordi- 
nate in the scale (O a , 1 Z9 <x> z ) 
is c. The numbers a, &, c 
are then the nonhomo- 
geneous #-, y- 9 and z-coor- 
dinates of the point P. 
Conversely, any three 
numbers a 9 1 9 c determine 
three points A 9 B 9 C on 
the x- 9 y-> and #-axes respectively, and the three planes AaOyOo^ B<x> x <x> a , 
Cco x <x> y meet in a point P whose coordinates are a 9 6, c. Thus every 
point not on the singular plane of the coordinate system determines 
and is determined by three coordinates. The point P is then repre- 
sented by the symbol (a 9 &, c). 

The dual process gives rise to the coordinates of a plane. Point 
and plane coordinates may then be put into a convenient relation, as 
was done in the case of point and line coordinates in the plane, thus 
giving rise to a system of simultaneous point and plane coordinates 
in space. We will describe the system of plane coordinates with 
reference to this relation. Given the system of nonhomogeneous point 
coordinates described above, establish in each of the pencils of planes 
on the lines VW 9 UW, UV a scale by choosing the plane UVW as 
the zero plane M = O r = O w in each of the scales, and letting the planes 
VW 9 UW 9 UV be the planes oo tt , oo , oo^ respectively. In the ^scale 




FIG. 84 



192 COORDINATE SYSTEMS [CHAP.VH 

let that plane through. 7W be the plane l u , which meets the aj-axis 
in the point l a . Similarly, let the plane l v meet the ?/-axis in the 
point l y ; and let the plane l w meet the z-axis in the point !. 
The swscale, inscale, and w-scale being now completely determined, 
any plane TT not on the point (which is called the singular point 
of this system of plane coordinates) meets the #-, y-, and #-axes in 
three points L, M> N which determine in the u-, v-, and ^-scales planes 
whose coordinates, let us say, are l } m, n. These three numbers are 
called the nonhomogeneous plane coordinates of TT. They completely 
determine and are completely determined by the plane TT. The plane 
TT is then denoted by the symbol [I, m, n]. 

In this system of coordinates it is now readily seen that the con- 
dition that the point (a, 5, c) be on the plane \l 9 m,ri\ is that the relation 
la + mb + nc + l = Q le satisfied. It follows readily, as in the planar 
case, that the plane |7, m> n] meets the #-, y-, and #-axes in points 
whose coordinates on these axes are 1/, 1/m, and 1/n respec- 
tively.* In deriving the above condition we will suppose that the 
plane TT = |7, m 9 n] does not contain two of the points U, F, IF, leav- 
ing the other case as an exercise for the reader. Suppose, then, that 
U= co x and F= ao y are not on TT. By projecting the ys-plane with 
U as center upon the plane TT, and then projecting TT with F as center 
on the ##-plane, we obtain the following perspectivities : 



where (x, y, z) represents any point on ?r. The product of these two 
perspectivities is a projectivity between the y^-plane and the ##-plane, 
by which the singular line of the former is transformed into the sin- 
gular line of the latter. Denoting the ^-coordinate of points in the 
2/2-plane by z f , this projectivity is represented (according to Theorem 
8, Cor. 2, and 68) by relations of the form 



We proceed to determine the coefficients a v l v c v The point of 
intersection of TT with the y-axis is (0, 1/m, 0), and is clearly 

* This statement remains valid even if one or two of the numbers Z, m, n are 
zero (they cannot all be zero unless the plane in question is the singular plane 
which we exclude from consideration), provided the negative reciprocal of be 
denoted by the symbol o>. 



09] COOBDINATES DT SPACE 193 

transformed by the projectivity in question into the point (0, 0, 0). 
Hence (1) gives 

CI= -,IT 

The point of intersection of TT with the z-axis is, if w=0, (0, 0, 1/n) 
and is transformed into itself. Hence (1) gives 



n m 

7 M 

or 5. = --- 

m 

If n = 0, we have at once \ = 0. 



Finally, the point of intersection of TT with the #-axis is ( 1/7, 0, 0), 
and the transform of the point (0, 0, 0). Hence we have 



or a, = --- 

1 m 

Hence (1) becomes y = -- x -- z -- > 

m m m 

a relation which must be satisfied by the coordinates (x, y, z) of any 
point on TT. This relation is equivalent to 

Ix + my 4- nz -j- 1 = 0. 
Hence (a, 6, c) is on [7, m,n], if 

(2) Za + m& + nc + 1 = 0. 

Conversely, if (2) is satisfied by a point (a, 6, c), the point (0, 5, e) = P 
is transformed by the projectivity above into (#, 0, c) = $, and hence 
the lines JP 27 and Q V which meet in (, &, c) meet on TT. 

DEFINITION. An equation which DEFINITION. An equation which 
is satisfied by all the points (x, y, ) is satisfied by all the planes [u,v,w] 
of a plane and by no other points on a point and by no other planes 
is called the point equation of the is called the plane equation of the 
plane. point. 

The result of the preceding discussion may then be stated as follows : 
THEOREM 9. The point equation THEOREM 9'. The plane equation 
of the plane [I, m, n] is of the point (a, I, c) is 

= Q. au + bv + cw + 1 = 0. 



194 COORDINATE SYSTEMS [OHAP.VII 

70. Homogeneous coordinates in space. Assign to the vertices 0, U, 
V, W of any tetrahedron of reference the symbols (0, 0, 0, 1), (1, 0, 0, 0), 
(0, 1, 0, 0), (0, 0, 1, 0) respectively, and assign to any fifth point T 
not on a face of this tetrahedron the symbol (1, 1, 1, 1). The five 
points 0, U 9 V, W, T are called the frame of reference of the system 
of homogeneous coordinates now to be described. The four lines join- 
ing T to the points 0, U, V, W meet the opposite faces in four points, 
which we denote respectively by (1, 1, 1, 0), (0, 1, 1, 1), (1, 0, 1, 1), 
(1, 1, 0, 1). The planar four-point (0, 0, 0, 1), (0, 0, 1, 0), (0, 1, 0, 0), 
(0, 1, 1, 1) we regard as the frame of reference (0, 0, 1), (0, 1, TO), 
(1, 0, 0), (1, 1, 1) of a system of homogeneous coordinates in the plane. 
To any point in this plane we assign the coordinates (0, # 2 , # 3 , # 4 ), if 
its coordinates in the planar system just indicated are (# 2 , # 3 , # 4 ). In 
like manner, to the points of the other three faces of the tetrahedron of 
reference we assign coordinates of the forms (x v 0, # 3 , # 4 ), (x v x 2t 0, # 4 ), 
and (x v x z> x s , 0). The coordinates of the points in the faces opposite 
the vertices (1, 0, 0, 0), (0, 1, 0, 0), (0, 0, 1, 0), (0, 0, 0, 1) satisfy respec- 
tively the equations x^ = 0, # 2 = 0, x z = 0, x = 0. 

To the points of each edge of the tetrahedron of reference a notation 
has been assigned corresponding to each of the two faces which meet 
in the edge. Consider, for example, the line of intersection of the 
planes x l =Q and # 2 = 0. Eegarding this edge as a line of x t =^ 0, the 
coordinate system on the edge has as its fundamental points (0, 0, 1, 0), 
(0, 0, 0, 1), (0, 0, 1, 1). The first two of these are vertices of the tetra- 
hedron of reference, and the third is the trace of the line joining 
(0, 1, 0, 0) to (0, 1, 1, 1). On the other hand, regarding this edge as a 
line of # 2 =0, the coordinate system has the vertices (0, 0, 1, 0) and 
(0, 0, 0, 1) as two fundamental points, and has as (0, 0, 1, 1) the trace 
of the line joining (1, 0, 0, 0) to (1, 0, 1, 1). But by construction the 
plane (0, 1, 0, 0)(1, 0, 0, 0)(1, 1, 1, 1) contains both (0, 1, 1, 1) and 
(1, 0, 1, 1), so that the two determinations of (0, 0, 1, 1) are identical. 
Hence the symbols denoting points in the two planes x l = and 
# 2 =0 are identical along their line of intersection. A similar result 
holds for the other edges of the tetrahedron of reference. 

THEOREM 10, DEFINITION. If P is any point not on a face of the 
tetrahedron of reference, there exist four numbers x v # 2 , # 3 , x^ all 
different from zero, suck that the projections of P from the four vertices 
(1, 0, 0, 0), (0, 1, 0, 0), (0, 0, 1, 0), (0, 0, 0, 1) respectively upon their 



70] COORDINATES IN SPACE 195 

opposite faces are (0, <e a , x v xj, (x v 0, X B) a?J, (^ re s , 0, # 4 ), (x v x 2 , x z , 0). 
l^ese four numbers are called the homogeneous coordinates of P and 
P is denoted ly (x v x 2 , X B , a?J. Any ordered set of four members, not 
all zero, determine uniquely a point in space whose coordinates they are. 

Proof. The line joining P to (1, 0, 0, 0) meets the opposite face in 
a point (0, x z , x 2 , # 4 ), which is not an edge of the tetrahedron of refer- 
ence, and such therefore that none of the numbers # 2 , X B , x 4 is zero. 
Likewise the line joining P to (0, 1, 0, 0) meets the opposite face in 
a point (as/, 0, # 8 ', # 4 ), such that none of the numbers x[, # 8 ', x is zero. 
But the plane P(l, 0, 0, 0) (0, 1, 0, 0) meets x l = in the line joining 
(0, 1, 0, 0) to (0, x 2 , # 3 , # 4 ), and meets x 2 = in the line joining 
(1, 0, 0, 0) to (as/, 0, # 3 ', xl). By the analytic methods already devel- 
oped for the plane, the first of these lines meets the edge common 
to x 1 = and x 2 = in the point (0, 0, a? 8 , # 4 ), and the second meets 
it in the point (0, 0, #/, xl). But the points (0, 0, a? 8 , # 4 ) and 
(0, 0, # 8 ', x^) are identical, and hence, by the preceding paragraph, we 
have # s /# 4 = %l/%l* Hence, if we place x l = x^xjxl, the point 
(as/, 0, x, xl) is identical with (x v 0, x 9 , x). The line joining P to 
(0, 0, 1, 0) meets the face # 8 = in a point (x? 9 x, 0, x^. By the 
same reasoning as that above it follows that we have x/xl r = xjx 
and %2/xl 1 ttj%# so that the point (as/', x^ 0, x^) is identical with 
(x v # 2 , 0, a? 4 ). Finally, the line joining P to (0, 0, 0, 1) meets the face 
# 4 = in a point which a like argument shows to be (x v x 2 , X B , 0). 

Conversely, if the coordinates (x v # 2 , a? 3 , # 4 ) are given, and one of 
them is zero, they determine a point on a face of the tetrahedron 
of reference. If none of them is zero, the lines joining (1, 0, 0, 0) 
to (0, x z9 x 39 # 4 ) and (0, 1, 0, 0) to (x v 0, # 8 , # 4 ) are in the plane 
(1, 0, 0, 0) (0, 1, 0, 0) (0, 0, x s) x 4 ), and hence meet in a point which, 
by the reasoning above, has the coordinates (x v x z , # 8 , # 4 ). 

COEOLLAEY. The notations (x^x^x^x^) and (kx v Jcx 2 , Jcx B , kx^) 
denote the same point for any value of 7c not equal to zero. 

Homogeneous plane coordinates in space arise by the dual of the 
above process. The four faces of a tetrahedron of reference are denoted 
respectively by [1, 0, 0, 0], t [0, 1, 0, 0], [0, 0, 1, 0], and [0, 0, 0, 1]. 
These, together with any plane [1, 1, 1, 1] not on a vertex of the 
tetrahedron, form the frame of reference. The four lines of inter- 
section of the plane [1, 1, 1, 1] with the other four planes in the order 



196 COOKDINATE SYSTEMS [CHAP.VII 

above are projected from the opposite vertices by planes which are 
denoted by [0,1,1,1], [1,0,1,1], [1,1,0,1], [1,1,1,0] respectively. 
The four planes [0, 1, 0, 0], [0, 0, 1, 0], [0, 0, 0, 1], and [0, 1, 1, 1] form, 
if the first in each of these symbols is suppressed, the frame of 
reference of a system of homogeneous coordinates in a bundle (the 
space dual of such a system in a plane). The center of this bundle 
is the vertex of the tetrahedron of reference opposite to [1, 0, 0, 0]. 
To any plane on this point is assigned the notation [0, u v u# ^J, if 
its coordinates in the bundle are [u z , u s> wj. In like manner, to the 
planes on the other vertices are assigned coordinates of the forms 
[u v 0, u s , u], [u v u v 0, ttj, [u v u 2 , u 3 , 0]. The space dual of the last 
theorem then gives : 

THEOKEM 10'. DEFINITION. Ifir is any plane not on a vertex of the 
tetrahedron of reference, there exist four numbers u v u^ u 9) u *> a ^ differ- 
ent from zero, such that the traces of IT on the four faces [1, 0, 0, 0], 
[0, 1, 0, 0], [0, 0, 1, 0], [0, 0, 0, 1] respectively are projected from the 
opposite vertices "by the planes [0, u^, u %9 wj, [u v 0, w s , wj, [u v u z) 0, % 4 ], 
\u v u z ,u 2 , 0], These four members are called the homogeneous coordinates 
0/7T, and TT is denoted ly [u v u z , u s , u\. Any ordered set of four num- 
lers, not all zero, determine uniquely a plane whose coordinates they are. 

By placing these systems of point and plane coordinates in a proper 
relation we may now readily derive the necessary and sufficient con- 
dition that a point (x v a? a , x s> a? 4 ) be on a plane \u v u%, u &) u^\. This 
condition will turn out to be 



We note first that in a system of point coordinates as described above 
the sis points (- 1, 1, 0, 0), (- 1, 0, 1, 0), (- 1, 0, 0, 1), (0, - 1, 1, 0), 
(0, 0, 1, 1), (0, 1, 0, 1) are coplanar, each being the harmonic con- 
jugate, with respect to two vertices of the tetrahedron of reference, of 
the point into which (1, 1, 1, 1) is projected by the line joining the 
other two vertices. The plane containing these is, in fact, the polar 
of (1, 1, 1, 1) with respect to the tetrahedron of reference (cf. Ez. 3, 
p. 47). Now choose 

as the plane [1, 0, 0, 0] the plane ^= 0, 
as the plane [0, 1, 0, 0] the plane # 2 = 0, 
as the plane [0, 0, 1, 0] the plane x s = 0, 
as the plane [0, 0, 0, 1] the plane x 4 = 0, 



70] COORDINATES IN SPACE 197 

as the plane [1, 1, 1, 1] the plane containing the points ( 1, 1, 0, 0), 
(-1,0,1,0), (-1,0,0,1)- 

With this choice of coordinates the planes [1, 0, 0, 0], [0, 1, 0, 0], 
[0, 0, 1, 0], and [1, 1, 1, 0] through the vertex F 4 , say, whose point 
coordinates are (0, 0, 0, 1), meet the opposite face x = in lines 
whose equations in that plane are 



Hence the first three coordinates of any plane [u v ^ 2 , u v 0] on F 4 
are the line coordinates of its trace on # 4 = 0, in a system so chosen 
that the point (SG V x 2 , # 3 ) is on the line [u v u S9 u s ] if and only if the 
relation u^ + U Q X & + u s x s = is satisfied. Hence a point (x v x z , x v 0) 
lies on a plane [u l9 u 2 , u^ 0] if and only if we have ufa + UyfB 9 + 
^ 3 # s = 0. But any point (sc v # 2 , # 8 , # 4 ) on the plane [u v u^ u^ 0] has, 
by definition, its first three coordinates identical with the first three 
coordinates of some point oa the trace of this plane with the plane 
x = 0. Hence any point (x 19 x 2 , x 3 , x 4 ) on [u v u 2 , u s , 0] satisfies the 
condition ufa + u^x z + u z x s + ufa*= 0. Applying this reasoning to 
each of the four vertices of the tetrahedron of reference and dualizing, 
we find that if one coordinate of \u v u 2 , ^ 3 , %J is zero, the necessary 
and sufficient condition that this plane contain a point (x v % 2 , # 8 , x 4 ) 
is that the relation 

U^ + U& + Ufa + Ufa = 

le satisfied; and if one coordinate of (x v x z , X B , x 4 ) is zero, the neces- 
sary and sufficient condition that this point be on the plane [^ v u &9 u^u^\ 
is likewise that tJie relation just given "be satisfied. 

Confining our attention now to points and planes no coordinate of 
which is zero, let xjx^x 9 xjx^y^ a? 8 /# 4 = 2, and let uju^% 9 
w a /w 4 =0, u z f%^w. Since x 9 y, z are the ratios of homogeneous 
coordinates on the lines # 2 = # 3 = 0, x^ = x 3 = 0, and x = x% = respec- 
tively, they satisfy the definition of nonhomogeneous coordinates 
given in 69. And since the homogeneous coordinates have been 
so chosen that the plane (u v % z , u^ u^ meets the line x z = x z = in 
the point ( u 9 0, 0, u^ = ( 1/u, 0, 0, 1), it follows that u, v } w are 
nonhomogeneous plane coordinates so chosen that a point (x, y, z) 9 
none of whose coordinates is zero, is on a plane [%, v> w\ none of 
whose coordinates is zero, if and only if we have (Theorem 9) 

ux + *oy -I- ^^ + 1 = ; 



198 



COORDINATE SYSTEMS 



[CHAP. VII 



that is, if and only if we have 

U& 4- Ufa + u s x 3 + Ufa = 0. 

This completes for all cases the proof of 

THEOREM 11. The necessary and sufficient condition that a point 
(Xy x 2i X B) # 4 ) le on a plane [u v u^ u s> u^ is that the relation 

Ufa + Ufa + Ufa + Ufa = 
be satisfied. 

By methods analogous to those employed in 64 and 65 we may 
now derive the results of Exs. 1-8 below. 



EXERCISES 

1. The equation of the plane through the three points A = (a v a 2 , # 3 , a 4 ), 
B = (1 19 b z , 3 , Z> 4 ), C = (c x , c 2 , c 3 , c 4 ) is 

#1 2T 5.' a 



5 1 & 2 5 3 & 4 



= 0. 



Dualize. 

2. The necessary and sufficient condition that four points -4, B, C 7 D be 
coplanar is the vanishing of the determinant 



b l 5 3 b z 5 4 



3. The necessary and sufficient condition that three points A, B, C be 
collinear is the vanishing of the three-rowed determinants of the matrix 



4. Any point of a pencil of points containing A and -B may be represented by 



P = 



5. Any plane of a pencil of planes containing m = [%, m 2 , w 8 , wi 4 ] and 
n = [n lf n 2 , TZ S , n 4 ] may be represented by 



6. Any projectivity between two one-dimensional primitive forms (of points 
or planes) in space is expressed by a relation between their parameters X, p, 
of the form 



If the base elements of the pencil are homologous, this relation reduces to 



70,71] LINEAE TRAITSFOBMATION 1 199 

7. If \ v X 21 \ 3 , \ 4 are the parameters of four points or planes of a pencil, 
tlieir cross ratio is 



8. Any point (plane) of a plane of points (bundle of planes) containing 
the noncollinear points A, B, C (planes a, /?, y) may be represented by 

P = (A^ + X^ + Ag^, A.ja 2 4- Xj&j -f X 3 c 2 , X^g + X 2 Z> 8 -f A 8 <? 8 , A.^ + X^ -f \ 3 c 4 ). 

9. Derive the equation of the polar plane of any point with regard to the 
tetrahedron of reference. 

10. Derive the equation of a cone. 

*11. Derive nonhomogeneous and homogeneous systems of coordinates in 
a space of four dimensions. 

71. Linear transformations in space. The properties of a linear 
transformation in space 



are similar to those found in 68 for the linear transformations in a 
plane. If the determinant of the transformation 



is different from zero, the transformation (1) will have a unique in- 
verse, viz.: 

< ; 



where the coefi&cients ^t y . are the cof actors of the elements a y . respec- 
tively in the determinant A. 

The transformation is evidently a collineation, as it transforms the 

^ G %!#! + ^ 2 # 2 + w 8 # s + i6 4 a; 4 = 

into the plane 



2 4- A BB u s -h 



200 COOKDINATE SYSTEMS [CHAP, vn 

Hence the collineation (1) produces on the planes of space the trans- 
formation 



( ' a-iol 

aru[ = A^Ut 4- 

To show that the transformation (1) is projective consider any 
pencil of planes 



In accordance with (2) this pencil is transformed into a pencil of the 
form 

(a^x 1 + a fa + ax B + a 4 \) + X (6^ + 1 fa + I fa + 1^ = 0, 

and these two pencils of planes are projective (Ex. 6, p. 198). 

Finally, as in 67, we see that there is one and only one trans- 
formation (1) changing the points (0, 0, 0, 1), (0, 0, 1, 0), (0, 1, 0, 0), 
(1, 0, 0, 0), and (1, 1, 1, 1) into the vertices of an arbitrary complete 
five-point in space. Since this transformation is a projective collinea- 
tion, and since there is only one projective collineation transforming 
one five-point into another (Theorem 19, Chap. IV), it follows that 
every projective collineation in space may be represented by a linear 
transformation of the form (1). This gives 

THEOREM 12. Any projective collineation of space may le repre- 
sented in point coordinates ly equations of the form (1), or in plane 
coordinates by equations of the form (3). In each case the determinant 
of the transformation is different from zero. Conversely, any trans- 
formation of this form in which the determinant is different from zero 
represents a protective collineation of space. 

COEOLLAET 1. In nonhomogeneous point coordinates a projective 
collineation is represented ly the linear fractional equations 



. 



in which the determinant A is different from zero. 



71,72] FINITE SPACES 201 

COROLLARY 2. If the singular plane of the nonhomogeneous system 
is transformed into itself, these equations reduce to 
x f = a 



72. Finite spaces. It will be of interest at this point to emphasize 
again the generality of the theory which we are developing. Since 
all the developments of this chapter are on the basis of Assumptions 
A, E, and P only, and since these assumptions imply nothing regard- 
ing the number system of points on a line, except that it be commu- 
tative, it follows that we may assume the points of a line, or, indeed, 
the elements of any one-dimensional form, to be in one-to-one recip- 
rocal correspondence with the elements of any commutative number 
system. We may, moreover, study our geometry entirely by analytic 
methods. From this point of view, any point in a plane is simply a 
set of three numbers (x v # 2 , # 3 ), it being understood that the sets 
(x v x 2> X B ) and (Jcx v Jcx 2 , Jcx s ) are equivalent for all values of A in the 
number system, provided k is different from 0. Any line in the plane 
is the set of all these points which satisfy any equation of the form 
2^ + u z x 2 + u 8 x B = 0, the set of all lines being obtained by giving 
the coefficients (coordinates) [u v u^ u$] all possible values in the 
number system (except [0, 0, 0]), with the obvious agreement that 
[u^u^Uz] and [ku 19 ku s , JcUi] represent the same line (^=^0). By 
letting the number system consist of all ordinary rational numbers, 
or all ordinary real numbers, or all ordinary complex numbers, we 
obtain respectively the analytic form of ordinary rational, or real, or 
complex projective geometry in the plane. All of our theory thus 
far applies equally to each of these geometries as well as to the 
geometry obtained by choosing as our number system any field 
whatever (any ordinary algebraic field, for example). 

In particular, we may also choose a finite field, i.e. one which con- 
tains only a finite number of elements. The simplest of these are 
the modular fields, the modulus being any prime number^.* If we 

* A modular field with modulus p is obtained as follows : Two integers n, n f 
(positive, negative, or zero) are said to be congruent modulo p, written n=n', mod.p, 
if the difference n n' is divisible by p. Every integer is then congruent to one 
and only one of the numbers 0, 1, 2, , p 1. These numbers are taken as the 
elements of our field, and any number obtained from these ly addition, subtraction* 



202 COORDINATE SYSTEMS [CHAP.VH 

consider, for example, the case p = 2, our number system contains 
only the elements and 1. There are then seven points, which we 
will label A, B, C, D, E, F, G, as follows : A = (0, 0, 1), B = (0, 1, 0), 
<? = (!, 0, 0), D = (0, 1, 1), J0 = (l, 1, 0), ^ = (1, 1, 1), ff (1, 0, 1). 
The reader will readily verify that these seven points are arranged 
in lines according to the table 

A B C D F G 
B C D E F G A 
D E F G A B C, 

each column constituting a line. Tor example, the line a? 1 = clearly 
consists of the points (0, 0, 1) = A, (0, 1, 0) = B, and (0, 1, 1) = D, these 
"being the only points whose first coordinate is 0. We have labeled 
the points of this finite plane in such a way as to exhibit clearly its 
abstract identity with the system of triples used for illustrative pur- 
poses in the Introduction, 2.* 

EXERCISES 

1. Verify analytically that two sides of a complete quadrangle containing a 
diagonal point are harmonic with the other two diagonal points. 

2. Show analytically that if two projective pencils of lines in a plane have 
a self -corresponding line, they are perspective. (This is equivalent to Assump- 
tion P.) 

3. Show that the lines whose equations are x 1 + \x 2 = 0, x% 4- pjc 8 = 0, and 
#s +vx t = are concurrent if X/w = 1 ; and that they meet the opposite 
sides of the triangle of reference respectively in collinear points, if X/xv = 1. 

4. Find the equations of the lines joining (c v c 2 , c 8 ) to the four points 
(1, 1, 1), and determine the cross ratios of the pencil. 

and multiplication, if not equal to one of these elements, is replaced by the element 
to which it is congruent. The modular field with modulus 5, for example, consists of 
the elements 0, 1, 2, 3, 4, and we have as examples of addition, subtraction, and 
multiplication 14-3 = 4, 2 + 3 = (since 5 = 0, mod. 5), 1 - 4 = 2, 2 . 3 = 1, etc, 
Furthermore, if a, 5 are any two elements of this field (a^ 0), there is a unique 
element x determined by the congruence ax = 6, mod. p; this element is defined 
as the quotient b/a (For the proof of this proposition the reader may refer to any 
standard text on the theory of numbers.) In the example discussed we have, for 
example, 4/3 = 3. 

* For references and a further discussion of finite projective geometries see a 
paper by O. Veblen and W. H. Bussey, Finite Projective Geometries, Transactions 
of the American Mathematical Society, Vol. VII (1906), pp. 241-259. Also a sub- 
sequent paper by 0. Veblen, Collineations in a Finite Projective Geometry, Trans- 
actions of the American Mathematical Society, Vol. VIII (1907), pp. 266-268. 



72] EXERCISES 203 

5. Show that the throw of lines determined on (c l9 c 2 , c s ) by the four 
points (1, 1, 1) is projective with (equal to) the throw of lines determined 
on (b 19 & 3 , & 3 ) by the points (a l9 a 2 , 3 ), if the following relations hold: 

a 3 = 0, 



ajajb? + a^bj + a^ajb} = 0, 

and that the sis cross ratios are a 3 /a 3 , a 9 /a 19 a^/a^ , a s /a, - aja^ 
2 /i (C. A. Scott, Mod. Anal. Geom., p. 50). 

6. Write the equations of transformation for the five types of planar col- 
lineations described in 40, Chap. IV, choosing points of the triangle of 
reference as fixed points. 

7. Generalize Ex. 6 to space. 

8. Show that the set of values of the parameter X of the pencil of lines 
m + \n = is isomorphic with the scale determined in this pencil by the lines 
for which the fundamental lines are respectively the lines X = 0, 1, oo 

9. Show directly from the discussion of 61 that the points whose non- 
homogeneous coordinates x 9 y satisfy the equation y = x are on the line joining 
the origin to the point (1, 1). 

10. There is then established on this line a scale whose fundamental points 
are respectively the origin, the point (!?!) and the point in which the line meets 
the line fc . The lines joining any point P in the plane to the points oo y , 003. 
meet the line y = x in two points whose coordinates in the scale just determined 
are the nonhomogeneous coordinates of P, so that any point in the plane 
(not on loo) is represented by a pair of points on the line y = x. Hence, show 
that in general the points (a;, y) of any line in the plane determine on the 
line y = x a projectivity with a double point on Z ; and hence that the equa- 
tion of any such line is of the form y = ax -f- b. What lines are exceptions to 
this proposition ? 

11. Discuss the modular plane geometry in which the modulus is^? = 3 ; 
and by properly* labeling the points show that it is abstractly identical with 
the system of quadruples exhibited as System (2) on p. 6. 

12. Show in general that the modular projective plane with modulus p 
contains p* + p + 1 points and the same number of lines ; and that there are 
p 4- 1 points (lines) on every line (point). 

13. The diagonal points of a complete quadrangle in a modular plane pro- 
jective geometry are collinear if and only if p = 2. 

14 * Show that the points and lines of a modular plane all belong to the 
same net of rationality. Such a plane is then properly projective without the 
use of Assumption P. 

15. Show how to construct a modular three-space. If the modulus is 2, 
show that its points may be labeled 0, 1, . . . , 14 in such a way that the 
planes are the sets of seven obtained by cyclic permutation from the set 
1 4 6 11 12 13 (i.e. 1 2 5 7 12 13 14, etc.), and that the lines are ob- 
tained from the lines 014, 1 2 8, 5 10 by cyclic permutations. (For a 



204 GOOEDINATE SYSTEMS [CHAP.VII 

study of this space, see G. M, Conwell, Annals of Mathematics, Vol. 11 
(1910), p. 60.) 

16. Show that the ten diagonal points of a complete five-point in space 
(0, 0, 0, 1), (0, 0, 1, 0), (0, 1, 0, 0), (1, 0, 0, 0), (1, 1, 1, 1) are given by the 
remaining sets of coordinates in which occur only the digits and 1. 

17. Show that the ten diagonal points in Ex. 16 determine in all 45 planes, 
of which each of a set of 25 contains four diagonal points, while each of the 
remaining 20 contains only three diagonal points. Through any diagonal 
point pass 16 of these planes. The diagonal lines, i.e. lines joining two 
diagonal points, are of two kinds : through each of the diagonal lines of the 
first kind pass five diagonal planes ; through each line of the second kind pass 
four diagonal planes. 

18. Show how the results of Ex. 17 are modified in a modular space with 
modulus 2 ; with modulus 3. Show that in the modular space with modulus 
5 the results of Ex. 17 hold without modification. 

* 19. Derive homogeneous and nonhomogeneous coordinate systems for 
a space of n dimensions, and establish the formulas for an n-dimensional 
projective collineation. 



CHAPTER VIII 

PROJECTTVITIES IN ONE-DIMENSIONAL FORMS* 

73. Characteristic throw and cross ratio. 

THEOREM 1. If M 9 N are double points of a projectivity on a line, 
and AA f , BB f are any two pairs of homologous points (i.e. if 
MNAB -^ MNA'ff), then MNAA -^ MNBB 1 . 

Proof. Let S 9 S f be any two distinct points on a line through 
M (fig. 85), and let the lines SA and S f A r meet in A" 9 and SB and 

S 




N 



S'B' meet in B". The points A", B", Nsxe then collinear (Theorem 23, 
Chap. IV). If the line A n B" meets SS' in a point Q> we have 

A" B" 

MNAA f = MQ8S 1 = MNBB f . 

A A 

This proves the theorem, which may also be stated as follows : 

The throws consisting of the pair of double points in a given order 
and any pair of homologous points are all equal. 

DEFINITION. The throw "T(MN 9 AA 1 ), consisting of the double points 
and a pair of homologous points of a projectivity, is called the charac- 
teristic throw of the projectivity ; and the cross ratio of this throw 
is called the characteristic cross ratio of the projectivity. | 

* All the developments of this chapter are on the basis of Assumptions A, E, P, H . 

t Since the double points enter symmetrically, the throws T (MN^ AA 7 ) and 
T(JO(f, AA") may be used equally well for the characteristic throw. The corre- 
sponding cross ratios R (MN, A A') and B (NM, AA*) are reciprocals of each other 
(cf. Theorem 13, Cor. 3, Chap. VI). 

205 



206 ONE-DIMENSIONAL PROJECTIVITIES [CHAP.VIII 

COROLLARY 1. A projectivity on a line with two given distinct 
double points is uniquely determined ly its characteristic throw or 
cross ratio. 

COROLLARY 2. The characteristic cross ratio of any involution with 
double points is 1. 

This follows directly from Theorem 27, Cor. 1, Chap. IV, and 
Theorem 13, Cor. 2, Chap. VI. 

If m, n are nonhomogeneous coordinates of the double points, and 
Jc is the characteristic cross ratio of a projectivity on a line, we have 

x t m f x m_ , 
x 1 n ' x n 

for every pair of homologous points x, x f . This is the analytic expres- 
sion of the above theorem, and leads at once to the following analytic 
expression for a projectivity on a line with two distinct double points 
m, n : 

COROLLARY 3. Any projectivity on a line with two distinct double 
points m, n may be represented ~by the equation 



x'n x n 
x r , x being any pair of homologous points. 

For when cleared of fractions this is a bilinear equation in x j , x 
which obviously has m, n as roots. Moreover, since any projectivity 
with two given distinct double points is uniquely determined by one 
additional pair of homologous elements, it follows that any projec- 
tivity of the kind described can be so represented, in view of the fact 
that one such pair of homologous points will always determine the 
multiplier k These considerations offer an analytic proof of Theo- 
rem 1, for the case when the double points Jf, 2f are distinct. 

It is to be noted, however, that the proof of Theorem 1 applies 
equally well when the points M, N coincide, and leads to the follow- 
ing theorem : 

THEOREM 2. If in a paralolic projectivity with double point M the 
points A A 1 and BB f are two pairs of homologous points , the parabolic 
projectivity with double point M which puts A into B also puts A 1 
into B'. 

COROLLARY. The characteristic cross ratio of any parabolic projcc- 
timty is unity. 



73] CHARACTERISTIC THROW 207 

The characteristic cross ratio together with the double point is 
therefore not sufficient to characterize a parabolic projectivity com- 
pletely. Also, the analytic form for a projectivity with double points 
m, n, obtained above, breaks down when m = n. We may, however, 
readily derive a characteristic property of parabolic projectivities, 
from which will follow an analytic form for these projectivities. 

THEOREM 3. If a parabolic projectivity with doiible point M trans- 
forms a point A into A! and A f into A fl , the pair of points A, A rf is 
harmonic with the pair A f M; i.e. we have H (MA r , AA ff ). 

Proof. By Theorem 23, Chap. IV, we have Q(MAJL*, MA" A'). 

Analytically, if the coordinates of M 9 A, A r , A rf are m, %, x r , x n 
respectively, we have, by Theorem 13, Cor. 4, Chap. VI, 

2 = 1 1 

x ffit x " in x - yn 
This gives 



x'm x m x n m x'm 

which shows that if each member of this equation be placed equal to 
t, the relation 

(1) 



' m x 



is satisfied by every pair of homologous points of the sequence obtained 
by applying the projectivity successively to the points A, A r , A", 
It is, however, readily seen that this relation is satisfied by every pair 
of homologous points on the line. Tor relation (1), when cleared of 
fractions, clearly gives a bilinear form in x f and x, and is therefore a 
projectivity; and this projectivity clearly has only the one double 
point m. It therefore represents a parabolic projectivity with the 
double point m y and must represent the projectivity in question, since 
the relation is satisfied by the coordinates of the pair of homologous 
points A, A\ which are sufficient with the double point to determine 
the projectivity. 

We have then : 

COROLLABY 1. Any parabolic projectivity with a double point, M, 
may be represented by the relation (1). 

DEFINITION. The number t is called the characteristic constant of 
the projectivity (1). 



208 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP. VIII 

COROLLARY 2. Conversely, if a projectivity with a double point 
M transforms a point A into A', and A* into A !r , such that we have 
H (MA 1 , -4-4"), the projectwity is parabolic. 

Proof. The double point M and the two pairs of homologous 
points AA r , A'A n are sufficient to determine the projectivity uniquely; 
and there is a parabolic projectivity satisfying the given conditions. 

74. Projective projectivities. Let TT be a projectivity on a line I, 
and let TT I be a projectivity transforming the points of I into the 
points of another or the same line V. The projectivity Ti^Tnrf 1 is then 
a projectivity on V. For Tr^ 1 transforms any point of V into a point 
of Z, TT transforms this point into another point of I, which in turn is 
transformed into a point of V by ir v Thus, to every point of V is made 
to correspond a unique point of l f , and this correspondence is projec- 
tive, since it is the product of projective correspondences. Clearly, 
also, the projectivity ir^ transforms any pair of homologous points of 
TT into a pair of homologous points of w^w-Trf 1 . 

DEFINITION. The projectivity Tj^Trrrf 1 is called the transform of V 
"by TT X ; two projectivities are said to be projectile or conjugate if one 
is a transform of the other by a projectivity. 

The question now arises as to the conditions under which two pro- 
jectivities are projective or conjugate. A necessary condition is evi- 
dent. If one of two conjugate projectivities has two distinct double 
points, the other must likewise have two distinct double points; if 
one has no double points, the other likewise can have no double points ; 
and if one is parabolic, the other must be parabolic. The further 
conditions are readily derivable in the case of two projectivities with 
distinct double points and in the case of two parabolic projectivities. 
They are stated in the two following theorems : 

THEOREM 4. Two projectivities each of which has two distinct double 
points are conjugate if and only if their characteristic throws are equal. 

Proof. The condition is necessary. For if TT, TT' are two conjugate 
projectivities, any projectivity T^ transforming TT into TT' transforms 
the double points M, N of TT into the double points M', N f of TT', and 
also transforms any pair of homologous points A, A l of TT into a pair 
of homologous points A 1 , A[ of ir 1 ; i.e. 

Tr^MNAA^ = M'N'A'Af. 
But this states that their characteristic throws are equal. 



74, 75] GKOUPS ON A LINE 209 

The condition is also sufficient ; for if it is satisfied, the projec- 
tivity TT^ defined by 



clearly transforms TT into TT'. 

COROLLARY. Any two involutions with doiible points. are conjugate. 
THEOREM 5. Any two parabolic projectivities are conjugate. 
Proof. Let the two parabolic projectivities be defined by 
ir(MMA) = MMA V and 7r r (M r M'A r ) = Jf'Jf^'. 
Then the projectivity ^ defined by 

ir^MAAJ = M'A'Af 

clearly transforms TT into TT'. 

Since the characteristic cross ratio of any parabolic projectivity is 
unity, the condition of Theorem 4 may also be regarded as holding 
for parabolic projectivities. 

75. Groups of projectivities on a line. DEFINITION. Two groups G 
and G ; of projectivities on a line are said to be conjugate if there 
exists a projectivity ir^ which transforms every projectivity of G into a 
projectivity of G', and conversely. We may then write 7r 1 G'7r~ ;l =G / ; 
and G' is said to be the transform of G ly nr. 

We have already seen (Theorem 8, Chap. Ill) that the set of all 
projectivities on a line form a group, which is called the general pro- 
jective group on the line. The following are important subgroups : 

1. TJie set of all projectivities leaving a given point of the line 
invariant. 

Any two groups of this type are conjugate. For any projectivity 
transforming the invariant point of one group into the invariant point 
of the other clearly transforms every projectivity of the one into 
some projectivity of the other. Analytically, if we choose % = oo as 
the invariant point of the group, the group consists of all projectivities 

of the form 

x r = ase + 'b. 

2. The set of all projectivities leaving two given distinct points 
invariant 

* Any two groups of this type are conjugate. Tor any projectivity 
transforming the two invariant points of the one into the invariant 
points of the other clearly transforms every projectivity of the one 



210 ONE-DIMENSIONAL PEOJECTIYITIES [CHAP.VIII 

into a projectivity of tlie other. Analytically, if x v x 2 are the two 
invariant points, the group consists of all projectivities of the form 

AU( ____ 

x f ~ 

The product of two such projectivities with multipliers k and Id is 
clearly given by 

x f x z 

This shows that any two projectivities of this group are commuta- 
tive. This result gives 

THEOREM 6. Any two projectivities which have two double points 
in common are commutative. 

This theorem is equivalent to the commutative law for multiplication. 
If the double points are the points and oo, the group consists of all projec- 
tivities of the form x f = ax. 

3. The set of all parabolic projectivities with a common double point 
In order to show that this set of projectivities is a group, it is only 
necessary to show that the product of two parabolic projectivities 
with the same double point is parabolic. This follows readily from 
the analytic representation. The set of projectivities above described 
consists of all transformations of the form 



^ l 

where x^ is the common double point (Theorem 3, Cor. 1). If 



___ 

/VM .. fy> f)i __ /y 

ifj " " IAJ, tfj ^^ tAj* 

are two projectivities of this set, the product of the first by the second 
is given by j j 

-. =: ' ~r & T" ^o, 

a/ x l x %i x 

which is clearly a projectivity of the set. It shows, moreover, that 
any two projectivities of this group are commutative. Whence 

THEOREM 7. Any two parabolic projectivities on a line with the 
same double point are commutative. 

This theorem is independent of Assumption P, although this assumption* 
is implied in the proof we have given. The theorem has already been proved 
without this assumption in Example 2, p. 70. 



75] GBOUPS ON A LINE 211 

Any two groups of this type are conjugate. For every projectivity 
transforming the double point of one group into the double point of 
the other transforms the one group into the other, since the projec- 
tive transform of a parabolic projectivity is parabolic. 

DEFINITION. Two subgroups of a group G are said to be conjugate 
under G if there exists a transformation of G which transforms one 
of the subgroups into the other. A subgroup of G is said to be self- 
conjugate or invariant under G if it is transformed into itself by 
every transformation of G; i.e. if every transformation in G trans- 
forms any transformation of the subgroup into another (or the same) 
transformation of the subgroup. 

We have seen that any two groups of any one of the three types 
are conjugate subgroups of the general projective group on the line. 
We may now give an example of a self-conjugate subgroup. 

Tlie set of all parabolic projectivities in a group of Type 1 alove is 
a self-conjugate subgroup of this group. It is clearly a subgroup, since 
it is a group of Type 3. And it is self-conjugate, since any conjugate 
of a parabolic projectivity is parabolic, and since every projectivity of 
the group leaves the common double point invariant. 

EXERCISES 

1. Write the eqtiations of all the projective transformations which permute 
among themselves (a) the points (0, 1), (1,0), (1, 1) ; (&) the points (0, 1), 
(1,0), (1,1), (a,&); (c) the points (0,1), (1,0), (1,1), (-1,1). What 
are the equations of the self-conjugate subgroup of the group of transforma- 
tions (a)? 

2. If a projectivity x r (ax + I}/ (ex -f d) having two distinct double ele- 
ments be written in the form of Cor. 3, Theorem 1, show that 



and that 

ad be 

3. If a parabolic projectivity x f = (ax + b)/(cx + d) be written in the form 
of Theorem 3, Cor. 1, show that x^ = (a d)/2 c, and t = 2 c/(a + ?). 

4. Show that a projectivity with distinct double points x v x z and charac- 
teristic cross ratio k can be written in the form 

1 



or 1 
a?! 1 1 
ar k 1 



212 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP.VIII 

5. Show that the parabolic projectivity of Theorem 3, Cor. 1, may be 
written in the form 



x 

i ^i 


1 
1 
1 




x 
1 

1 i 


1 
1 







6. If by means of a suitably chosen transformation of a group any of the 
elements transformed may be transformed into any other element, the group 
is said to be transitive. If by a suitably chosen transformation of a group any 
set of n distinct elements may be transformed into any other set of n distinct 
elements, and if this is not true for all sets of n + 1 distinct elements, the 
group is said to be n-ply transitive. Show that the general projective group on 
a line is triply transitive, and that of the subgroups listed in 75 the first 
is doubly transitive and the other two are simply transitive. 

7. If two projecti vities on a line, each having two distinct double points, 
have one double point in common, the characteristic cross ratio of their prod- 
uct is equal to the product of their characteristic cross ratios. 

76. Projective transformations between conies. "We have consid- 
ered hitherto projectivities between one-dimensional forms of the 
first degree only. We shall now see how projectivities exist also be- 
tween one-dimensional forms of the second degree, and also between 
a one-dimensional form of the first and one of the second degree. 
Many familiar theorems will hereby appear in a new light. 

As typical for the one-dimensional forms of the second degree we 
choose the conic. The corresponding theorems for the cone then 
follow by the principle of duality. 

Let TTj be a projective collineation between two planes a, a v and 
let C z be any conic in a. Any two projective pencils of lines in a 
are then transformed by ^ into two projective pencils of lines in a v 
such that any two homologous lines of the pencils in a are trans- 
formed into a pair of homologous lines in o^; for if TT be the projec- 
tivity between the pencils in a', T^TTTrf 1 will be a projectivity between 
the pencils in a (cf. 74). Two projective pencils of lines generating 
the conic <7 2 thus correspond to two pencils of lines in a generating 
a conic Cf. The transformation ir^ then transforms every point of C 2 
into a unique point of Cf. Similarly, it is seen that ir^ transforms 
every tangent of (7 2 into a unique tangent of Cf. 

DEFINITION. Two conies are said to be protective if to every point of 
one corresponds a poinjb of the other, and to every tangent of one 



76] 



TEAJNTSFOEMATION OF COJSTICS 



213 



corresponds a tangent of the other, in such a way that this correspond- 
ence may be brought about by a projective collineation between the 
planes of the conies. The projective collineation is then said to 
generate the projectivity between the conies. 

Two conies in different planes are projective, for example, if one is the pro- 
jection of the other from a point on neither of the two planes. If the second 
of these is projected back on the plane of the first from a new center, we 
obtain two conies in the same plane that are projective. We will see presently 
that two projective conies may also coincide, in which case we obtain a pro- 
jectivity on a conic. 

THEOREM 8. Two conies that are projectile with a third are 
projective. 

Proof. This is an immediate consequence of the definition and the 
fact that the resultant of two collineations is a collineation. 

We proceed now to prove the fundamental theorem for projec- 
tivities between two conies. 

THEOREM 9. 'A projectivity "between two conies is uniquely deter- 
mined if three distinct points (or tangents) of one are made to corre- 
spond to three distinct points (or tangents] of the other. 





FIG. 86 



Proof. Let <7 2 , <7 2 be the two conies (fig. 86), and let A, B, be 
three points of (7 2 , and A f , B r , C r the corresponding points of Cf. Let 
P and P f be the poles of AB and A'Jff with respect to <7 2 and <7 X 
respectively. If now the collineation TT is defined by the relation 
7r(ABCP)=zA f B r C'P f (Theorem 18, Chap. IV), it is clear that the 
conic (7 2 is transformed by TT into a conic through the points A', B f , C r 9 
with tangents A'P* and B f P f . This conic is uniquely determined by 
these specifications, however, and is therefore identical with CJ. The 
collineation TT then transforms C 2 into Cf in such a way that the 
points A, B, C are transformed into A 1 , B f , C f respectively. Moreover, 



214 ONE-DIMENSIONAL PEOJECTIYITIES [CHAP.VIII 

suppose 7r f were a second collineation transforming C 2 into C* in the 
way specified. Then 7r'"V would be a collineation leaving A, B, C, P 
invariant ; i.e. TT = TT'. 

The argument applies equally well if A f B'C f are on the conic O 2 , 
i.e. when the two conies 2 , C? coincide. In this case the projectivity 
is on the conic C. This gives 

COROLLARY 1. A projectivity on a conic is uniquely determined when 
three pairs of homologous elements (points or tangents] are given. 

Also from the proof of the theorem follows 

COROLLARY 2. A collineation in a plane which transforms three 
distinct points of a conic into three distinct points of the same conic and 
which transforms the pole of the line joining two of the first three 
points into the pole of the line joining the two corresponding points 
transforms the conic into itself. 

The two following theorems establish the connection between pro- 
jectivities between two conies and projectivities between one-dimen- 
sional forms of the first degree. 

THEOREM 10. If A and B 1 are THEOREM 10'. If a and V are 
any two points of two projective any two tangents of two protective 
conies C* and Of respectively, the conies C z and C* respectively, the 
pencils of lines with centers at A pencils of points on a and V are 
and B ! are protective if every pair protective if every pair of homol- 
of homologous lines of these pencils ogous points on these lines is on 
pass through a pair of homologous a pair of homologous tangents of 
points on the two conies respectively, the conies respectively. 

Proof. It will suffice to prove the theorem on the left. Let A' be 
the point of (7f homologous with A. The collineation which generates 
the projectivity between the conies then makes the pencils of lines at 
A and A f protective, in such a way that every pair of homologous 
lines contains a pair of homologous points of the two conies. The pen- 
cil of lines at B f is projective with that at A 1 if they correspond in 
such a way that pairs of homologous lines intersect on C* (Theorem 
2, Chap. V). This establishes a projective correspondence between 
the pencils at A and B r in which any two homologous lines pass 
through two homologous points of the conies and proves the theorem. 

It should be noted that in this projectivity the tangent to C 2 at A 
corresponds to the line of the pencil at B' passing through A f . 



70] 



TBAJtfSFOKMATION OF COLICS 



215 



COROLLARY. Conversely, if two 
conies correspond in such a way 
that every pair of homologous tan- 
gents is on a pair of homologous 
points of two projective pencils of 
points whose axes are tangents of 
the conies, they are protective. 



COROLLARY. Conversely, if two 
conies correspond in such a way 
that every pair of homologoits 
points is on a pair of homologous 
lines of two protective pencils of 
lines whose centers are on the 
conics 9 they are projective. 

Proof. This follows from the fact that the projectivity between the 
pencils of lines is uniquely determined by three pairs of homologous 
lines. A projectivity between the conies is also determined by the 
three pairs of points (Theorem 9), in which three pairs of homolo- 
gous 4 lines of the pencils meet the conies. But by what precedes 
and the theorem above, this projectivity is the same as that described 
in the corollary on the left. The corollary on the right may be proved 
similarly. If the two conies are in the same plane, it is simply the 
plane dual of the one on the left. 

By means of these two theorems the construction of a projectivity 
between two conies is reduced to the construction of a projectivity 
between two primitive one-dimensional forms. 

It is now in the spirit of our previous definitions to adopt the 
following : 

DEFINITION. A Hue conic and 
a pencil of points whose axis is 
a line of the conic are said to be 
perspective if they correspond in 
such a way that every line of the 
conic passes through the homolo- 
gous point of the pencil of points. 
A line conic and a pencil of lines 
are said to be perspective if every 
two homologous lines meet in a 
point of a pencil of points whose 



DEFINITION. A point conic and 
a pencil of lines whose center is a 
point of the conic are said to be 
perspective if they correspond in 
such a way that every point of 
the conic is on the homologous 
line of the pencil. A point conic 
and a pencil of points are said to 
be perspective if every two homol- 
ogous points are on the same line 
of a pencil of lines whose center 
is a point of the conic. 



axis is a line of the conic. 



The reader will now readily verify that with this extended use of 
the term perspective, any sequence of perspectivities leads to a pro- 
jectivity. For example, two pencils of lines perspective with the same 
point conic are projective by Theorem 2, Chap. V; two point conies 




216 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP.VIII 

perspective with the same pencil of lines or with the same pencil of 
points are projective by Theorem 10, Cor., etc. 

Another illustration of this extension of the notion of perspectivity 
leads readily to the following important theorem : 

THEOEEM 11. Two conies which are not in the same plane and have 
a common tangent at a point A are sections of one and the same cone. 

Proof. If the two conies (7 2 , C* (fig. 87) are made to correspond 
in such a way that every tangent x of one is associated with that 

tangent #'of the other 
which meets x in a 
point of the common 
tangent a of the conies, 
they are projective. 
For the tangents of 
the conies are then 

87 perspective with the 

same pencil of points 

(cf. Theorem 10', Cor.). Every pair of homologous tangents of the two 
conies determines a plane. If we consider the point of intersection 
of three of these planes, say, those determined by the pairs of tangents 
W, erf, dd f , and project the conic C* on the plane of <7 2 from 0, there 
results a conic in the plane of C*. This conic has the lines 6, c, d for 
tangents and is tangent to a at A ; it therefore coincides with C 2 
(Theorem 6', Chap. V). The two conies (7 2 , C? then have the same 
projection from 0, which proves the theorem.* 

EXERCISES 

1. State the theorems concerning cones dual to the theorems of the preced- 
ing sections. 

2. By dualizing the definitions of the last article, define what is meant by 
the perspectivity between cones and the primitive one-dimensional forms. 

3. If two projective conies have three self-corresponding points, they are 
perspective with a common pencil of lines. 

4. If two projective conies have four self -corresponding elements, they 
coincide. 

5. State the space duals of the last two propositions. 

*It will be seen later that this theorem leads to the proposition that any conic 
may be obtained as the projection of a circle tangent to it in a different plane. 



76, 77] 



PEOJECTIVITIES OX A CONIC 



217 



6. If a pencil of lines and a conic in the plane of the pencil are projective, 
but not perspective, not more than three lines of the pencil pass through their 
homologous points on the conic. (Hint. Consider the points of intersection of 
the given conic with the conic generated by the given pencil and a pencil of 
lines perspective with the given conic.) Dualize. 

7. The homologous lines of a line conic and a projective pencil of lines in 
the same plane intersect in points of a " curve of the third order " such that 
any line of the plane has at most three points in. common with it. (This fol- 
lows readily from the last exercise.) 

8. The homologous elements of a cone of lines and a projective pencil of 
planes meet in a " space curve of the third order" such that any plane has 
at most three points in common with it. 

9. Dualize the last two propositions. 

77. Projectivities on a conic. We have seen that two protective' 
conies may coincide (Theorems 8-10), in which case we obtain a 
projective correspondence among the points or the tangents of the 




conic. The construction of the projectivity in this case is very 
simple, and leads to many important results. It results from the 
following theorems": 

THEOREM 12. If A, A f are any THEOREM 12'. If a, a f are any 

two distinct homologous points of two distinct homologous tangents 

a projectivity on a conic, and B,3 ! ; of a projectivity on a conic, and 

C, C r ; etc., are any other pairs of I, V; c, c f ; etc., are any other pairs 



218 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP, vin 

homologous points, the lines A r B of homologous tangents, the points 
and AB', A f C and AC 1 , etc., meet a f b and aV, a f c and ac f , etc., are 
in points of the same line; and collinear with the same point; 
this line is independent of the pair and this point is independent of 
AAl chosen. the pair aa 1 chosen. 

Proof. The pencils of lines A* (ABC- - ) and A(A'B f C f < - ) are pro- 
jective (Theorem 10), and since they have a self-corresponding line 
AA 1 , they are perspective, and the pairs of homologous lines of these 
two pencils therefore meet in the points of a line (fig. 88). This 
proves the first part of the theorem on the left. That the line thus 
determined is independent of the homologous pair AA! chosen then 
follows at once from the fact this line is the Pascal line of the simple 
hexagon AB'CA'BC 1 , so that the lines B J G and BC f and all other 
analogously formed pairs of lines meet on it. The theorem on the 
right follows by duality. 

DEFINITION. The line and the point determined by the above dual theo- 
rems are called the axis and the center of the projectivity respectively. 

COROLLARY 1. A (nonidentical) COROLLARY 1'. A (nonidenti- 

projectivity on a conic is wiiquely cal) projectivity on a conic is 

determined when the axis of ho- wniquely determined when the 

mology and one pair of distinct center and one pair of distinct 

homologous points are given. homologous tangents are given. 

These corollaries follow directly from the construction of the pro- 
jectivity arising from the above theorem. This construction is as 
follows : Given the axis o and a pair of distinct homologous points 
AA f , to get the point P 1 homologous with any point P on the conic ; 
join P to A! ; the point P f is then on the line joining A 'to the point 
of intersection of A'P with o. Or, given the center and a pair of 
distinct homologous tangents aa f , to construct the tangent p f homolo- 
gous with any tangent p ; the line joining the point a r p to the center 
meets a in a point of p r . 

COROLLARY 2. Every double COROLLARY 2'. Every doulle 

point of a projectivity on a conic line of a projectivity on a conic 

is on the axis of the projectivity ; contains the center of the projec- 

and, conversely, every point com- tivity; and, conversely, every tan- 

mon to the axis and the conic is gent of a conic through the center 

a double point. is a double line of the projectivity. 



77] 



PEOJECTIYITIES ON A CONIC 



219 



COROLLARY 3. A projectimty COROLLARY 3'. A projectimty 

among the points on a conic is among the tangents to a conic is 

parabolic if and only if the axis parabolic if and only if the center 

is tangent to the conic. . is a point of the conic. 

THEOREM 13. A projectimty among the points of a conic determines 
a projectimty of the tangents in which the tangents at pairs of homol- 
ogous points are homologous. 

Proof. This follows at once from the fact that the collineation in 
the plane of the conic which generates the projectivity transforms 
the tangent at any point of the conic into the tangent at the homol- 
ogous point, and hence also generates a projectivity between the 
tangents. 

THEOREM 14. The center of a projectivity of tangents on a conic 
and the axis of the corresponding projectimty of points are pole and 
polar with respect to the conic. 

A 




FIG. 89 



Proof. Let AA ! , BB ! t CC ! (fig. 89) he three pairs of homologous 
points (AA 1 being distinct), and let A'B and AB ! , A'C and AC f , meet 
in points E and S respectively, which determine the axis of the pro- 
jectivity of points. Now the polar of E with respect to the conic is 
determined by the intersections of the pairs of tangents at A f , B and 
A) B 1 respectively; and the polar of S is ' determined by the pairs of 
tangents at A 1 , C and A, C r respectively (Theorem 13, Chap. V). The 
pole of the axis JZS is then determined as the intersection of these 



220 ' ONE-DIMENSIONAL PEOJEOTIVITIES [CHAP.VIII 

two polars (Theorem 17, Chap. V), But by definition these two polars 
also determine the center of the projectivity of tangents. 

This theorem is obvious if the projectivity has double elements ; the proof 
given, however, applies to all cases. 

The collineation generating the projectivity on the conic transforms 
the conic into itself and clearly leaves the center and axis invariant. 
The set of all collineations in the plane leaving the conic invariant 
form a group (cf. p. 67). In determining a transformation of this 
group, any point or any line of the plane may be chosen arbitrarily 
as a double point or a double line of the collineation ; and any two 
points or lines of the conic may be chosen as a homologous pair of 
the collineation. The collineation is then, however, uniquely deter- 
mined. In fact, we have already seen that the projectivity on the 
conic is uniquely determined by its center and axis and one pair of 
homologous elements (Theorem 12, Cor. 1); and the theorem just 
proved shows that if the center of the projectivity is given, the axis 
is uniquely determined, and conversely. 

COROLLARY 1. A plane protective collineation which leaves a non- 
degenerate conic in its plane invariant is of Type I if it lias two 
doutle points on the conic, tmless it is of period tivo, in which case it 
is of Type IV; and is of Type III if the corresponding projectivity 
on the conic is parabolic. 

COROLLARY 2. An elation or a collineation of Type II transforms 
every nondegenerate conic of its plane into a different conic. 

COROLLARY 3. A plane $rojective collineation which leaves a conic 
in its plane invariant and has no double point on the conic has one 
and only one double point in the plane. 

THEOREM 15. The group of protective collineations in a plane leav- 
ing a nondegenerate conic invariant is simply isomorphic* with the 
general protective group on a line. 

Proof. Let A be any point of the invariant conic. Any projectivity 
on the conic then gives rise to a projectivity in the flat pencil at A in 
which two lines are homologous if they meet the conic in a pair of 
homologous points. And, conversely, any projectivity in the flat 

* Two groups are said to be simply isomorpMc if it is possible to establish a (1,1) 
correspondence between the elements of the two groups such that to the product of 
any two elements of one of the groups corresponds the product of the two corre- 
sponding elements of the other. 



77,78] INVOLUTIONS 221 

pencil at A gives rise to a projectivity on the conic. The group of all 
projectivities on a conic is therefore simply isomorphic with the group 
of all projectivities in a fiat pencil, since it is clear that in the corre- 
spondence described between the projectivities in the flat pencil and 
on the conic, the products of corresponding pairs of projectivities will 
be corresponding projectivities. Hence the group of plane collineations 
leaving the conic invariant is simply isomorphic with the general pro- 
jective group in a flat pencil and hence with the general projective 
group on a line. 

78. Involutions. An involution was defined (p. 102) as any projec- 
tivity in a one-dimensional form which is of period two, ie. by the 
relation I 2 = 1 (I = 1), where I represents an involution. This relation 
is clearly equivalent to the other, I = I~ 1 (l= 1), so that any projec- 
tivity (not the identity) in a one-dimensional form, which is identical 
with its inverse, is an involution. It will be recalled that since an in- 
volution makes every pair of homologous elements correspond doubly, 
i.e. A to A! and A f to A, an involution may also be considered as a 
pairing of the elements of a one-dimensional form ; any such pair is 
then called a conjugate, pair of the involution. We propose now to 
consider this important class of projectivities more in detail. To this 
end it seems desirable to collect the fundamental properties of invo- 
lutions which have been obtained in previous chapters. They are as 
follows : 

1. If the relation 7r z (A) A Jwlds for a single element A (not a 
double element of TT) of a one-dimensional form, the projectivity TT is 
an involution, and the relation holds for every element of the form 
(Theorem 26, Chap. IV). 

2. An involution is uniquely determined when two pairs of conjib- 
gate elements are given (Theorem 26, Cor., Chap. IV). 

3. The opposite pairs of any quadrangular set are three pairs of 
an involution (Theorem 27, Chap. IV). 

4. If M, N are distinct double elements of any projectivity in a 
one-dimensional form and A, A r and B, B l are any two pairs of 
homologous elements of the projectivity, the pairs of elements MN 9 AB J 
A f B are three pairs of an involution (Theorem 27, Cor. 3, Chap. IV), 

5. If M, N are double elements of an involution, they are distinct, 
and every conjugate pair of the involution is harmonic with M, N 
(Theorem 27, Cor. 1, Chap. IV). 



222 



ONE-DIMENSIONAL PEOJECTIVITIES [CHAP, vm 



6. An involution is uniquely determined, if two double elements are 
given, or if one double element and another conjugate pair are given. 
(This follows directly from the preceding.) 

7. An involution is represented analytically ly a bilinear form 
cscx r a (JG + x') & = 0, or ly the transformation 



ex a 

(Theorem 12, Cor. 3, Chap. YI). 

8. An involution with double elements m, n may be represented ly 
the transformation 



x' 



(Theorem 1, (Tors. 2, 3, Chap. VIII). 

We recall, finally, the Second Theorem of Desargues and its various 
modifications ( 46, Chap. V), which need not be repeated at this 
place. It has been seen in the preceding sections that any projec- 
tivity in a one-dimensional primitive form may he transformed into a 
projectivity on a conic. We shall find that the construction of an in- 
volution on a conic is especially simple, and may be used to advantage 
in deriving further properties of involutions. Under duality we may 

confine our consideration 

*\ to the case of an involu- 

' tion of points on a conic. 
y THEOREM 16. The lines 
joining the conjugate points 
of an involution on a conic 
all pass through the center 
of the involution. 

Proof. Let A, A' (fig. 90) 
be any conjugate pair (A 
not a double point) of an 
involution of points on a 
conic C 2 . The line A A! is then an invariant line of the collineation gener- 
ating the involution. Every line joining a pair of distinct conjugate 
points of the involution is therefore invariant, and the generating 
collineation must be a perspective collineation, since any collineation 
leaving four lines invariant is either perspective or the identity 




. 90 



78] INVOLUTIONS 223 

(Theorem 9, Cor. 3, Chap. III). It remains only to show that the 
center of this perspective collineation is the center of the involution. 
Let B y B f (B not a double point) be any other conjugate pair of the 
involution, distinct from A, A 1 . Then the lines AB f and A'B inter- 
sect on the axis of the involution. But since B t B r correspond to each 
other doubly, it follows that the lines AB and A f B f also intersect 
on the axis. This axis then joins two of the diagonal points of the 
quadrangle AA ! BB r . The center of the perspective collineation is 
determined as the intersection of the lines AA r and BB ! , i.e. it is 
the third diagonal point of the quadrangle AA'BB*. The center of 
the collineation is therefore the pole of the axis of the involution 
(Theorem 14, Chap. V) and is therefore (Theorem 14, above) the center 
of the involution. 

Since this center of the involution is clearly not on the conic, the 
generating collineation of any involution of the conic is a homology, 
whose center and axis o are pole and polar with respect to the conic. 
A homology of period two is sometimes called a harmonic homol- 
ogy, since it transforms any point P of the plane into its harmonic 
conjugate with respect to and the point in which OP meets 
the axis. It is also called a protective reflection or a point-line reflec- 
tion. Clearly this is the only kind of homology that can leave a conic 
invariant. 

The construction of the pairs of an involution on a conic is now 
very simple. If two conjugate pairs A 9 A* and B, B f are given, the lines 
AA 1 and BB J determine the center of the involution. The conjugate 
of any other point C on the conic is then determined as the intersec- 
tion with the conic of the line joining C to the center. If the involu- 
tion has double points, the tangents at these points pass through the 
center of the involution ; and, conversely, if tangents can be drawn to 
the conic from the center of the involution, the points of contact of 
these tangents are double points of the involution. 

The great importance of involutions is in part due to the following 
theorem : 

THEOREM 17. Any projectimty in a one-dimensional form may le 
obtained, as the product of two involutions. 

Proof. Let II be the projectivity in question, and let A be any 
point of the one-dimensional form which is not a double point. 



224 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP.VIII 

Further, let U(A) = A! and ft (A 1 ) = A". Then, if I, is the involution 
of which A 1 is a double point and of which AA n is a conjugate pair 
(Prop. 6, p. 222), we have 



so that in the projectivity I^II the pair AA f corresponds to itself 
doubly. I^II is therefore an involution (Prop. 1, p. 221). If it be 
denoted by I 2 , we have 1^11 = 1^ or TL^I^I^ which was to be 
proved. 

This proof gives at once : 

COROLLARY 1. Any projectivity II may le represented as the prod- 
uct of two involutions, II = I 1 -I 2 , either of which (but not lotli) has 
an arbitrary point (not a double point of II) for a double point. 

Proof. We have seen above that the involution ^ may have an 
arbitrary point (A 1 ) for a double point. If in the above argument we 
let I 2 be the involution of which A f is a double point and AA rf is a 
conjugate pair, we have II I & (A'A' r ) = A r! A r ; whence II I 2 is an invo- 
lution, say I r We then have II = 1^ I 2 , in which I 2 has the arbitrary 
point A f for a double point. 

The argument given above for the proof of the theorem applies 
without change when A = A ff , i.e. when the projectivity II is an in- 
volution. This leads readily to the following important theorem : 

COROLLARY 2. If A A! is a conjugate pair of an involution I, the 
involution of which A, A! are double points transforms I into itself, 
and the two involutions are commutative. 

Proof. The proof of Theorem 17 gives at once I I^Lj, where \ 
is determined as the involution of which A,A f are double points. We 
have then I x - 1 = I 2 , from which follows, by taking the inverse of both 
sides of the equality, I ^ = I^ 1 = I 2 , or l l - 1 = ! I a , or I x I I x = I. 

As an immediate corollary of the preceding we have 

COROLLARY 3. The product of two involutions with double points 
A, A f and B, B ! respectively transforms into itself the involution in 
which A A f and B B' are two conjugate pairs. 

Involutions related as are the two in Cor. 2 above are worthy of 
special attention. 

DEFINITION. Two involutions are said to be harmonic if their 
product is an involution. 



78,79] INVOLUTIONS 225 

THEOREM 18. Two harmonic involutions are commutative. 

Proof. If I v I 2 are harmonic, we have, by definition, I x - 1 2 = I 3 , where 
I 3 is an involution. This gives at once the relations 1^ I 2 - 1 3 = 1 and 

COROLLARY. Conversely, if two distinct involutions are commutative, 
they are harmonic. 

For from the relation Ij-I^Ig-^ follows (I 1 I 2 ) 2 = 1; i.e. I^I^ 
is an involution, since I^I^l. 

DEFINITION. The set of involutions harmonic with a given involu- 
tion is called a pencil of involutions. 

It follows then from Theorem 17, Cor. 2, that the set of all involu- 
tions in which two given elements form a conjugate pair is a pencil. 
Thus the double points of the involutions of such a pencil are the 
pairs of an involution. 

79. Involutions associated with a given projectivity* In deriving 
further theorems on involutions we shall find it desirable to suppose 
the projectivities in question to be on a conic. 

THEOREM 19. If a projectivity on a conic is represented as the product 
of two involutions, the axis of the projectivity is the line joining the 
centers of the two involutions. 

Proof. Let the given projec- 
tivity be II = I 2 - I x ; I v I 2 being 
two involutions. Let O v 2 be 
the centers of I x , I 2 respectively 
(fig. 91), and let "A and B be 
any two points on the conic 
which are not double points of 
either of the involutions \ or I 2 
and which are not a conjugate 
pair of I t or I 3 . If, then, we 
have II (AB) = A'B 1 , we have, by 
hypothesis, \(AB) = A^B^ and 

) = A f B'- ? AV B 1 being uniquely determined points of the conic, 




B 



such that the lines AA V BB^ intersect in 1 and the lines A^AI 9 
intersect in 2 . The Pascal line of the hexagon AA^BB^ff then 
passes through O v O z and the intersection of the lines AB 1 and A'B. 
But the latter point is a point on the axis of II. This proves the theorem. 



226 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP, vni 

COROLLARY. A projectivity on a conic is the product of two involu- 
tions, the center of one of which may le any arbitrary point (not a 
double point) on the axis of the projectivity ; the center of the other 
is then uniquely determined. 

Proof. Let the projectivity II be determined by its axis I and any 
pair of homologous points A, A 1 (fig. 91). Let 1 be any point on the 
axis not a double point of II, and let I x be the involution of which 
O x is the center. If, then, l l (A) = A v the center 2 of the involution 
I such that II = I 3 - I x is clearly determined as the intersection of the 
line A^A! with the axis. For by the theorem the product / 2 J a is a 
projectivity having I for an axis, and it has the points A, A f as a homol- 
ogous pair. This shows that the center of the first involution may 
be any point on the axis (not a double point). The modification of 
this argument in order to show that the center of the second involu- 
tion may be chosen arbitrarily (instead of the center of the first) is 
obvious. 

THEOREM 20. There is one and only one involution commutative 
with a given nonparabolic noninvolutoric projectivity. If the projec- 
tivity is represented on a conic, the center of this involution is the 
center of the projectivity. 

Proof. Let the given nonparabolic projectivity II be on a conic, 
and let I be any involution commutative with H ; Le. such that we 
have II I = I- II. This is equivalent to II I II"" 1 = I. That is to say, 
I is transformed into itself by II. Hence the center of I is transformed 
into itself by the collineation generating II. But by hypothesis the 
only invariant points of this collineation are its center and the points 
(if existent) in which its axis meets the conic. Since the center of I 
cannot be on the conic, it must coincide with the center of II. More- 
over, if the center of I is the same as the center of II, I is trans- 
formed into itself by the collineation generating II, II -LIT" * = I. 
Hence H I = I H. Hence I is the one and only involution commu- 
tative with II, 

COROLLARY 1. There is no involution commutative with a parabolic 
projectivity. 

DEFINITION. The involution commutative with a given nonpara- 
bolic noninvolutoric projectivity is called the involution belonging to 
the given projectivity. An involution belongs to itself. 



79] INVOLUTION'S 227 

COROLLARY 2. If a nonparabolic projectivity has double points, the 
involution belonging to the projectivity has the same double points. 

For if the axis of the projectivity meets the conic in two points, 
the tangents to the conic at these points meet in the pole of the axis. 

It is to be noted that the involution I belonging to a given projec- 
tivity II transforms II into itself, and is transformed into itself by II. 
Indeed, from the relation II I = I II follow at once the relations 
I.II-I = II and DM'II~' 1 = L Conversely, from the equation 

iM-n- 1 follows n.i = Mi. 

THEOREM 21. The necessary and sufficient condition that two invo- 
lutions on a conic be harmonic is that their centers be conjugate with 
respect to the conic. 

Proof. The condi- 3 \^^-^ A 

tion is sufficient. For 
let I 1? I 2 be two invo- 
lutions on the conic 
whose centers O v 2 
respectively are con- 
jugate with respect 
to the conic (fig. 92). 
Let A be any point 
of the conic not a 
double point of either involution, and let I 1 (A)^A 1 and I Z (AJ=A'. 
If, then, l l (A f ) = A[, the center O l is a diagonal point of the quadrangle 
AA^A!A[, and the center 2 is on the side A^A 1 . Since, by hypothesis, 
2 is conjugate to O l with respect to the conic, it must be the diago- 
nal point on AjA f , i.e. it must be collinear with AA[. We have then 
l z >I^(AA J )A f A, i.e. the projectivity l^\ is an involution I 8 . The 
center 3 of the involution I 3 is then the pole of the line 1 3 with 
respect to the conic (Theorem 19). The triangle 1 a 8 is therefore 
self-polar with respect to the conic. It follows readily also that the 
condition is necessary. Tor the relation I^Ig Ig leads at once to 
the relation 1^ = 1^1,. If O v 3 , 3 are the centers respectively of 
the involutions I v I 2 , I 3 , the former of these two relations shows 
(Theorem 19) that B is the pole of the line 0^; while the latter 
shows that 3 is the pole of the line 0^0 y The triangle 0^0. 2 9 is 
therefore self-polar. 




228 



ONE-DIMENSIONAL PEOJECTIVITIES [CHAP, vm 



COROLLARY 1. Given any two involutions, there exists a third invo- 
lution which is harmonic with each of the given involutions. 

For if we take tlie two involutions on a conic, the involution whose 
center is the pole with respect to the conic of the line joining the 
centers of the given involutions clearly satisfies the condition of the 
theorem for each of the latter. 

COROLLARY 2. Three involutions each of which is harmonic to the 
other two constitute, together with the identity, a group. 

COROLLARY 3. The centers of all involutions in a pencil of involu- 
tions are collinear. 

THEOREM 22. The set of all projectivities to which belongs the same 
involution I forms a commutative group. 

Proof. If II, IIj are two projectivities to each of which belongs the 
involution I, we have the relations I -II -I = 13 and I-II 1 -I=:II 1 , 
from which follows I-II~ 1 I = n~ 1 and, by multiplication, the rela- 
, which shows that the set 
forms a group. To show 
that any two projectivities 
of this group are commu- 
tative, we need only sup- 
pose the projectivities 
given on a conic. Let A 
be any point on this 
conic, and let II (A) = A ! 
and TL l (A r ) = A, so that 
^ 1 / . Since the 



Hani I.II.M.IVI=] 

A A, 




same involution I belongs, 

by hypothesis, both to II and 11^ these two projectivities have the 
same axis ; let it be the line I (fig. 93). The point IL^ (A) = A 1 is now 
readily determined (Theorem 12) as the intersection with the conic of 
the line joining A f to the intersection of the line AA[ with the axis I. 
In like manner, H(A^) is determined as tie intersection with the 
conic of the line joining A to the intersection of the line A^A r with 
the axis I Hence H (AJ = A[, and hence II U^A) = A[. 

It is noteworthy that when the common axis of the projectivities 
of this group meets the conic in two points, which are then common 
double points of all the projectivities of the group, the group is the 



79] INVOLUTIONS 229 

same as the one listed as Type 2, p. 209. If, however, our geometry 
admits of a line in the plane of a conic but not meeting the conic, the 
argument just given proves the existence of a commutative group 
none of the projectivities of which have a double point. 

THEOREM 23. Two involutions have a conjugate pair (or a double 
point) in common if and only if the product of the two invohctions 
has two double points (or is parabolic). 

Proof. This follows at once if the involutions are taken on a conic. 
For a common conjugate pair (or double point) must be on the line 
joining the centers of the two involutions. This line must then meet 
the conic in two points (or be tangent to it) in order that the involu- 
tions may have a conjugate pair (or a double point) in common. 

EXERCISES 

1 . Dualize all the theorems and corollaries of the last two sections. 

2. The product of two involutions on a conic is parabolic if and only if the 
line joining the centers of the involutions is tangent to the conic. Dualize. 

3. Any involution of a pencil is uniquely determined when one of its con- 
jugate pairs is given. 

4. Let II be a noninvolutoric projectivity, and let I be the involution be- 
longing to II; further, let TL(AA') = A' A", A being any point on which the 
projectivity operates which is not a double point, and let 1(4') = A. Show, 
by taking the projectivity on a conic, that the points A'A are harmonic 
with the points A A". 

5. Derive the theorem of Ex. <4 directly as a corollary of Prop. 4, p. 221, 
assuming that the projectivity n has two distinct double points. 

6. From the theorem of Ex. 4 show how to construct the involution be- 
longing to a projectivity n on a line without making use of any double points 
the projectivity may have. 

7. A projectivity is uniquely determined if the involution belonging to it 
and one pair of homologous points are given. 

8. The product of two involutions I 1? I 2 is a projectivity to which belongs 
the involution which is harmonic with each of the involutions I x , I 2 . 

9. Conversely, every projectivity to which a given involution I belongs can 
be obtained as the product of two involutions harmonic with I. 

10. Show that any two projectivities n i9 II 3 may be obtained as the 
product of involutions in the form H l = I'I 19 II 3 = I 3 -I; and hence that the 
product of the two projectivities is given by IVI^ = I 2 -I r 

11. Show that a projectivity H = I-Ii may also be written IE = I 2 -I, I 2 
being a uniquely determined involution ; and that in this case the two invo- 
lutions Ij, I 2 are distinct unless n is involutoric. 



230 ONE-DIMENSIONAL PEOJECTIVITIES [CHAP.VIII 

12. Show that if I x , I 2 , I 3 are three involutions of the same pencil, the 
relation (I^^-Ig) 2 ^ 1 must hold. 

13. If aa', W, cc' are the coordinates of three pairs of points in involution, 

u J.T. 4. '- & &'- c c'-a - 

show that r = 1. 

a 7 c 6 a c' b 

80. Harmonic transformations. The definition of harmonic involu- 
tions in the section above is a special case of a more general notion 
which can he defined for (1, 1) transformations of any kind whatever. 

DEFINITION. Two distinct transformations A and B are said to be 
harmonic if they satisfy the relation (AB" 1 ) 2 = 1 or the equivalent 
relation (BA- 1 )** 1, provided that AB" 1 ^ 1. 

A number of theorems which are easy consequences of this defini- 
tion when taken in conjunction with the two preceding sections are 
stated in the following exercises. (Of. C. Segre, Note sur les homo- 
graphies binaires et leur faisceaux, Journal fur die reine und ange- 
wandte Mathematik, Vol. 100 (1887), pp. 317-330, and H. Wiener, 
Ueber die aus zwei Spiegelungen zusammengesetzten Verwandt- 
schaften, Berichte d. K. sachsischen Gesellschaft der Wissenschaften, 
Leipag, Vol. 43 (1891), pp. 644-673.) 

EXERCISES 

1. If A and B are two distinct involutoric transformations, they are har- 
monic to their product AB. 

2. If three involutoric transformations A, B, T satisfy the relations 
(ABF) 2 = 1 , ABF 5* 1 , they are all three harmonic to the transformation AB. 

3. If a transformation S is the product of two involutoric transformations 
A, B (i.e. 2 = AB) and T is an involutoric transformation harmonic to 21, then 
we have (ABr) 2 =l. 

4. If A, J3, C, A', B', C' are six points of a line, the involutions A, B, T, 
such that T(AA') ~B'B, A(') = CC, *&(CC?)=A'A, are all harmonic to 
the same projectivity. Show that if the six points are taken on a conic, this 
proposition is equivalent to Pascal's theorem (Theorem 3, Chap. V). 

5. The set of involutions of a one-dimensional form which are harmonic 
to a given nonparabolic projectivity form, a pencil. Hence, if an involution 
with double points is harmonic to a projectivity with two double points, the 
two pairs of double points form a harmonic set. 

6. Let be a fixed point of a line I, and let C be called the mid-point of a 
pair of points A, J3 } provided that C is the harmonic conjugate of with 
respect to A and B. If A, B, C, A', B', C'are any six points of I distinct 
from 0, and AB' have the same mid-point as A'B, and BC' have the same 
mid-point as B'C> then CA' will have the same mid-point as C?A* 



80,81] SCALE ON A CONIC 231 

7. Any two involutions of the same one-dimensional form determine a 
pencil of involutions. Given two involutions A, B and a point M, show how 
to construct the other double point of that involution of the pencil of which 
one double point is M. 

8. The involutions of conjugate points on a line I with regard to the conies 
of any pencil of conies in a plane with I form a pencil of involutions. 

9. If two nonparabolic projectivities are commutative, the involutions 
belonging to them coincide, unless both projectivities are involutions, in which 
case the involutions may be harmonic. 

10. If [II] is the set of projectivities to which belongs an involution I and 
A and B are two given points, then we have [II (.4)] -^ [II (By]. 

11. A conic through two of the four common points of a pencil of conies 
of Type J meets the conies of the pencil in pairs of an involution. Extend 
this theorem to the other types of pencils of conies. Dualize. 

12. The pairs of second points of intersection of the opposite sides of a 
complete quadrangle with a conic circumscribed to its diagonal triangle are in 
involution (Sturm, Die Lehre von den Geometrischen Verwandtschaften, 
Vol. I, p. U9). 

81. Scale on a conic. The notions of a point algebra and a scale 
which we have developed hitherto only for the elements of one- 
dimensional primitive forms may also be studied to advantage on a 
conic. The constructions for the sum and the product of two points 
(numbers) on a conic is remarkably simple. As in the case on the 
line, let 0, 1, co be any three arbitrary distinct points on a conic C\ 
Regarding these as the fundamental points of our scale on the conic, 
the sum and the product of any two points x, y on the conic (which 
are distinct from co) are denned as follows : 

DEFINITION. The conjugate of in the involution on the conic 
having oo for a double point and x, y for a conjugate pair is called 
the sum of the two points x 9 y and is denoted by x + y (fig. 94, left). 
The conjugate of 1 in the involution determined on the conic by the 
conjugate pairs 0, oo and x, y is called the product of the points x, 
y and is denoted by x - y (fig. 94, right). 

It will be noted that under Assumption P this definition is entirely 
equivalent to the definitions of the sum and product of two points on 
a line, previously given (Chap. VI). To construct the point x + y on 
the conic (fig. 94), we need only determine the center of the involution 
in question as the intersection of the tangent at oo with the line joining 
the points x, y. The point x + y is then determined as the intersection 
with the conic of the line joining the center to the point 0. Similarly, 



232 



ONE-DIMENSIONAL PKOJECTIYITIES [CHAP. VIII 



to obtain the product of the points x, y we determine the center of the 
involution as the intersection of the lines Oo> and xy. The point x - y 
is then the intersection with the conic of the line joining this center to 





FIG. 94 



the point 1. The inverse operations (subtraction and division) lead to 
equally simple constructions. Since the scale thus defined is obviously 
projective with the scale on a line, it is not necessary to derive again 
the fundamental properties of addition and subtraction, multiplication 
and division. It is clear from this consideration that the points of a 
conic form a field with reference to the operations just defined. This 
fact will be found of use in the analytic treatment of conies. 

At this point we will make use of it to discuss the existence of the 
square root of a number in the field of points. It is clear from the 

x 




FIG. 95 



preceding discussion that if a number x satisfies the equation x* = a, 
the tangent to the conic at the point x must pass through the inter- 
section of the lines Ooo and 1 a (fig. 95). A number a will therefore 
have a square root in the field if and only if a tangent can le drawn to 



81] 



SCALE ON A CONIC 



233 



the conic from the intersection of the lines Oco and 1 a; and, conversely, 
if the number a has a sqiiare root in the field, a tangent can be drawn 
to the conic from this point of intersection. It follows at once that if 
a number a has a square root x, it also has another which is obtained 
by drawing the second tangent to the conic from the point of inter- 
section of the lines Oco and la. Since this tangent meets the conic 
in a point which is the harmonic conjugate of x with respect to Oco, 
it follows that this second square root is x. It follows also from 
this construction that the point 1 has the two square roots 1 and 1 
in any field in which 1 and 1 are distinct, i.e. whenever R Q is satisfied. 

We may use these considerations to derive the following theorem, 
which will be used later. 

THEOREM 2-4. If AA r , BB* are any two distinct pairs of an involu- 
tion, there exists one and only one pair CC f distinct from BB 1 such 
that the cross ratios 
%(AA r , BB>) 
B (AA r , CC r ) 
equal. 

Proof. Let the 
involution be taken 
on a conic, and let 
the pairs AA! and 
BB f be represented 
by the points Oco 
and la respectively (fig. 96). Let xx 1 be any other pair of the invo- 
lution. We then have, clearly from the above, xx* = a. Further, the 
cross ratios in question give 

U (Oco, 1 a) = - , B (Oco, 005') = -.. 
x a x' 

These are equal, if and only if x j = ax, or if xx f = ax*. But this implies 
the relation a = ax 2 , and since we have a = 0, this gives x* = 1. The 
only pair of the involution satisfying the conditions of the theorem 
is therefore the pair (7(7' = 1, a. 

EXERCISES 

1. Show that an involution which has two harmonic conjugate pairs has 
double points if and only if 1 has a square root in the field. 

2. Show that any involution may be represented by the equation afx a. 




_ .,. 
FIG. 90 



234 



ONE-DIMENSIONAL PEOJECTIVITIES [CHAP. VIII 



3. The equation of Ex. 13, p. 230, is the condition that the lines joining 
the three pairs of points aa' ? W/, cc' on a conic are concurrent. 

4. Show that if the involution x'x = a has a conjugate pair W such that 
the cross ratio R(0o>, W) has the value X, the number a\ has a square root 
in the field. 

82. Parametric representation of a conic. Let a scale be established 
on a conic <7 2 by choosing three distinct points of the conic as the 
fundamental points, say, 0=0, M= oo, A= 1. Then let us establish 
a system of nonhomogeneous point coordinates in the plane of the 

conic as follows : Let 

- N the line OH be the x- 

axis, with as origin 
and M as <x> x (fig. 97). 
Let the tangents at 
and M to the conic 
meet in a point N 9 and 
let the tangent ON be 
the y-axis, with -2V" as 
oo y . Finally, let the 
point A be the point 
(1, 1), so that the line 
AN meets the #-axis 
in the point for which 
# = 1, and AM meets 
the ^-axis in the point for which y = 1. Now let P = X be any point 
on the conic. The coordinates (a?, y) of P are determined by the 
intersections of the lines PN and PM with the #-axis and the y-axis 
respectively. We have at once the relation 




FIG. 97 



since the points 0, oo, 1, X on the conic are perspective from M with 
points 0, oo, 1, y on the y-axis. To determine x in terms of X, we note, 
first, that from the constructions given, any line through N meets the 
conic (if at all) in two points whose sum in the scale is 0. In par- 
ticular, the points 1, 1 on the conic are collinear with N and the 
point 1 on the #-axis, and the points X, X on the conic are collinear 
with N and the point x on the re-axis. Since the latter point is also 
on the line joining and co on the conic, the construction for multi- 
plication on the conic shows that any line through, the point x on 



82] PAEAMETEIC BEPEESENTATION 235 

the #-axis meets the conic (if at all) in two points whose product is 
constant, and hence equal to X 2 . The line joining the point x on the 
#-axis to the point 1 on the conic therefore meets the conic again in 
the point X 2 . But now we have 0, co, 1, X 2 on the conic perspective 
from the point 1 on the conic with the points 0, co, 1, $ on the 
#-axis. This gives the relation 

x = X 2 . 

We may now readily express these relations in homogeneous form. 
If the triangle OMN is taken as triangle of reference, ON being 
x^ = 0, 031 being os a = 0, and the point A being the point (1, 1, 1), 
we pass from the nonhomogeneous to the homogeneous by simply 
placing x = xjx v y = xjx y The points of the conic C' 2 may then 
le represented ly the relations 

(1) a^ia^ifljBssX^X,:!. 

This agrees with our preceding results, since the elimination of \ 
between these equations gives at once 

a: '-^053=0, 

which we have previously obtained as the equation of the conic. 

It is to be noted that the point M on the conic, which corresponds 
to the value X = oo, is exceptional in this equation. This exceptional 
character is readily removed by writing the parameter X homogene- 
ously XsaX^Xy Equations (1) then readily give 

THEOREM 25. A conic may le represented analytically ly the equa- 
tions x^ : X 2 : x s = \f : \\ : X 2 2 . 

This is called a parametric representation of a conic. 

EXERCISES 

1. Show that the equation" of the line joining two points X 19 X 2 on the conic 
(1) above is x l (X x -f X 2 ) # 2 + \iX 2 x s = ; and that the equation of the tan- 
gent to the conic at a point \i is x l 2 A^g + X 2 # 3 = 0. Dualize. 

2. Show that any collineation leaving the conic (1) invariant is of the form 



(Hint. Use the parametric representation of the conic and let the projectivity 
generated on the conic by the collineation be Xf = a\i + /JXg, X^ = y\i + 8X 2 .) 



CHAPTER IX 

GEOMETRIC CONSTRUCTIONS. INVARIANTS 

83. The degree of a geometric problem. The specification of a line 
by two of its points may be regarded as a geometric operation* The 
plane dual of this operation is the specification of a point by two 
lines. In space we have hitherto made use of the following geometric 
operations : the specification of a line by two planes (this is the 
space dual of the first operation mentioned above) ; the specification 
of a plane by two intersecting lines (the space dual of the second 
operation above) ; the specification of a plane by three of its points 
or by a point and a line ; the specification of a point by three planes 
or by a plane and a line. These operations are known as linear 
operations or operations of the first degree, and the elements deter- 
mined by them from a set of given elements are said to be obtained 
by linear constrictions, or by constrictions of the first degree. The 
reason for this terminology is found in the corresponding analytic 
formulations. Indeed, it is at once clear that each of the two linear 
operations in a plane corresponds analytically to the solution of a 
pair of linear equations ; and the linear operations in space clearly 
correspond to the solution of systems of three equations, each of the 
first degree. Any problem which can be solved by a finite sequence 
of linear constructions is said to be a linear prollem or a problem 
of the first* degree. Any such problem has, if determinate, one and 
only one solution. 

In the usual representation of the ordinary real projective geometry in a 
plane by means of points and lines drawn, let us say, with a pencil on a sheet 
of paper, the linear constructions .are evidently those that can be carried 
out by the use of a straightedge alone. There is no familiar mechanical 

* An operation on one or more elements is defined as a correspondence whereby 
to the set of given elements corresponds an element of some sort (cf. 48). If the 
latter element is uniquely defined by the set of given elements (in general, the order 
of the given elements is an essential factor of this determination), the operation is 
said to be one-valued. The operation referred to in the text is then a one-valued 
operation defined for any two distinct points and associating with any such pair 
(the order of the points is in this case immaterial) a new element, viz. a line. 

236 



8B] DEGEEE OF A GEOMETBIC PEOBLEM 237 

device for drawing lines and planes in space. But a picture (which is the 
section by a plane of a projection from a point) of the lines and points of 
intersection of linearly constructed planes may be constructed with a straight- 
edge (cf. the definition of a plane). 

As examples of linear problems we mention : (a) the determination 
of the point homologous with a given point in a projectivity on a 
line of which three pairs of homologous points are given; (6) the 
determination of the sixth point of a quadrangular set of which five 
points are given; (c) the determination of the second double point 
of a projectivity on a line of which one double point and two pairs of 
homologous points are given (this is equivalent to (&) ) ; (d) the deter- 
mination of the second point of intersection of a line with a conic, one 
point of intersection and four other points of the conic being given, etc. 

The analytic relations existing between geometric elements offer 
a convenient means of classifying geometric problems.* Confining 
ourselves, for the sake of brevity, to problems in a plane, a geometric 
problem consists in constructing certain points, lines, etc., which bear 
given relations to a certain set of points, lines, etc., which are sup- 
posed given in advance. In fact, we may suppose that the elements 
sought are points only ; for if a line is to be determined, it is sufficient 
to determine two points of this line ; or if a conic is sought, it is suffi- 
cient to determine five points of this conic, etc. Similar considera- 
tions may also be applied to the given elements of the problem, 
to the effect that we may assume these given elements all to be 
points. This merely involves replacing any given elements that are 
not points by certain sets of points having the property of uniquely 
determining these elements. Confining our discussion to problems in 
which this is possible, any geometric problem may be reduced to 
one or more problems of the following form : Given in a plane a 
certain finite number of points, to construct a point which shall bear 
to the given points certain given relations. 

In the analytic formulation of such a problem the given points 
are supposed to be determined by their coordinates (homogeneous or 
nonhomogeneous), referred to a certain frame of reference. The ver- 
tices of this frame of reference are either points contained among the 
given points, or some or all of them are additional points which we 

* The remainder of this section follows closely the discussion given in Castel- 
nuovo, Lezioni di geometria, Rome-Milan, Vol. I (1904), pp. 467 fl. 



238 G-EOMETEIC CONSTEUCTIONS [CHAP.IX 

suppose added to the given points. The set of all given points then 
gives rise to a certain set of coordinates, which we will denote by 
1, a, &, c, ,* and which are supposed known. These numbers to- 
gether with all numbers obtainable from them by a finite number of 
rational operations constitute a set of numbers, 



which we will call the domain, of rationality defined ly the data.^ In 
addition to the coordinates of the known points (which, for the sake 
of simplicity, we will suppose given in nonhomogeneous form), the 
coordinates (x, y) of the point sought must be considered. The con- 
ditions of the problem then lead to certain analytic relations which 
these coordinates x, y and a, 6, c - - must satisfy. Eliminating one 
of the variables, say #, we obtain two equations, 

/i()-o, /.fayHO, 

the first containing x but not y ; the second, in general, containing 
both x and y. The problem is thus replaced by two problems : the 
first depending on the solution of f (x) = to determine the abscissa 
of the unknown point ; the second to determine the ordinate, assum- 
ing the abscissa to be known. 

In view of this fact we may confine ourselves to the discussion of 
problems depending on a single equation with one unknown. Such 
problems may be classified according to the equation to which they 
give rise. A problem is said to be algebraic if the equation on which 
its solution depends is algebraic, i.e. if this equation can be put in 
the form 

(1) ar+a l ar~ l +ajf-*+> + a n = 0, 

in which the coefficients a l9 & 2 , - , a n are numbers of the domain of 
rationality defined by the data. Any problem which is not algebraic 
is said to be transcendental Algebraic problems (which alone will 
be considered) may in turn be classified according to the degree n of 

* In case homogeneous coSrdinates are used, a, &, c, . denote the mutual ratios 
of the coordinates of the given elements. 

t A moment's consideration will show that the points whose coordinates are 
numbers of this domain are the points obtainable from the data by linear construc- 
tions. Geometrically, any domain of rationality on a line may be defined as any 
class of points on a line which is closed under harmonic constructions ; i.e. such 
that if A, 2J, C are any three points of the class, the harmonic conjugate of A with 
respect to 3 and C is a point of the class, 



83] DEGREE OF A GEOMETEIC PEOBLEM 239 

the equation on which their solutions depend. We have thus problems 
of the first degree (already referred to), depending merely on the solution 
of an equation of the first degree; problems of the second degree, 
depending on the solution of an equation of the second degree, etc. 

Account must however be taken of the fact that equation (1) 
may be reducible within the domain K ; in other words, that the left 
member of this equation may be the product of two or more poly- 
nomials whose coefficients are numbers of K. In fact, let us suppose, 
for example, that this equation may be written in the form 



where j> 19 fa are two polynomials of the kind indicated, and of degrees 
n^ and n 2 respectively (n t + n z = ri). Equation (1) is then equivalent 
to the two equations 

<W*) = o> &(*) = o. 

Then either it happens that one of these two equations, e.g. the first, 
furnishes all the solutions of the given problem, in which case fa being 
assumed irreducible in K, the problem is not of degree n, but of degree 
n^ < n ; or, both equations furnish solutions of the problem, in which 
case fa also being assumed irreducible in K, the problem reduces to 
two problems, one of degree ^ and one of degree n y In speaking of 
a problem of the nth degree we will therefore always assume that 
the associated equation of degree n is irreducible in the domain of 
rationality defined by the data. Moreover, we have tacitly assumed 
throughout this discussion that equation (1) has a root ; we shall see 
presently that this assumption can always be satisfied by the intro- 
duction, if necessary, of so-called improper elements. It is important 
to note, however, since our Assumptions A, E, P do not in any way 
limit the field of numbers to which the coordinates of all elements 
of our space belong, and since equations of degree greater than one 
do not always have a root in a given field when the coefficients of 
the equation belong to this field, there exist spaces in which problems 
of degree higher than the first may have no solutions. Thus in the 
ordinary real projective geometry a problem of the second degree 
will have a (real) solution only if the quadratic equation on which 
it depends has a (real) root. 

The example of a problem of the second degree given in the next 
section will serve to illustrate the general discussion given above. 



240 GEOMETRIC CONSTRUCTIONS [CHAP. DC 

84. The intersection of a given line with a given conic. Given a 
conic defined, let us say, by three points A, B, and the tangents at 
A and B ; to find the points of intersection of a given line with this 
conic. Using nonhomogeneous coordinates and choosing as #-axis one 
of the given tangents to the conic, as y-axis the line joining the points 
A and J5, and as the point (1, 1) the point C, the equation of the conic 
may be assumed to be of the form 



The equation of the given line may then be assumed to be of the form 



The domain of rationality defined by the data is in this case 

K -[!,*, 2]. 

The elimination of y between the two equations above then leads to 
the equation 

(1) # 2 j;# 2= 0. 

This equation is not in general reducible in the domain K. The 
problem of determining the points of intersection of an arbitrary line 
in a plane with a given conic in this plane is then a problem of the 
second degree. If equation (1) has a root in the field of the geometry, it 
is clear that this root gives rise to a solution of the problem proposed ; 
if this equation has no root in the field, the problem has no solution. 
If, on the other hand, one point of intersection of the line with the 
conic is given, so that one root of equation (1), say a? ~ r, is known, 
the domain given by the data is 



and in this domain. (1) is reducible ; in fact, it is equivalent to the 

equation 

(x r + p) (a? r) = 0. 



The problem of finding the remaining point of intersection then 
depends merely on the solution of the linear equation 

x T + 2> = ; 

* There is no loss in generality in assuming this form ; for if in the choice of 
coordinates the equation of the given line were of the form x = c, we should merely 
have to choose the other tangent as se-axis to bring the problem into the form here 
assumed. 



84,85] PKOPOSITION K 2 241 

that is, the problem is of the first degree, as already noted among 
the examples of linear problems. 

It is important to note that equation (1) is the most general form 
of equation of the second degree. It follows that every problem of the 
second degree in a plane can ~be reduced to the construction of the points 
of intersection of an arbitrary line with a particular conic. We 
shall return to this later ( 86). 

85. Improper elements. Proposition K 2 . We have called attention 
frequently to the fact that the nature of the field of points on a line 
is not completely determined by Assumptions A, E, P, under which 
we are working. We have seen in particular that this field may be 
finite or infinite. The example of an analytic space discussed in the 
Introduction shows that the theory thus far developed applies equally 
well whether we assume the field of points on a line to consist of all 
the ordinary rational numbers, or of all the ordinary real numbers, 
or of all the ordinary complex numbers. According to which of these 
cases we assume, our space may be said to be the ordinary rational 
space, or the ordinary real space, or the ordinary complex space. 
Now, in the latter we know that every number has a square root. 
Moreover, each of the former spaces (the rational and the real) are 
clearly contained in the complex space as subspaces. Suppose now 
that our space S is one in which not every number has a square 
root. In such a case it is often convenient to be able to think of our 
space S as forming a subspace in a more extensive space S', in which 
some or all of these numbers do have square roots. 

We have seen that the ordinary rational and ordinary real spaces 
are such that they may be regarded as subspaces of a more exten- 
sive space in the number system associated with which the square 
root of any number always exists. In fact, they may be regarded as 
subspaces of the ordinary complex space which has this property. 
For a general field it is easy to prove that if a v a z , -, a n are any 
finite set of elements of a field F, there exists a field F, containing 
all the elements of F, such that each of the elements a v & 2 , , a n is a 
square in P. This is, of course, less general than the theorem that 
a field P exists in which every element of F is a square, but it is 
sufficiently general for many geometric purposes. In the presence of 
Assumptions A, E, P, H it is equivalent (cf. 54) to the following 
statement : 



242 GEOMETRIC CONSTKUCTIONS [CHAP.IX 

PROPOSITION K 2 . If any finite number of involutions arc given in 
a space S satisfying Assumptions A, E, P, there exists a space S' of 
which S is a subspace,* such that all the given involutions have 
double points in S'. 

A proof of this theorem will be found at the end of the chapter. 
The proposition is, from the analytic point of view, that the domain 
of rationality determined by a quadratic problem may be extended so 
as to include solutions of that problem. The space S' may be called 
an extended space. The elements of S may be called proper elements, 
and those of S' which are not in S may be called improper. A projec- 
tive transformation which changes every proper element into a proper 
element is likewise a proper transformation; one which transforms 
proper elements into improper elements, on the other hand, is called 
an improper transformation. Taking Proposition K 2 for the present as 
an assumption like A, E, P, and H , and noting that it is consistent 
with these other assumptions because they are all satisfied by the ordi- 
nary complex space, we proceed to derive some of its consequences. 

THEOREM 1. A proper one-dimensional projectivity without proper 
double elements may always be regarded in an extended space as 
having two improper double elements. (A, E, P, H , K 2 )f 

Proof. Suppose the projectivity given on a conic. If the involu- 
tion which belongs to this projectivity had two proper double points, 
they would be the intersections of the axis of the projectivity with 
the conic, and hence the given projectivity would have proper double 
points. Let S 7 be the extended space in which (K 2 ) the involution 
has double points. There are then two points of S' in which the 
axis of the projectivity meets the conic, and these are, by Theorem 20, 
Chap. VIII, the double points of the given projectivity. 

COROLLARY 1. If a line does not meet a conic in proper points, it 
may be regarded in an extended space as meeting it in two improper 
points. (A, E, P, H , K 2 ) 

COROLLARY 2. Every quadratic equation with proper coefficients has 
two roots which, if distinct, are both proper or both improper. (A, E, 
P, H , K 2 ) 

* We use the word subspa.ce to mean any space, every point of which is a point 
of the space of which it is a subspace. With this understanding the subspace may 
be identical with the space of which it is a subspace. The ordinary complex space 
then satisfies Proposition K 2 . t Of. Ex., p. 261. 



85] PBOPOSITION K 2 243 

For the double points of any projectivity satisfy an equation of 
the form car + (d a) x I = (Theorem 11, Cor. 4, Chap. VI), and 
any quadratic equation may be put into this form. 

THEOREM 2. Any two involutions in the same one-dimensional form 
have a conjugate pair in common, which may be proper or improper. 
(A, E, P, H , K 2 ) 

This follows at once from the preceding and Theorem 23, Chap. VIII. 

COROLLARY. In any involution there exists a conjugate pair, proper 
or impro%)er, which is harmonic with any given conjugate pair. (A, 
E, P, H , K 2 ) 

For the involution which has the given pair for double elements 
has (by the theorem) a pair, proper or improper, in common with the 
given involution. The latter pair satisfies the condition of the theorem 
(Theorem 27, Cor. 1, Chap. IV). 

We have seen earlier (Theorem 4, Cor., Chap. VIII) that any two 
involutions with double points are conjugate. Under Proposition K 2 
we may remove the restriction and say that any two involutions are 
conjugate in an extended space dependent on the two involutions. If 
the involutions are on coplanar lines, we have the following : 

THEOREM 3. Two involutions on distinct lines in the same plane 
are perspective (the center of perspectivity leing proper or improper), 
provided the point of intersection of the lines is a double point for 
loth or for neither of the involutions. (A, E, P, K 3 ) 

Proof. If the point of intersection of the two lines be a double 
point of each of the involutions, let Q and E be an arbitrary pair 
of one involution and Q 1 and R 1 an arbitrary pair of the other involu- 
tion. The point of intersection of the lines QQ r and RE J is then a 
center of a perspectivity which transforms elements which determine 
the first involution into elements which determine the second. If 
the point is a double point of neither of the two involutions, let 
M be a double point of one and M r of the other (these double points 
are proper or else exist in an extended space S r which exists by 
Proposition K 2 ). Also let .Fand N* be the conjugates of in the two 
involutions. Then by the same argument as before, the point of 
intersection of the lines MM f , NN f may be taken as the center of 
the perspectivity. 



244 GEOMETRIC CONSTRUCTIONS [CHAP, ix 

It was proved in 66, Chap. VII, that the equation of any point 
conic is of the form 



(1) a n x* + a,^* + a^&l + 2 a u ^ a + 2 a^x^ + 2 a 2S x 2 x s = ; 

but it was not shown that every equation of this form represents a 
conic. The line x l = contains the point (0, x v x 5 ) satisfying (1), 
provided the ratio x 2 : x s satisfies the quadratic equation 

a^p* + 2 a#B 9 x 9 + a^x* = 0. 



Similarly, the lines # 2 = and x z = contain points of the locus 
denned by (1), provided two other quadratic equations are satisfied. 
By Proposition K 2 there exists an extended space in which these 
three quadratic equations are solvable. Hence (1) is satisfied by the 
coordinates of at least two distinct points P, Q (proper or improper).* 

A linear transformation 



(2) 



evidently transforms the points satisfying (1) into points satisfying 
another equation of the second degree. If, then, (2) is so chosen as 
to transform P and Q into the points (0, 0, 1) and (0, 1, 0) respec- 
tively, (1) will be transformed into an equation which is satisfied by 
the latter pair of points, and which is therefore of the form 

(3) ax* + CjXfo + c z x : x s + c 8 o^ 2 = 0. 

If c l = 0, the points satisfying (3) lie on the two lines 
x l = 0, ax 1 + c 2 # 8 + c s # 2 = ; 

and hence (1) is satisfied by the points on the lines into which these 
lines are transformed by the inverse of (2). If c x 3* 0, the trans- 
formation 

^i = < 

(4) a^- 



* Proposition K 2 has been used merely to establish the existence of points satis- 
fying (1). In case there are proper points satisfying (1), the whole argument can be 
made without K2. 



85,86] PEOBLEMS OF THE SECOND DEGREE 245 

transforms the points (x v x y a? 8 ) satisfying (3) into points (x[ 9 xj, x) 
satisfying 

(5) L - *-& x* + ( 6l xi + c,x>) x> = 0. 



But (5) is in the form which was proved in Theorem 1, Chap. VII, 
to be the equation of a conic. As the points which satisfy (5) are 
transformed by the inverse of the product of the collineations (2) and 
(4) into points which satisfy (1), we see that in all cases (1) repre- 
sents a point conic (proper or improper, degenerate or nondegenerate). 

This gives rise to the two following dual theorems : 

THEOREM 4. Every equation of the form 

a^x* + a^%* + a^xl + 2 a^x^ + 2 a^x z 4- 2 a 25 x 2 x B = 
represents a point conic (proper or improper) which may, liowever, 
degenerate ; and, conversely, every point conic may be represented "by 
an equation of this form. (A, E, P, H , K 2 ) 

THEOEEM 4'. Every equation of the form 

AX + A ^l + A **< + 2 -W 2 + 2 4As + 2 As Vs = 
represents a line conic (proper or improper) which, may, however, de- 
generate; and, conversely, every line conic may be represented "by an 
equation of this form. (A, E, P, H , K 2 ) 

86. Problems of the second degree. "We have seen in 83 that 
any problem of the first degree can be solved completely by means 
of linear constructions ; but that a problem of degree higher than the 
first cannot be solved by linear constructions alone. In regard to 
problems of the second degree in a plane, however, it was seen in 
84 that any such problem may be reduced to the problem of find- 
ing the points of intersection of an arbitrary line in the plane with 
a particular conic in the plane. This result we may state in the 
following form: 

THEOEEM 5. Any problem of the second degree in a plane may be 
solved by linear constructions if the intersections of every line in the 
plane with a single conic in this plane are assumed "known. (A, E, 
P, H , K 2 ) 

In the usual representation of the protective geometry of a real plane by 
means of points, lines, etc., drawn with a pencil, say, on a sheet of paper, the 
linear constructions, as has already been noted, are those that can be per- 
formed with the use of a straightedge alone. It will be shown later that any 



246 GEOMETRIC CONSTEUCTIONS [CHAP.IX 

conic in the real geometry is equivalent protectively to a circle. The instru- 
ment usually employed to draw circles is the compass. It is then clear that 
in this representation any problem of the second degree can be solved ly means of 
a straightedge and compass alone. The theorem just stated, however, shows that 
if a single circle is drawn once for all in the plane, the straightedge alone 
suffices for the solution of any problem of the second degree in this plane. 
The discussion immediately following serves to indicate briefly how this may 
be accomplished. 

We proceed to show how this theorem may be used in the solution 
of problems of the second degree. Any such problem may be reduced 
more or less readily to the first of the following : 

PROBLEM 1. To find the double points of a projectivity on a line of 
which three pairs of homologous points are given. We may assume 




PA' B' C Q B A 

FIG. 98 

that the given pairs of homologous points all consist of distinct points 
(otherwise the problem is linear). In accordance with Theorem 5, 
we suppose given a conic (in a plane with the line) and assume 
known the intersections of any line of the plane with this conic. Let 
be any point of the given conic, and with as center project the 
given pairs of homologous points on the conic (fig. 98). These define 
a projectivity on the conic. Construct the axis of this projectivity 
and let it meet the conic in the points P, Q. The lines OP, OQ then 
meet the given line in the required double points. 

PROBLEM 2. To find the points of intersection of a given line with 
a conic of which five points are given. Let A, B, C 9 D 9 H be the given 
points of the conic. The conic is then defined by the projectivity 
D(A 9 B, C)j^E(A 9 B, 0) between the pencils of lines at D and E. 



SEXTUPLY PEESPECTIVE TBIANGLES 



247 



This projectivity gives rise to a projectivity on the given line of 
which three pairs of homologous points are known. The double 
points of the latter projectivity are the points of intersection of the 
line with the conic. The problem is thus reduced to Problem 1. 

PROBLEM 3. We have seen that it is possible for two triangles in 
a plane to be perspective from four different centers (ef. Ex. 8, p. 105). 
The maximum number of ways in which it is conceivable that two 
triangles may be perspective is clearly equal to the number of per- 
mutations of three things three at a time, ie. six. The question then 
arises, Is it possible to construct two triangles that are perspective from 
six different centers? Let the two triangles be ABC and A!ffG^ and let 

x l = 0, x z = 0, x s = 

be the sides of the first opposite to A, B, C respectively. Let the 
sides of the second opposite to -4', B f , C' respectively be 



, = 0, 



Fas, = 0, 



0. 



The condition for ABC = A f B r C f is that the points of intersection of 



A 



corresponding sides be collinear, ie. 

(i) 



01-1 
-F 1 
-J' 1 



In like manner, the condition for BCA~ A r B r C f is 



(2) 



- 1" I' 
-101 
-V 1 



From these two conditions follows 

-k" U 
-I" 1 
1-10 



which is the condition for CAB = A f B f C f . Hence, if two triangles are 
in the relations ABC = A r r C f and BCA = A'B'C J , they are also in 

A A 

the relation CAB = A r B f C f . Two triangles in this relation are said to 
be triply perspective (cf. Ex. 2, p. 100). The 'domain of rationality 
defined by the data of our problem is clearly 



248 GEOMETRIC CONSTRUCTIONS [CHAP.IX 

Since numbers in this domain may be found which satisfy equations 
(1) and (2), the problem of constructing two triply perspective tri- 
angles is linear. 

The condition for ACB = A'B'C' is 

A 

(3) V-V'=Q. 

If relations (1), (2), and (3) are satisfied, the triangles will be per- 
spective from four centers. Let Jc be the common value of W and l n 
(3), and let I be the common value of V and W (1). Relation (2) then 
gives the condition k* I = 0. The relations 



then define two quadruply perspective triangles. The problem of 
constructing two such triangles is therefore still linear. 

If now we add the condition for CBA = A'B ! C' 9 the two triangles 
will, by what precedes, be perspective from six different centers. The 
latter condition is 

(4) A"J'-Z"=0. 

With the preceding conditions (1), (2), (3) and the notation adopted 
above, this leads to the condition 

L3 _ 78 _ I 

A; (/ = JL 

The equation IP 1 = is, however, reducible in K ; indeed, it is 

equivalent to 

-1=0, # + * + != 0. 



The first of these equations leads to the condition that A ( , J3 f , C r are 
collinear, and does not therefore give a solution of the problem. The 
problem of constructing two triangles that are sextuply perspective 
is therefore of the second degree. The equation 



has two roots w, w* (proper or improper and, in general,* distinct). 
Hence our problem has two solutions. One of these consists of the 



triangles 



8 == 0, X L + w*% 2 + wx^= 0. 



* They can coincide only if the number system is such that 1 + 1 + 1 = 0; e.g. m 
a finite space involving the modulus 3. 



86] 



SEXTUPLY PERSPECTIVE TRIANGLES 



249 



Two of the sides of the second triangle may be improper.* The 
points of intersection of the sides of one of these triangles with the 
sides of the other are the following nine points : 

(0, -1, 1) ( 0, w*, -w) ( 0, w, -w 2 ) 

(5) (-1, 0, 1) (~w\ 0, 1 ) (-to, 0, 1 ) 

(-1, 1, 0) ( w, -1, ) ( < -1, ) 

They form a configuration 



which contains four configurations 




of the kind studied in 36, Chap. IV. All triples of points in the 
same row or column or term of the determinant expansion of their 
matrix are collinear.f If one line is omitted from a finite plane (in 
the sense of 72, Chap. VII) having four points on each line, the 
remaining nine points and twelve lines are isomorphic with this 
configuration. 

EXERCISES 

The problems in a plane given below that are of the second degree are to be solved 
by linear constructions ^ with the assumption that the points of intersection of any line 
in the plane with a given fixed conic in the plane are known : i.e. "with a straight- 
edge and a given circle in tJie plane" 

1. Construct the points of intersection of a given line with a conic deter- 
mined by (i) four points and a tangent through one of them ; (ii) three points 
and the tangents through two of them ; (iii) five tangents. 

2. Construct the conjugate pair common to two involutions on a line. 

3. Given a conic determined by five points, construct a triangle inscribed 
in this conic whose sides pass through three given points of the plane. 

* It may be noted that in the ordinary real geometry two sides of the second 
triangle are necessarily improper, so that in this geometry our problem has no 
real solution. 

t They all lie on any cubic curve of the form x* + xj + xj -j- 3 \xix^x 9 = for 
any value of X, and are, in fact, the points of inflexion of the cubic. This configura- 
tion forms the point of departure for a variety of investigations leading into many 
different branches of mathematics. 



250 GEOMETRIC CONSTRUCTIONS 



4. Given a triangle A^B Z C 2 inscribed in a triangle A l B 1 C l . In how 
many ways can a triangle A 3 B S C 3 be inscribed in A 2 B Z C 2 and circumscribed 
to A^C^t Show that in one case, in which one vertex of A S B^C S may be 
chosen arbitrarily, the problem is linear (cf. 36, Chap. IV) ; and that in 
another case the problem is quadratic. Show that this problem gives all COn- 

rt g 

figurations of the symbol . Give the constructions for all cases (cf. 




S. Ivantor, Sitzungsberichte dermathematisch-naturwissenschaftlichen Classe 
der Kaiserlichen Akademie der Wissenschaften zu Wien, Vol. LXXXIV 
(1881), p. 915). 

5. If opposite vertices of a simple plane hexagon PiP 2 P s P 4 P 5 P G are on 
three concurrent lines, and the lines P-JP^ PJP^ Pf^ are concurrent, then the 
lines P 2 P 3 , PPV P Q Pi are also concurrent, and the figure thus formed is a 
configuration of Pappus. 

6. Show how to construct a simple ?i-point inscribed in a given simple 
n-point and circumscribed to another given simple n-point. 

7. Show how to inscribe in a given conic a simple n-point whose sides 
pass respectively through n given points. 

8. Construct a conic through four points and tangent to a line not meeting 
any of the four points. 

9. Construct a conic through three points and tangent to two lines not 
meeting any of the points. 

10. Construct a conic through four given points and meeting a given line 
in two points harmonic with two given points on the line. 

11. If A is a given point of a conic and A r , Y are two variable points of the 
conic such that AX, A Y always pass through a conjugate pair of a given 
involution on a line I, the line XY will always pass through a fixed point B. 
The line AB and the tangent to the conic at A pass through a conjugate pair 
of the given involution. 

12. Given a collineation in a plane and a line which does not contain a 
fixed point of the collineation ; show that there is one and only one point on 
the line which is transformed by the collineation into another point on the line. 

13. Given four skew lines, show that there are in general two lines which 
meet each of the given four lines ; and that if there are three such lines, there 
is one through every point on one of the lines. 

14. Given in a plane two systems of five points A^A^A^A^A^ and 
BiB^B^B^ ; given also a point X in the plane, determine a point Y such 
that we have X (A^^A^A^A^j- F(B 1 J5 3 B 8 J5 4J B 6 ). In general, there is one 
and only one such point F. Under what condition is there more than one? 
(R. Sturm, Mathematische Annalen, Vol. I (1869), p. 533.*) 

* This is a special case of the so-called problem of projectivity. For references 
and a systematic treatment see Sturm, Die Lehre von den geometrischen Ver- 
wandtschaften, Vol. I, p. 348. 



37] 



USTVABIANTS 



251 



87. Invariants of linear and quadratic binary forms. An expres- 
sion of the form a^ -h a z x 2 is called a linear "binary form in the 
two variables x v x 2 . The word linear refers to the degree in the 
variables, the word Unary to the number (two) of the variables. A 
convenient notation for such a form is a^ The equation 



defines a unique element A of a one-dimensional form in which a 
scale has been established, viz. the element whose homogeneous co- 
ordinates are (x v x 2 ] = (a v a x ). If l x = \x^ + 6 2 # 2 is another linear 
binary form determining the element B, say, the question arises 
as to the condition under which the two elements A and B coincide. 
This condition is at once obtained as the vanishing of the determinant 
A formed by the coefficients of the two forms ; i.e. the elements A 
and B will coincide if and only if we have 



A = 



a, a, 



= 0. 



Now suppose the two elements A and B are subjected to any pro- 
jective transformation II : 



=0. 



The forms a x and l x will be transformed into two forms a x , and & a ', 
respectively, which, when equated to 0, define the points A r 3 B r into 
which the points A, B are transformed by II. The coefficients of 

are readily calculated as 



the forms a x , 9 l x , in terms of those of a x) l 
follows: 



which gives 
Similarly, we find 



=a (aa l 



i -h 



Now it is clear that if the elements -4, B coincide, so also will the 
new elements A 1 , B 1 coincide. If we have A = 0, therefore we should 



also have A' = 
We have 



,' 



= 0. That this is the case is readily verified. 



A' = 



252 GEOMETEIC CONSTRUCTIONS [CHAP, ix 

by a well-known theorem in determinants. Tins relation may also 

be written 

a ft 



A' = 



7 8 



A. 



The determinant A is then a function of the coefficients of the forms 
a *> b*> wikh tt 16 property that, if the two forms are subjected to a lin- 
ear homogeneous transformation of the variables (with nonvamshing 
determinant), the same function of the coefficients of the new forms 
is equal to the function of the coefficients of the old forms multiplied 
by an expression which is a function of the coefficients of the trans- 
formation only. Such a function of the coefficients of two forms is 
called a (simultaneous) invariant of the forms. 

Suppose, now, we form the product a x - l x of the two forms a x , l a ,. 
If multiplied out, this product is of the form 

a = # 2 + 2 a 



Any such form is called a quadratic Unary form. Under Proposi- 
tion K 2 every such form may be factored into two linear factors 
(proper or improper), and hence any such form represents two ele- 
ments (proper or improper) of a one-dimensional form. These two 
elements will coincide, if and only if the discriminant D a = a^ 
a n ' a w ^ ^ e quadratic form vanishes. The condition D a =s there- 
fore expresses a property which is invariant under any projectivity. 
If, then, the form a% be subjected to a projective transformation, the 
discriminant D a , of the new form a r x must vanish whenever D n van- 
ishes. There must accordingly be a relation of the form D a , = k - D a . 
If &l be subjected to the transformation II given above, the coefficients 
new form aj are readily found to be 



2 
(1) < - a^aft + a u (a$ + /3y) + a 



By actual computation the reader may then verify the relation 

a /3 3 
7 8 

The discriminant D a of a quadratic form a 2 is therefore called an 
invariant of the form. 



87] INVARIANTS 253 

Suppose, now, we consider two binary quadratic forms 



Each of these (under JK 2 ) represents a pair of points (proper or im- 
proper). Let us seek the condition that these two pairs be harmonic. 
This property is invariant under protective transformations ; we may 
therefore expect the condition sought to be an invariant of the two 
forms. We know that if a v a 2 are the nonliomogeneous coordinates 
of the two points represented by a* = 0, we have relations 

= ^ss, a +a = 2 *V 

a n " a n 

with similar relations for the nonhomogeneous coordinates & 1? 6 2 of 
the points represented by &| = 0. The two pairs of points a v a z ; l v 6 2 
will be harmonic if we have (Theorem 13, Cor. 2, Chap. VI) 

* = -!. 

L 

This relation may readily be changed into the following : 

which, on substituting from the relations just given, becomes 

1 -2o 1 A.= 0. 



This is the condition sought. If we form the same function of the 
coefficients of the two forms a^ 9 V x * obtained from a*, I* by subjecting 
them to the transformation II, and substitute from equations (1), we 
obtain the relation 



In the three examples of invariants of binary forms thus far 
obtained, the function of the new coefficients was always equal to 
the function of the old coefficients multiplied by a power of the 
determinant of the transformation. This is a general theorem regard- 
ing invariants to which we shall refer again in 90, when a formal 
definition of an invariant will be given. Before closing this section, 
however, let us consider briefly the cross ratio B (a-fl 29 &A) f ^ e 
two pairs of points represented by a% = 0, J = 0. This cross ratio 



254 GEOMETRIC CONSTRUCTIONS [CHAP, ix 

is entirely unchanged when the two forms are subjected to a pro- 
jective transformation. If, therefore, this cross ratio be calculated in 
terms of the coefficients of the two forms, the resulting function of 
the coefficients must be exactly equal to the same function of the 
coefficients of the forms a^ b f x ; the power of the determinant referred 
to above is in this case zero. Such an invariant is called an absolute 
invariant; for purposes of distinction the invariants which when 
transformed are multiplied by a power = of the determinant of 
the transformation are then called relative invariants. 



EXERCISES 

1. Show that the cross ratio Bs (a 1 <7 2 , b^ referred to at the end of the 
last section is 



and hence show, by reference to preceding results, that it is indeed an absolute 
invariant. 

2. Given three pairs of points defined by the three binary quadratic forms 
#2 Q } jj = o, c = ; show that the three will be in involution if we have 

H 12 22 _ 

Hence show that the above determinant is a simultaneous invariant of the 
three forms (cf. Ex. 13, p. 230). 

88. Proposition K^ If we form the product of n linear binary 

forms a x a' x a'J a~ l \ we obtain an expression of the form 



An expression of this form is called a "binary homogeneous form or 
quant of the nth degree. If it is obtained as the product of n linear 
forms, it will represent a set of n points on a line (or a set of n ele- 
ments of some one-dimensional form). 

If it is of the second degree, we have, by Proposition K 2 , that there 
exists an extended space in which it represents a pair of points. At 
the end of this chapter there will be proved the following generali- 
zation of K 2 : 



88,89] PBOPOSITION K n 255 

PROPOSITION K n . If a, a l x) - are a finite number of Unary "homo- 
geneous forms whose coefficients are proper in a space S which satisfies 
Assumptions A, E, P, there exists a space S', of which S is a sub- 
space, in the number system of which each of these forms is a product 
of linear factors. 

As in 85, S' is called an extended space, and elements in S' but 
not in S are called improper elements. Proposition K n thus implies 
that an equation of the form a * = can always be thought of as 
representing n (distinct or partly coinciding) improper points in an 
extended space in case it does not represent any proper points. 

Proposition K n could be introduced as an (not independent) assump- 
tion in addition to A, E, P, and H . Its consistency with the other 
assumptions would be shown by the example of the ordinary com- 
plex space in which it is equivalent to the fundamental theorem 
of algebra, 

89. Taylor's theorem. Polar forms. It is desirable at this point 
to borrow an important theorem from elementary algebra. 

DEFINITION. Given a term Ax" of any polynomial, the expression 
nAx?~ l is called the derivative of Ax? with respect to x t in symbols 

Ax* = nAx?-\ 

The derivative of a polynomial with respect to x i is, by definition, the 

sum of the derivatives of its respective terms. 

o 
This definition gives at once A == 0, if A is independent of # . 

u%i 

Applied to a term of a binary form it gives 

-^ 



With this definition it is possible to derive Taylor's theorem for the 
expansion of a polynomial. *We state it for a binary form as follows : 
Given the binary form 



A 



* For the proof of this theorem on the basis of the definition just given, cf . Fine, 
College Algebra, pp. 460-462. 



256 GEOMETEIC CONSTRUCTIONS [CHAP.IX 

If herein we substitute for x v x 2 respectively the expressions x l + \y v 
t + *# 2 > we obtain, 

/ n o 

+ * 



Here the parentheses are differential operators. Thus 



where -^ means means rH etc. It is readily 

daf ^iL^iJ ^2^1 && 2 |_teJ 

proved for any term of a polynomial (and hence for the polynomial 
itself) that the value of such a higher derivative as d z f/dx z 8x l is 
independent of the order of differentiation ; i.e. that we have 

a 2 / = a 2 / 



DEFINITION. The coefficient of X in the above expansion, viz. 
y l df/8x 1 + 2/2#//d# 2 is called the first polar form of (y l} y z ) with 
respect to f (x v x z ) ; the coefficient of X 2 is called the second ; the 
coefficient of \ n is called the nth polar form of (y v # 2 ) ivitli respect 
to the form f. If any polar form be equated to 0, it represents a set 
of points which is called the first) second, , nth polar of the point 
(y 1? y 3 ) with respect to the set of points represented by f (x v # 2 ) = 0. 

Consider now a binary ioim.f(x v cc 2 ) = and the effect upon it of 
a projective transformation 



If we substitute these values in / (x v o? 3 ), we obtain a new form 
&(8l> %) A point (x v x z ) represented by/(^, o? 2 )= will be trans- 
formed into a point (a?/, x$ represented by the form F(x' v x)= 0. 
Moreover, if the point (y l9 y 3 ) be subjected to the same projectivity, 
it is evident from the nature of the expansion given above that the 
polars of (y l9 y z ) with respect to f(x v # 2 ) = are transformed into 
the polars of (y[, y a ') with respect to F(x[, x) = 0. 



89,90] INVAKLAJSTTS 257 

We may summarize the results thus obtained as follows : 

THEOREM 6. If a Unary form f is transformed ly a protective 
transformation into the form F, the set of points represented byf=Q 
is transformed into the set represented ly F=Q. Any polar of a 
point (y v y^ with respect to f = is transformed into the correspond- 
ing polar of the point (y[, y) with respect to J?= 0. 

The following is a simple illustration of a polar of a point with 
respect to a set of points on a line. 

The form x^ = represents the two points whose nonhomo- 
geneous coordinates are and oo respectively. The first polar of any 
point (y v y^) with respect to this form is clearly y^ & + y^ = 0, and 
represents the point ( y v y z ) ; in other words, the first polar of a 
point P with respect to the pair of points represented by the given 
form is the harmonic conjugate of this point with respect to the pair. 

EXERCISE 

Determine the geometrical construction, of the (n l)th polar of a point 
with respect to a set of n distinct points on a line (cf. Ex. 3, p. 51). 

90. Invariants and covariants of binary forms. DEFINITION. If a 
binary form a = a Q x" + na l x"~~ l x z + + a> n %S be changed by the 



transformation , _ 



_ , 
* 



into a new form A = A^x[ n + A$*~' L xl H ----- h A n x n , any rational 
function I(a Q) a v - , a n ) of the coefficients such that we have 

I(A 4 V ..-, A n ) = <j>(a, A 7, 8) I(a Q> a v , a n ) 
is called an invariant of the form a. A function 

tfK^, -.,.; x lt x z ) 

of the coefficients and the variables such that we have 
C(A A v .>.,A n \ x[, xf) ^ (a, /3, 7, 8) . C(a , a v -.- 9 a n ; x v x^) 

is called a covariant of the form a. The same terms apply to func- 
tions I and C of the coefficients and variables of any finite number 
of binary forms with the property that the same function of the 
coefficients and variables of the new forms is equal to the original 
function multiplied by a function of a, & 7, S only ; they are then 
called simultaneous invariants or covariants. 



258 GEOMETRIC CONSTRUCTIONS [CHAP.IX 

In 87 we gave several examples of invariants of binary forms, 
linear and quadratic. It is evident from the definition that the con- 
dition obtained ly equating to any invariant of a form (or of a 
system of forms] must determine a property of tlie set of points 
represented by the form (or forms) which is invariant under a pro- 
jective transformation. Hence the complete study of the projective 
geometry of a single line would involve the complete theory of invari- 
ants and covariants of binary forms. It is not our purpose in this 
book to give an account of this theory. But we will mention one 
theorem which we have already seen verified in special cases. 

Tlie functions <(#, /3, y, 8) and -^(a, /3, % 8) occurring in the 
definition above are always powers of tlie determinant aS /3<y of 
the projective transformation in question.* 

Before closing this section we will give a simple example of a cova- 
riant. Consider two binary quadratic forms a*, b* and form the new 
quantic 

c a * = K & i - <* A) x i + (<*>& - & A) ^ A + ( a A - a A) x l- 

By means of equations (1), 87, the reader may then verify without 
difficulty that the relation 



holds, which proves C ab to be a covariant. The two points represented 
by C& = are the double points (proper or improper) of the involu- 
tion of which the pairs determined by a* = 0, b* = are conjugate 
pairs. This shows why the form should be a covariant. 

EXERCISE 

Prove the statement contained in the next to the iast sentence. 

91. Ternary and quaternary forms and their invariants. The remarks 
which have been made above regarding binary forms can evidently be 
generalized. A p-aryform of the nth degree is a polynomial of the nth 
degree homogeneous in p variables. When the number of variables is 
three or four, the form is called ternary or quaternary respectively. 
The general ternary form of the second degree when equated to zero 
has been shown to be the equation of a conic. In general, the set of 
points (proper and improper) in a plane which satisfy an equation 

- = 



* For proof, cf M for example, Grace and Young, Algebra of Invariants, pp. 21, 22. 



01] INVARIANTS 259 

obtained by equating to zero a ternary form of the nth degree is 
called an algebraic curve of the nth degree (order). Similarly, the set 
of points determined in space by a quaternary form of the nth degree 
equated to zero is called an algebraic surface of the nth degree. 

The definitions of invariants and covariants of jp-ary forms is pre- 
cisely the same as that given above for binary forms, allowance being 
made for the change in the number of variables. Just as in the 
binary case, if an invariant of a ternary or quaternary form vanishes, 
the corresponding function of the coefficients of any protectively 
equivalent form also vanishes, and consequently it represents a prop- 
erty of the corresponding algebraic curve or surface which is not 
changed when the curve or surface undergoes a projective transforma- 
tion. Similar remarks apply to covariants of systems of ternary and 
quaternary forms. 

Invariants and covariants as defined above are with respect to the 
group of all projective collineations. The geometric properties which 
they represent are properties unaltered by any projective collineatiou. 
Like definitions can of course be made of invariants with respect to 
any subgroup of the total group. Evidently any function of the 
coefficients of a form which is invariant under the group of all col- 
lineations will also be an invariant under any subgroup. But there 
will in general be functions which remain invariant under a subgroup 
but which are not invariant under the total group. These correspond 
to properties of figures which are invariant under the subgroup with- 
out being invariant under the total group. We thus arrive at the 
fundamental notion of a geometry as associated with a given group, 
a subject to which we shall return in detail in a later chapter. 

EXERCISES 

1. Define by analogy with the developments of 80, the n - 1 polars of a 
ternary or quaternary form of the nth degree. 

2. Regarding a triangle as a curve of the third degree, show that the second 
polar of a point with regard to a triangle is the polar line defined on page 46. 

3. Generalize Ex. 2 in the plane and in space, and dualize. 

^11 ^12 ^13 

4. Prove that the discriminant a la a 22 a 23 of the ternary quadratic form 
is an invariant. What is its geometrical interpretation? Cf. Ex., p. 187. 



260 GEOMETEIC CONSTRUCTIONS [CHAP. IX 

92. Proof of Proposition K n . Given a rational integral function 

<f> (x) = a tf + a^"- 1 -{ ----- h a n , a^ 0, 

whose coefficients belong to a given field F, and which is irreducible in 
F, there exists a field F', containing F, in which the equation <f> (x) = 
has a root. 

Let f(x) be any rational integral function of x with coefficients in 
F, and let / be an arbitrary symbol not an element of F. Consider 
the class F^ = [/(/)] of all symbols f(j)> where [/(&)] is the class of 
all rational integral functions with coefficients in F. We proceed to 
define laws of combination for the elements of Fj which render the 
latter a field. The process depends on the theorem * that any poly- 
nomial f(x) can be represented uniquely in the form 



where q(x) and r(x) are polynomials belonging to F, i.e. with 
coefficients in F, and where r(x) is of degree lower than the degree 
n of <j> (x). If two polynomials f l9 / 2 belonging to F are such that 
their difference is exactly divisible by <f>(x), then they are said to be 
congruent modulo <f>(x), in symbols / t = / 2 , mod. <p(x). 

1. Two elements /^j), / 2 (/) of F, are said to be equal, if and only 
if/!^) andj^as) are congruent mod. <p(x). By virtue of the theorem 
referred to above, every element /(/) of F y is equal to one and only 
one element f f (j) of degree less than n. We need hence consider only 
those elements f(j) of degree less than n. Further, it follows from 
this definition that <f> (/) = 0. 

2. If /!(*)+/, (a) s/ t (a), mod * (as), then/ 1 (y)+/ 2 (y)-/ 8 (/). 

3. If /, (x) /, (x) s/ 8 (*), mod. * (x), then / t (/) ./, (/) =/ 3 (/). 
Addition and multiplication of the elements of F y having tlius 

been defined, the associative and distributive laws follow as immedi- 
ate consequences of the corresponding laws for the polynomials /(#). 
It remains merely to show that the inverse operations exist and are 
unique. That addition has a unique inverse is obvious. To prove 
that the same holds for multiplication (with the exception of 0) we 
need only recall f that, since <(#) and any polynomial f(x) have no 
common factors, there exist two polynomials h(x) and Jc(x) with 
coefficients in F such that 

A (a?) -/() + ft (a?). ^ () = !. 
* Fine, College Algebra, p. 156. f Fine, loc. cit., p. 208. 



92] PEOOF OF K n 261 

This gives at once h (/) /(/) = 1, 

so that every element /(/) distinct from has a reciprocal. The class 
F y is therefore a field with respect to the operations of addition and 
mutiplication defined above (cf. 52), such that <(/)= 0. It follows 
at once* that xj is a factor of $(x) in the field F, which is there- 
fore the required field F'. The quotient <f>(x)/(xj) is either irre- 
ducible in Fj, or, if reducible, has certain irreducible factors. If the 
degree of one of the latter is greater than unity, the above process may 
be repeated leading to a field F^,,/' being a zero of the factor in 
question. Continuing in this way, it is possible to construct a field 
^7,y,...,/ m) > w ^ ere fl&^w 1, in which <f>(x) is completely reducible, 
i.e. in which <(#) may be decomposed into n linear factors. This 
gives the following corollary : 

Given a polynomial <f> (x) "belonging to a given field F, there exists a 
field F 7 containing F in which <j> (x) is completely reducible. 

Finally, an obvious extension of this argument gives the corollary : 

Given a finite number of polynomials each of which belongs to a 
given field F, there exists a field F', containing F, in which each of the 
given polynomials is completely reducible. 

This corollary is equivalent to Proposition K n . For if S be any 
space, let F be the number system on one of its lines. Then, as in 
the Introduction (p. 11), F' determines an analytic space which is 
the required space S' of Proposition K n . 

The more general question at once presents itself: Given a field 
F, does there exist a field F, containing F, in which every polynomial 
belonging to F is completely reducible ? The argument used above 
does not appear to offer a direct answer to this question. The ques- 
tion has, however, recently been answered in the affirmative by an 
extension of the above argument which assumes the possibility of 
" well ordering " any class, f 

EXERCISE 

Many theorems of this and other chapters are given as dependent on 
A, E, P, H , whereas they are provable without the use of H . Determine 
which theorems are true in those spaces for which H is false. 

* Fine, College Algebra, p. 1G9. 

t Cf. E. Steinitz, Algebraische Theorie der Kbrper, Journal fur reine u. ange- 
wandte Mathematik, Yol. CXXXVII (1909), p. 167 ; especially pp. 271-286. 



CHAPTER 

PROJECTIVE TRANSFORMATIONS OF TWO-DIMENSIONAL FORMS 

93. Correlations between two-dimensional forms. DEFINITION. A 
projective correspondence between the elements of a plane of points 
and the elements of a plane of lines (whether they be on the same 
or on different bases) is called a correlation. Likewise, a projective 
correspondence between the elements of a bundle of planes and the 
elements of a bundle of lines is called a correlation.^ 

Under the principle of duality we may confine ourselves to a con- 
sideration of correlations between planes. In such a correlation, then, 
to every point of the plane of points corresponds a unique line of the 
plane of lines ; and to every pencil of points in the plane of points 
corresponds a unique projective pencil of lines in the plane of lines. 
In particular, if the plane of points and the plane of lines are on the 
same base, we have a correlation in a planar field, whereby to every 
point P of the plane corresponds a unique line p of the same plane, 
and in which, if Jf, P& P& P are collinear points, the corresponding 
lines p v p 2 , p z) p^ are concurrent and such that 



That a correlation T transforms the points [P] of a plane into the 
lines [p] of the plane, we indicate as usual by the functional notation 



The points on a line I are transformed by F into the lines on a 
point L. This determines a transformation of the lines [I] into the 
points [], which we may denote by T', thus: 

T'(l) = L. 

That F ; is also a correlation is evident (the formal proof may be 
supplied by the reader). The transformation T r is called the correla- 
tion induced by T. If a correlation T transforms the lines [I] of a 

* All developments of this chapter are on the basis of Assumptions A, E, P, and 
H . Cf . the exercise at the end of the last chapter. 

t The terms reciprocity and dudlity are sometimes used in place of correlation. 

262 



93] CORRELATIONS 263 

plane into the points [L] of the plane, the correlation which trans- 
forms the points [ll r ] into the lines [LL f ] is the correlation induced 
by T. If F is induced by F, it is clear that T is induced by P. 
For if we have 



we have also 



and hence the induced correlation of F transforms P 2 into p^ etc. 

That correlations in a plane exist follows from the existence of the 
polar system of a conic. The latter is in fact a projective transforma- 
tion in which to every point in the plane of the conic corresponds a 
unique line of the plane, to every line corresponds a unique point, 
and to every pencil of points (lines) corresponds a projective pencil 
of lines (points) (Theorem 18, Cor., Chap. V). This example is, how- 
ever, of a special type having the peculiarity that, if a point P corre- 
sponds to a line p, then in the induced correlation the line p will 
correspond to the point P ; i.e. in a polar system the points and lines 
correspond doubly. This is by no means the case in every correlation. 

DEFINITION. A correlation in a plane in which the points and 
lines correspond doubly is called a polarity. 

It has been found convenient in the case of a polarity defined by 
a conic to study a transformation of points into lines and the induced 
transformation of lines into points simultaneously. Analogously, in 
studying collineations we have regarded a transformation T of points 
P v P^ P B , P into points _Zf, P^ PJ 9 P^ and the transformation T' of 
the lines P& P 2 P B , P B P^ 1J1> into the lines IJ'JJ', PjPj, #2J', #?' as 
the same collineation. In like manner, when considering a trans- 
formation of the points and lines of a plane into its lines and points 
respectively, a correlation F operating on the points and its induced 
correlation T 1 operating on the lines constitute one transformation of 
the points and lines of the plane. For this sort of transformation we 
shall also use the term correlation. In the first instance a correlation 
in a plane is a correspondence between a plane of points (lines) and 
a plane of lines (points). In the extended sense it is a transformation 
of a planar field either into itself or into another planar field, in 
which an element of one kind (point or line) corresponds to an ele- 
ment of the other kind. 



264 TWO-DIMENSIONAL PBOJECTIVITIES [CHAP.X 

The following theorem is an immediate consequence of the defini- 
tion and the fact that the resultant of any two protective correspond- 
ences is a projective correspondence. 

THEOREM 1. The resultant of two correlations is a projective col- 
lineation, and the resultant of a correlation and a projective collinea- 
tion is a correlation. 

We now proceed to derive tins fundamental theorem for correlations 
between two-dimensional forms. 

THEOREM 2. A correlation letween two two-dimensional primitive 
forms is uniquely defined when four pairs of homologous elements are 
given, provided that no three elements of either form are on the same 
one-dimensional primitive form. 

Proof. Let the two forms be a plane of points a and a plane of 
lines #'. Let C* be any conic in a r , and let the four pairs of homol- 
ogous elements be A 9 3, C, D in a and a', V, c', dl in a!. Let A!, B f , 
C\ D f be the poles of a f , V, c f , d r respectively with respect to C 2 . If 
the four points A, B, C, D are the vertices of a quadrangle and the 
four points A', B r , C r , D* are likewise the vertices of a quadrangle 
(and this implies that no three of the lines a r , V, c r , d f are concurrent), 
there exists one and only one collineation transforming A into A r , B 
into B f , C into C 1 , and D into D f (Theorem 18, Chap. IV). Let tins 
collineation be denoted by T, and let the polarity defined by the conic 
C 2 be denoted by P. Then the projective transformation F which is 
the resultant of these two transforms A into a r , B into V, etc. More- 
over, there cannot be more than one correspondence effecting this 
transformation. For, suppose there were two, F and F r Then the 
projective correspondence I^" 1 F would leave each of the four points 
A, B, C, D fixed; i.e. would be the identity (Theorem 18, Chap. IY). 
But this would imply F x = F. 

THEOREM 3. A correlation which interchanges the vertices of a 
triangle with the opposite sides is a polarity. 

Proof. Let the vertices of the given triangle be A, B, C, and let 
the opposite sides be respectively a, I, c. Let P be any point of the 
plane ABC which is not on a side of the triangle. The line p into 
which P is transformed by the given correlation F does not, then, pass 
through a vertex of the triangle ABC. The correlation F is deter- 
mined by the equation F (ABCP) = alcp, and, by hypothesis, is such 



93] 



COEKELATIONS 



265 



that F (ale) ABC. The points [Q] of c are transformed into the 
lines [g] on C, and these meet c in a pencil [Q r ] projective with [Q] 
(fig. 99). Since A corresponds to B and to A in the projectivity 
[Q] ~^[Q f ], this projectivity is an involution L The point Q in which 




CP meets c is transformed by F into a line on the point cp\ and 
since Q and cp are paired in I, it follows that cp is transformed 
into the line CQ Q CP. In like manner, Ip is transformed into BP. 
Hence p = (cp, lp] is transformed into P = (CP, BP). 

THEOBEM 4. Atiy projective collineatwn, II, m a plane, <x, is the 
product of two polarities. 

Proof, Let Aa be a lineal element o! a, and let 
H (4a) =* jl V, n (4W) A" a". 

Unless II is perspective, Aa may be so chosen that A 9 A f , A n are not 
collinear, aa r a fr are not concurrent, and no line of one of the three 
lineal elements passes through the point of another. In this case there 
exists a polarity P such that J?(AA'A rr ) = a"a'a, namely the polarity- 
defined by the conic with regard to which AA"(aa ff ) is a self-polar tri- 
angle and to which o! is tangent at A f . If II is perspective,,the existence 
of P follows directly on choosing Aa, so that neither A nor a is fixed. 
We then have 



and hence the triangle AA[(aaf) is self-reciprocal. Hence (Theorem 3) 
PIT ss Pj is a polarity, and therefore II = PP r 



266 TWO-DIMENSIONAL PBOJECTIVITIES [CHAI>. x 

94. Analytic representation of a correlation between two planes. 
Bilinear forms. Let a system of simultaneous point-and-line coordi- 
nates be established in a planar field. We then have 

THEOREM 5. Any correlation in a plane is given as a transforma- 
tion of points into lines by equations of the form 



(1) pul = a^ + 



where the determinant A of the coefficients a l} is different from zero. 
Conversely, every transformation of this form in which the determinant 
A is different from zero represents a correlation. 

The proof of this theorem is completely analogous to the proof of 
Theorem 8, Chapter VII, and need not be repeated here. 
As a corollary we have 

COROLLARY 1. The transformation pu[~x v pu^x^ pu! } =x s in 
a plane represents a polarity in which to every side of the triangle of 
reference corresponds the opposite vertex. 

Also, if (u[, ii' 2) u^) be interpreted as line coordinates in a plane 
different from that containing the points (x lf # 2 , x s ) (and if the num- 
ber systems are so related that the correspondence X 1 = X between 
the two planes is projective), we have at once 

COROLLARY 2. The equations of Theorem 5 also represent a correla- 
tion between the plane of (x v w z > #3) and the plane of (%', ul, v. 8 '). 

Returning now to the consideration of a correlation in a plane 
(planar field), we have seen that the equations (1) give the coordi- 
nates (u[, tf a ', ui) of the line u'= T (X), which corresponds to the 
point -3T= (x v % z , a? 8 ). By solving these equations for x i9 



we obtain the coordinates of Z= T" 1 (vf) in terms of the coordinates 
u[ of the line to which Xis homologous in the inverse correlation T"\ 
If, however, we seek the coordinates of the point -5T' = T (u] which 
corresponds to any line u in the correlation T, we may proceed as 
follows : 



COKBELATI02TS 267 

Let the equation of the point JT' = (&[, x! 2 , x' A ) in line coordinates be 

= 0. 



Substituting in this equation from (1) and arranging the terms as a 
linear expression in x v x 2) # 3 , 



= 0, 
we readily find 



(3) 



The coordinates of X ! in terms of the coordinates of u are then 
given by 

(4) 



This is the analytic expression of the correlation as a transformation 
of lines into points ; i.e. of the induced correlation of F. These equa- 
tions clearly apply also in the case of a correlation between two 
different planes. 

It is perhaps well to emphasize the fact that Equations (1) express T as a 
transformation of points into lines, while Equations (4) represent the induced 
correlation of lines into points. Since we consider a correlation as a trans- 
formation of points into lines and lines into points, T is completely represented 
by (1) and (4) taken together. Equations (2) and (3) taken together repre- 
sent the inverse of T. 

Another way of representing F analytically is obtained by observ- 
ing that the point (x v x 2 , # 8 ) is transformed by F into the line whose 
equation in current coordinates (x[, x, x) is 



or, 

(5) (a^ + a 12 2 

= 0. 



The left-hand member of (5) is a general ternary bilinear form. We 
have then 

COROLLARY 3. Any ternary bilinear form in which the determinant 
A is different from zero represents a correlation in a plane. 



268 TWO-DIMENSIONAL PEOJECTIVITIES [CHAP.X 

95. General projective group. Representation by matrices. The 

general protective group of transformations in a plane (which., under 
duality, we take as representative of the two-dimensional primitive 
forms) consists of all projective collineations (including the identity) 
and all correlations in the plane. Since the product of two collinea- 
tions is a collineation, the set of all projective collineations .forms a 
subgroup of the general group. Since, however, the product of two 
correlations is a collineation, there exists no subgroup consisting 
entirely of correlations.* 

According to the point of view developed in the last chapter, the 
projective geometry of a plane is concerned with theorems which 
state properties invariant under the general projective group in the 
plane. In particular, the principle of duality may be regarded as a 
consequence of the presence of correlations in this group. 

Analytically, collineations and correlations may be regarded as 
aspects of the theory of matrices. The collineation 



may be conveniently represented by the matrix A of the coefficients a t/ : 



The product of two collineations A = (a v ) and B = (&J is then given 
by the product of their matrices : 




the element of the ith row and the yth column of the matrix BA 
being obtained by multiplying each element of the ith row of B by the 
corresponding element of the^th column of A and adding the products 
thus obtained. It is clear that two collineations are not in general 
commutative. 

* A polarity and the identity fcym a group ; but this forms no exception to the 
statement just made, since the identity must be regarded as a collineation. 



MATEICES 269 

Of the two matrices 

/ a il <*12 13\ /*n u ai\ 



^31 ^32 <W \^13 ^23 ^33/ 



either of which is obtained from the other by interchanging rows and 
columns, one is called the conjugate or transposed matrix of the 
other. The matrix 

/ A A 4 \ 

I -fX-. i L n i uTjL - I 




is called the adjoint matrix of the matrix A. The adjoint matrix is 
clearly obtained by replacing each element of the transposed matrix 
by its cof actor. Equations (2) of 67 show that the adjoint of a 
given matrix represents the inverse of the collineation represented "by 
the given matrix. Indeed, by direct multiplication, 



i\ 



Kl ^22 ^23 MlS 4*2 As H 




and the matrix just obtained clearly represents the identical col- 
lineation. Since, when a matrix is thought of as representing a 
collineation, we may evidently remove any common factor from all 
the elements of the matrix, the latter matrix is equivalent to the 
so-called identical matrix,* 

II 0\ 
1 . 

\0 I/ 

Furthermore, Equations (3), 67, show that if a given matrix 
represents a collineation in point coordinates, the conjugate of the 
adjoint matrix represents the same collineation in line coordinates. 
Also from the representation of the product of two matrices just 
derived, ' follows the important result: 

The determinant of the product of two matrices (collineations) is 
equal to the product of the determinants of the two matrices (col- 
lineations). 

* In the general theory of matrices these two matrices are not, however, re- 
garded as the same. It is only the interpretation of them as collineations which 
renders them equivalent. 



270 TWO-DIMENSIONAL PEOJECTIVITIES [CHAP.X 

From what has just been said it is clear that a matrix does not 
completely define a collineatiou, unless the nature of the coordinates 
is specified. If it is desired to exhibit the coordinates in the nota- 
tion, we may write the collineation ?/ = 20 v a ; in the symbolic form 



The matrix (a v ) may then be regarded as an operator transforming 
the coordinates x=s(x l9 # a , # 8 ) into the coordinates #' = (#/, h # 8 ). If 
we place a v = a jl9 the matrix conjugate to (aj is (a,,). Also by plac- 
ing JT = -4,,, the adjoint matrix of (c^) is ( Jj. The inverse of the 
above collineation is then written 



Furthermore, the collineation #'= (a v )# is represented in line coordi- 
nates by the equation 



This more complete notation will not be found necessary in gen- 
eral in the analytic treatment of collineations, when no correlations 
are present, but it is essential in the representation of correlations 
by means of matrices. 

The correlation (1) of 94 may clearly be represented symbolically 

by the equation 

'(0 t ,)0, 

where the matrix (a^) is to be regarded as an operator transforming 
the point & into the line vf. This correlation is then expressed as a 
transformation of lines into points by 



The product of two correlations u f ~ (a iy ) x and %'= (B v ) x is there- 
fore represented by 

d**(Bv)(av)x 

(d. Equations (4), 94), or by 



Also, the inverse of the correlation u 1 = (a ) x is given by 

^ = (l v )< 
or by 



TYPES OF COLLISrEATIOlSrS 



271 



EXERCISE 

Show that if [II] is the set of all collineations in a plane and I\ is any 
correlation, the set of all correlations in the plane is [Orj, so that the two 
sets of transformations [II] and [Iirj comprise the general projective group 
in the plane. By virtue of this fact the subgroup of all projective collineations 
is said to be of index in the general projective group.* 

96. Double points and double lines of a collineation in a plane. 

Referring to Equations (1) of 67 we see that a point (x v x z , x s ) 
which is transformed into'itself by the collineation. (1) must satisfy 
the equations 



which, by a simple rearrangement, may be written 



(1) 



If a point (x v C 2 , # 3 ) is to satisfy these three equations, the deter- 
minant of this system of equations must vanish ; i.e. p must satisfy 
the equation 



(2) 



= 0. 



Tliis is an equation of the third degree in p, which cannot have more 
than three roots in the number system of our geometry. 

Suppose that /> t is a root of this equation. The system of equa- 
tions (1) is then consistent (which means geometrically that the 
three lines represented by them pass through the same point), and 
the point determined by any two of them (if they are independent, 
i.e. if they do not represent the same line) is a double point Solving 
the first two of these equations, for example, we find as the coordi- 
nates (% v x z> o? 3 ) of a double point 



(3) iv-v^s 8 * 



2 Pi 



*u-/>i 

& 21 



*13 



* A subgroup [II] of a group is said to he of index n, if there exist n 1 trans- 
formations T l (i ss 1, 2, . . . n - l),-such that the n - 1 sets [nTf] of transformations 
together with the set [II] contain all the transformations of the group, while no two 
transformations within the same set or from any two sets are identical. 



272 TWO-DIMENSIONAL PEOJECTIVITIES [CHAP, x 

which represent a unique point, unless it should happen that all the 
determinants on the right of this equation vanish. Leaving aside 
this possibility for the moment, we see that every root of Equation 
(2), which is called the characteristic equation of the collineatioii (or 
of the representative matrix), gives rise to a unique double point. 
Moreover, every double point is obtainable in this way. This is the 
analytic form of the fact already noted, that a collineation which is 
not a Jiomology or an elation cannot have more than three do Me 
points, unless it is the identical collineation. 

If, however, all the determinants on the right in Equations (3) 
vanish, it follows readily that the first two of Equations (1) represent 
the same line. If the determinants formed analogously from the last 
two equations do not all vanish, we again get a unique double point ; 
but if the latter also vanish, then all three of the equations above 
represent the same line. Every point of this line is then a double point, 
and the collineation must be a homology or an elation. Clearly this 
can happen only if p l is at least a double root of Equation (2) ; for 
we know that a perspective collineation cannot have more than one 
double point which is not on the axis of the collineation. 

A complete enumeration of the possible configurations of double 
points and lines of a collineation can be made by means of a study 
of the characteristic equation, making use of the theory of elementary 
divisors.* It seems more natural in the present connection to start 
with the existence of one fixed point (Proposition K 8 ) and discuss 
geometrically the cases that can arise. 

By Theorem 4 a collineation is the product of two polarities. Hence, 
any double point has the same polar line in both polarities, and that 
polar line is a double line. Hence the invariant figitre of douUc points 
and lines is self-dual. 

Four points of the plane, no three of which are collineor, cannot 
be invariant unless the collineation reduces to the identity. If three 
noncollinear points are invariant, two cases present themselves. If 
the collineation reduces to the identity on no side of the invariant 
triangle, the collineation is of Type I (cf. 40, Chap. IV). If the 
collineatioia is the identity on one and only one side of the invariant 
triangle, the collineation is of Type JF.f If two distinct points are 

* Cf. Baeher, Introduction to Higher Algebra, Chaps. XX and XXI. 

t If it is the identity on more than one side, it is the identical collineation. 



no] 



TYPES OF COLLDTEATIONS 



273 



invariant, but no point not on the line I joining these two is invariant, 
two possibilities again arise. If the collineation does not leave every 
point of this line invariant, there is a unique other line through one 
of these points that is invariant, since the invariant figure is self-dual. 
The collineation is then of Type II. If every point of the line is 
invariant, on the other hand, all the lines through a point of the 
line I must be invariant, since the figure of invariant elements is 
self-dual. The collineation is then of Type V. 

If only one point is fixed, only one line can be fixed. The collinea- 
tion is then parabolic both on the line and on the point, and the 
collineation is of Type III. 

We have thus proved that every collineation different from the 
identity is of one of the five types previously enumerated. Type I 
may be represented by the symbol [1, 1, 1], the three 1's denoting 
three distinct double points. In Type IV there are also three distinct 
double points, but all points on the line joining two of them are fixed 
and Equation (1) has one double root. Type 7Fis denoted by [(1, 1), 1]. 
In Type //, as there are only two distinct double points, Equation 
(1) must have a double root and one simple root. This type is ac- 
cordingly denoted by the symbol [2, 1], the 2 indicating the double 
point corresponding to the double root. Type Fis then naturally repre- 
sented by [(2, 1)], the parentheses again indicating that every point 
of the line joining the two points is fixed. Type III corresponds to a 
triple root of (1), and may therefore be denoted by [3]. We have 
then the following : 

THEOREM 6. Every protective collineation in a plane is of one of 
the following five types : 



[1,1,1] 


[(1.1). 1] 






[2,1] 


[(2, 1)] 






[3] 



In this table the first column corresponds to three distinct roots 
of the characteristic equation, the second column to a double root, 
the third column to a triple root. The first row corresponds to the 
cases in which there exist at least three double points which are 



274 TWO-DIMENSIONAL PBOJEOTIVITIES [CHAP, x 

not collinear ; tlie second row to the case where there exist at least 
two distinct double points and all such points are on the same line ; 
the third row to the case in which there exists only a single -double 
point. 

With every collineation in a plane are associated certain projec- 
tivities on the invariant lines and in the pencils on the invariant 
points. In case the collineation is of Type /, it is completely deter- 
mined if the projectivities on two sides of the invariant triangle are 
given. There must therefore be a relation between the projectivities 
on the three sides of the invariant triangle (cf. Ex. 5, p. 276). In a 
collineation of Type II the projectivity is parabolic on one of the 
invariant lines but not on the other. The point in which the two 
invariant lines meet may therefore be called singly parabolic. The 
collineation is completely determined if the projectivities on the 
two invariant lines are given. In a collineation of Type /// the pro- 
jectivity on the invariant line is parabolic, as likewise the projectivity 
on the invariant point. The fixed point may then be called doubly 
parabolic. The projectivities on the invariant lines of a collineation 
of Type V are parabolic except the one on the axis which is the 
identity. The center is thus a singly parabolic point. In the table 
of Theorem 6 the symbols 3, 2, and 1 may be taken to indicate 
doubly and singly and nonparabolic points respectively.* 

We give below certain simple, so-called canonical forms of the 
equations defining collineations of these five types. 

Type L Let the invariant triangle be the triangle of reference. 
The collineation is then given by equations of the form 



in which a lv # 22 , & 33 are the roots of the characteristic equation and 
must therefore be all distinct. 

Type IV, Homology. If the vertices of the triangle of reference 
are taken as invariant points, the equations reduce to the form written 
above; but since one of the lines 3^= 0, # 3 = 0, # 3 = is pointwise 

* For a more detailed discussion of collineations, reference may "be made to 
Newson, A New Theory of Collineations, etc., American Journal of Mathematics, 
Vol. XXIV, p. 109. 



J] TYPES OF COLLIKEATIONS 275 

invariant, we must have either a 22 = a ss or a ss = a n or & n = a Z2 . Thus 
the homology, may be written 



A harmonic homology or reflection is obtained by setting & 83 = 1. 

Type IL The characteristic equation has one double root, p^ = p , 
say, and a simple root /3 3 . Let the double point corresponding to 
Pi Pz ^ e ^i (0> 0, 1), let the double point corresponding to p z be 
?7g = (1, 0, 0), and let the third vertex of the triangle of reference 
be any point on the double line u z corresponding to /> 3 , which line 
will pass through the point U r The collineation is then of the form 



since the lines ^=0 and a? 2 = are double lines and (1, 0, 0) is a 
double point. The characteristic equation of the collineation is clearly 



and since this must have a double root, it follows that two of the 
numbers a n> 22 , a 33 must be equal. To determine which, place 
p = a 00 ; using the minors of the second row, we find, as coordinates 
of the corresponding double point, 

(0, (a n - a 23 ) (a 22 - a S3 ), a 32 (a n - a M ) ), 

which is U v and hence we have #. 22 =ft 83 . The collineation then is 
of Type II, if a xl ^ a 2Z . Its equations are therefore 



where a 32 3=- and a n = a 22 . 

jf^e /JJ. The characteristic equation has a triple root, p 1 = p 2 = P z 
say. Let U : = (0, 0, 1) be the single double point, and the line x l = be 
the single double line. With this choice of coordinates the collineation 

has the form , 

P x i a 'n^i> 



276 TWO-DIMENSIONAL PROJECTIVITIES [OHAP.X 

By writing the characteristic equation we find, in view of the fact 
that the equation has a triple root, that a n =^ 25J =^ 33 . The form of 
the collineation is therefore 



where the numbers a 21 , a sz must be different from 0. * 

Type F, Elation. Choosing (0, 0, 1) as center and ^=0 as axis, 
the equations of the collineation reduce to the form given for Type ///, 
where, however, a 33 must be zero in order that the line 0^= be 
pointwise invariant. The equations for Type II also yield an elation 
in case & n = a 22 . Thus an elation may be written 



EXERCISES 

1. Determine the collineation which transforms the points A = (0, 0, 1), 
B = (0, 1, 0), C= (1, 0, 0), D = (1, 1, 1) into the points E, C, D, A respec- 
tively. Show that the characteristic equation of this collineation is (p I) 
(p 2 + 1) = 0, which in any field has one root. Determine the double point 
and double line corresponding to this root. Assuming the field of numbers to 
be the ordinary complex field, determine the coordinates of the remaining two 
double points and double lines. Verify, by actually multiplying the matrices, 
that this collineation is of period 4 (a fact which is evident from the defini- 
tion of the collineation). 

2. With the same coordinates for A, JS, C, D determine the collincation 
which transforms these points respectively into the points Z>, A , 7>, C. The 
resulting collineation must, from this definition, be a homology. Why? De- 
termine its center and its axis. By actual multiplication of the matrices 
verify that its square is the identical collineation. 

3. Express each of the collineations in Exs. 1 and 2 in terms of line 
coordinates. 

4. Show that the characteristic cross ratios of the one-dimensional projec- 
tivities on the sides of the invariant triangle of the collineation x'z=ax v 
x^ = bx z , z = c# 3 are the ratios of the numbers , /;, c. Hence show that tho 
product of these cross ratios is equal to unity, the double points being taken 
around the triangle in a given order. 

5. Prove the latter part of Ex. 4 for the cross ratios of tho projectivities 
on the sides of the invariant triangle of any collineation of Type /. 



TYPES OF COLLINEATIOjtfS 277 

6. Write the equations of a collineation of period 3;4;5;-..;n;-... 

7. By properly choosing the system of nonhomogeneous coordinates any 
collineation of Type / may be represented by equations x' = ax, y' = ly. The 
set of all collineations obtained by giving the parameters a, b all possible 
values forms a group. Show that the collineations #' = axj ?/' = a r y, where r 
is constant for all collineations of the set, form a subgroup. Show that every 
collineation of this subgroup leaves invariant every curve whose equation is 
y = cx r 9 where c is any constant. Such curves are called path curves of the 
collineations. 

8. If P is any point of a given path curve, p the tangent at P, and 
A y B, C the vertices of the invariant triangle, then & (p, PA, PJ3, PC) is a 
constant. 

9. For the values r = 1, 2, -| the path curves of the collineations of the 
subgroup described in Ex. 7 are conies tangent to two sides of the invariant 
triangle at two vertices. 

10. If r = 0, the subgroup of Ex. 7 consists entirely of homologies. 

11. Prove that any collineation of Type / may be expressed in the form 

x' = k(ax + by), 
/ = *(& -ay), 
with the restriction a 2 + b 2 = 1. 

12. Prove that any collineation can be expressed as a product of collinea- 
tions of Type /. 

13. Let the invariant figure of a collineation of Type // be A, JB, I, m, 
where I = AB, B = Im. The product of such a collineation by another of 
Type // with invariant figure A', B, Z, m' is in general of Type //, but may 
be of Types /// or 1 V. Under what conditions do the latter cases arise ? 

14. Using the notation of Ex. 13, the product of a collineation of Type II 
with invariant figure A, J5, Z, m by one with invariant figure -4,,J5', Z, m' is 
in general of Type II, but may be of Types /// or IV. Under what conditions 
do the latter cases arise ? 

15. Prove that any collineation can be expressed as a product of collinea- 
tions of Type II. 

16. Two collineations of Type III with the same invariant figure are not 
in general commutative. 

17. Any protective collineation can be expressed as a product of collinea- 
tions of Type ///. 

18. If II is an elation whose center is C, and P any point not on the 
axis, then P and C are harmonically conjugate with respect to n~ 1 ( J P) 
andH(P). 

19. If two coplanar conies are projective, the correspondence between the 
points of one and the tangents at homologous points of the other determines 
a correlation* 

20. If in a collineation between two distinct planes every point of the 
line of intersection of the planes is self-corresponding, the planes are per- 
spective. 



278 TWO-DIMENSIONAL PROJEOTIV1TIES [CHAP, x 

21. In nonhomogeneous coordinates a collineation of Type / with fixed 
points (o lf a a ), (b l9 6 a ) (c v c 2 ) may be written 





a; y 1 x y 1 
! #2 1 t ! a 2 1 cr 2 

(? 1 c . 2 ^ fc'c l c ~ I /j'^ 


Type II may be w 


x y 1 ' ^ a? y 1 
a x ij 1 1 ! n a 1 1 

b 1% 1 /J />i & 2 1 ^* 

Cj c 2 1 k' ci c z 1 /t' 
ritten 

ar y 1 2: y 1 
i /i ft ~\ ft 

/^ /; 3 1 ^j />! ^ 2 1 AV> a 


and Type III may 

f __ 


a? y 1 ?/ a,- y 1 
i i 11 
#! 2 1 1 rtj_ a 2 J. J. 

Z^ J 2 1 A' bi h n 1 X: 

! 5 2 f *! 4 t 

be written 
2- y 1 

w i 20 s (a* 2 + 2j8J)a 1 + 2as 1 ? + W' 1 


a; y 1 

^ x w a 0^ + 2^ 
a; y 1 

?^ w a (a 2 + 2^) 3 + 2a.V + ?a> 


y x y 1 

&* 1 3 o 

u\ w 9 ai 2 +2/fr 



97. Double pairs of a correlation. We inquire now regarding the 
existence of double pairs of a correlation in a plane. By a double pair 
is meant a point X and a line u such that the correlation transforms 
X into u and also transforms u into X; in symbols, if F is the cor- 
relation, such that r(Z)=w and F(w)=:X. We have already seen 
(Theorem 3) that if the vertices and opposite sides of a triangle are 
double pairs of a correlation, the correlation is a polarity. 

We may note first that the problem of finding the double pairs of 
a correlation is in one form equivalent to finding the double elements 



07] DOUBLE PAIRS OF A CORRELATION 279 

of a certain eollineation. In fact, a double pair X, u is such that 
F(JL")=^ and r 2 (J5T) = r(w)=JT, so that the point of a double pair 
of a correlation F is a double point of the coUineation F 2 Similarly, 
it may be seen that the lines of the double pairs are the double lines 
of the eollineation F 2 . It follows also from these considerations that 
F is a polarity, if F 2 is the identical eollineation. 

Analytically, the problem of determining the double pairs of a 
correlation leads to the question : For what values of (x v # 2 , # 8 ) are 
the coordinates 



of the line to which it corresponds proportional to the coordinates 

a x + 



of the line which corresponds to it in the given correlation ? If p is 
the unknown factor of proportionality, this condition is expressed by 
the equations 

K - P a u) x i + (* - P a n) + ( a is - P a ^ x* = 3 

pa S2 ) ^ 8 = 0, 



which must be satisfied by the coordinates (x lf # 2 , a? 8 ) of any point 
of a double pair. The remainder of the treatment of this problem is* 
similar to the corresponding part of the problem of determining the 
double elements of a eollineation ( 96). The factor of proportionality 
p is determined by the equation 



*p a *l 



(2) 



which is of the third degree and has (under Proposition K 2 ) three 
roots, of which one is 1, and of which the other two may be proper 
or improper. Every root of this equation when substituted for p in 
(1) renders these equations consistent. The coordinates (x v x v # 8 ) 
are then determined by solving two of these. 

If the reciprocity in question is a polarity, Equations (1) must be 
satisfied identically, i.e. for every set of values (x v # 2 , x 3 ). This would 
imply that all the relations 

S-/>^-0 (*,/=!, 2, 3) 

are satisfied. 



280 TWO-DIMENSIONAL PEOJECTIVITIES [CHAP. X 

Let us suppose first that at least one of the diagonal elements of the 
matrix of the coefficients (a i} ) be different from 0. If this be a lv the 
relation & u -/^n O gives at once /> = !; and this value leads at 
once to the further relations 

% = V ft/ = 1,2, 3). 

The matrix in question must then be symmetrical. If, on the other 
hand, we have a u = a^ = a^ = 9 there must be some coefficient a^ 
different from 0. Suppose, for example, # 12 * 0. Then the Delation 
a 12 Aa 21 = shows that neither k nor # 21 can be 0. The substitution 
of one in. the other of the relations a 12 =&& 21 and a^ Jea^ then gives 
Jc* = 1, or Jc = 1. The value 7c = 1 again leads to the condition that 
the matrix of the coefficients be symmetrical The value & = 1 
gives a K = 0, and # y = <^, which would render the matrix skew 
symmetrical. The determinant of the transformation would on this 
supposition vanish (since every skew-symmetrical determinant of odd 
order vanishes), which is contrary to the hypothesis. The value 
Jc BBS 1 is therefore impossible. We have thus been led to the fol- 
lowing theorem: 

THEOREM 7. TJie necessary and sufficient condition that a reci- 
procity in a plane le a polarity is that the matrix of its coefficients 
le symmetrical. 

If the coordinate system is chosen so that the point which corre- 
sponds to p =s 1 in Equation (2) is (1, 0, 0), it is clear that we must 
have a zl a 12 and # 81 = fl ia . If the line corresponding doubly to 
(1, 0, 0) does not pass through it, the coordinates [1, 0, 0] may be 
assigned to this line. The equations of the correlation thus assume 
the form 

(3) 



and Equation (2) reduces to 



(4) 



= 0. 



The roots, other than 1, of this equation clearly correspond to points 
on [1, 0, 0], Choosing one of these points (Proposition K 2 ) as (0, 0, 1), 
we have either # 28 = a S2 , which would lead to a polarity, or # 88 = 0. 



97] DOUBLE PAIES OF A COEBELATION 281 

In the latter case it is evident that (4) has a double root if # 32 = a 23 , 
but that otherwise it has two distinct roots. Therefore a correlation 
in which (1, 0, 0) and [1, 0, 0] correspond doubly, and which is not 
a polarity, may be reduced to one of the three forms : 



(a = 0, 1 = 0) 



K= - 



The squares of these correlations are collineations of Types /, //, IV 
respectively. 

If the line doubly corresponding to (1, 0, 0) does pass through it, 
the coordinates [0, 1, 0] may be assigned to this line, and the equa- 
tions of the correlation become 



Equation (2) at the same time reduces to 

(i-p)*-o, 

and the square of the correlation is always of Type III. There are 
thus five types of correlations, the polarity and those whose squares 
are collineations of Types /, II, III, IV. 

EXERCISES * 

1. The points which lie upon the lines to which they correspond in a cor- 
relation form a conic section C 2 , and the lines which lie upon the points to 
which they correspond are the tangents to a conic K 2 . How are C 2 and E? 
related, in each of the five types of correlations, to one another and to the 
doubly corresponding elements ? 

* On the theory of correlations see Seydewitz, Archiv der Mathematik, 1st series, 
Vol. VIII (1846), p. 32 ; and Schr5ter, Journal fur die reine und angewandte Mathe- 
matik, Vol. LXXVII (1874), p. 105. 



282 TWO-DIMENSIONAL PROJECTIVITIES [CHAP. X 

2. If a line a does not lie upon the point A' to which it corresponds in a 
' correlation, there is a projectivity between the points of a and tho points in 

which their corresponding lines meet a. In the case of a polarity this pro- 
jectivity is always an involution. In any other correlation the linos upon 
which this projectivity is involutoric all pass through a unique fixed point 0. 
The line o having the dual property corresponds doubly to 0. The doublo 
points of the involutions on the lines through are on the conic ( '-, and th< k 
double lines of the involutions on the points of /C 2 are tangent to A" 2 . and o 
are polar with respect to C 2 and K z . If a correlation determines involutions 
on three nonconcurrent lines, it is a polarity. 

3. The lines of K 2 through a point P of C 2 are the line which is transformed 
into P and the line into which P is transformed by the given correlation. 

4. In a polarity C 2 and K* are the same conic. 

5. A necessary and sufficient condition that a collineation be the product of 
two reflections is the existence of a correlation which is left invariant by the 
collineation.* 

98. Fundamental conic of a polarity in a plane. We have just 
seen that a polarity in a plane is given by the equations 



(1) 



DEFINITION. Two homologous elements of a polarity in a plane are 
called pole and polar, the point being the pole of the line and the 
line being the polar of the point. If two points are so situated that 
one is on the polar of the other, they are said to be conjugate. 

The condition that two points in a plane of a polarity he conju- 
gate is readily derived. In fact, if two points P = (M V x z , $ a ) and 
p* '=(#/, # a ', x) are conjugate, the condition sought is simply that 
the point P ! shall be on the line p f = \u[, u[, M 8 '], the polar of J\\ i.e. 
u[x{ + u&l + u&l = 0. Substituting for u[, ut, %[ their values in 
terms of x v x^ x s from (1), we obtain the desired condition, via. : 

(2) a^xl + 



As was to be expected, this condition is symmetrical in the coordi- 
nates of the two points P and P f . By placing x{ = as, we obtain the 

* This is a special case of a theorem of Dunham Jackson, Transactions of the 
American Mathematical Society, Vol. X (1909), p. 479, 



POLAR SYSTEM 283 

condition that the point /' be self -conjugate, ie. that it be on its polar. 
We thus uUuiu the result; 

THKOUBM 8. 77/0 sdfaonjugttte print* of the polarity (1) are on 
the- MUM whoso equation is 



ft, r<f + <typ* + a^l + 2 a^v,, + 2 fl w avcfr + 2 a 2g aj a a? 8 = ; 
( 9 conversely, every point of this conic is sdf -conjugate. 

This conic is called the fundamental conic of the polarity. All of 
it points may be improper, but it can never degenerate, for, if so, 
the determinant v | would have to vanish (cf. Ex., p. 187). By 
duality we obtain 

THEOREM 8 r . The self-c,o>iju<jate lines of tlie polarity (1) are lines 
of the, conic 

(4) A^ 4- A^l 4- J S X + 2' A^u^ 4- 2 A^n^ 4- 2 ^ 23 w 3 ^ 3 = ; 
<mrf, emwrMly, every line of this conic is self-conjugate. 

Every point A' of the conic (3) corresponds in the polarity (1) to 
the tangent to (3) at Jif. For if not, a point A of (3) would be polar 
to a line a through A and meeting (3) also in a point B> B would 
then be polar to a line I through B, and hence the line a^AB 
would, by the definition of a polarity, be polar to ab = B. This would 
require that a correspond both to A and to B. 

If now we recall that the polar system of a conic constitutes a 
polarity (Theorem 18, Cor., Chap. V) in which all the points and 
lines of the conic, and only these, are self-conjugate, it follows from 
the above that every polarity is given by the polar system of its 
fundamental conic. This and other results following immediately 
from it are contained in the following theorem : 

THEOREM 9. Mrery polarity w the polar &y$t&ni of a conic, the 
fwudaviental conic of the polarity. The self-conjugate points are 
the points and the self -conjugate lines are the tangents of this conic. 
Jtfvcry yoU and polar pair are pole and polar with respect to the 
fundamental conin. 

This establishes that Equation (4) represents the same conic as 
Equation (3). The last theorem may be utilized to develop the ana- 
lytic expressions for poles and polars, and tangents to a conic. This 
we take up in the next section. 



284 TWO-DIMENSIONAL PEOJECTIVITIES |CHAI>. x 

99. Poles and polars with respect to a conic. Tangents. We 

have seen that the most general equation of a conic in point coor- 
dinates may be written 

(1) a n x%+ & 22 # 2 2 + ^as&'s + 2 a u x l x 2 + 2 a 1 ^x 1 x s + 2 ^ 23 ^ 2 ^ 3 = 0. 

The result of the preceding section shows that the equation of the 
same conic in line coordinates is 

(2) 4X + A^ul + A^ul + 2 A^u, + 2 A^u, + 2 A n u t u s = 0, 
where A, is the cofactor of a,, in the determinant 



This result may also be stated as follows : 

THEOREM 10. The necessary and sufficient condition that the line 
w 1 a? 1 +^ a a; 3 + W 8 a? a =s le tangent to the conic (1) is that JSyitation (2) 
"be satisfied. 

COKOLLAKY. This condition may also le written in the form 



u, u n u 9 



= 0. 



Equation (2) of the preceding section expresses the condition that 
the points (x v # a , # 3 ) and (#/, x 9 ojj) be conjugate with respect to the 
conic (1). If in this equation (a?/, x, &) be supposed given, while 
(x v # 2 , # 3 ) is regarded as variable, this condition is satisfied by all the 
points of the polar of (x[, x^, x^) with respect to the conic and by no 
others. It is therefore the equation of this polar. When arranged 
according to the variable coordinates x it it becomes 

(3) (*!!/+ ajX+ *vf&*i + (*vPi + * M atf + ^<)^2 

, ; 



while if we arrange it according to the coordinates #?/, it becomes 
(4) (a lA + a lz x, + a lA ) x[ + (a, A + ^ 2 A + a^ < 

== 0. 



Now it is readily verified that the latter of these equations may 
be derived from the equation (1) of the conic by applying to the 
left-hand member of this equation the polar operator 



oi),ioo] VARIOUS DEFINITIONS OF CONICS 285 



( 89) and dividing the resulting equation by 2. Furthermore, if 

we define the symbols ~~ > -^ > -^ to be the result of substituting 
dx l dx^ 0a? 8 

(^/, 2 ', # 3 ') for (ajj, a? a , a? 8 ) in the expressions -^-> ~^> -^- (/ being any 

</i#.| ve/t'g </cH2g 

polynomial in x v % 2 , X Q ), it is readily seen that Equation (3) is 
equivalent to 



where now/ is the left-hand member of (1). 
This leads to the following theorem : 

THEOREM 11. Iff^Qis the equation of a conic in homogeneous 
point coordinates, the equation of the polar of any point (%{, x 9 x$ is 
given ly either of the equations 



= ( ) or Xif + x 

8x 8 *dx[ z d% o% 

If the point (%[, x^ x) is a point on the conic, either of these equa- 
tions represents the tangent to the conic f = at this point. 

100. Various definitions of conies. The definition of a (point) 
conic as the locus of the intersections of homologous lines of two 
protective flat pencils in the same plane was first given by Steiner in 
1832 and used about the same time by Ghasles. The considerations 
of the preceding sections at once suggest two other methods of defi- 
nition, one synthetic, the other analytic. The former begins by the 
synthetic definition of a polarity (cf. p. 263), and then defines a point 
conic as the set of all self-conjugate points of a polarity, and a line 
conic as the set of all self -con jugate lines of a polarity. This defini- 
tion was first given by wn Staudt in 1847. From it he derived the 
fundamental properties of conies and showed easily that his definition 
is equivalent to Steiner's. The analytic method is to define a (point) 
conic as the set of all points satisfying any equation of the second 
degree, homogeneous in three variables % v x z , x 3 . This definition (at 
least in its nonhomogeneous form) dates back to Descartes and Fermat 
(1637) and the introduction of the notions of analytic geometry. 



286 TWO-DIMENSIONAL PBOJECTIVITIBB [CHAP. X 

The oldest definition of conies is due to the ancient Greek geometers, who 
defined a conic as the plane section of a circular cone. This definition involves 
metric ideas and hence does not concern us at this point. We will return to it 
later. It is of interest to note in passing, however, that from this definition 
Apollonius (about 200 A.D.) derived a theorem equivalent to the one that the 
equation of a conic in point coordinates is of the second degree. 

The reader will find it a valuable exercise to derive for himself 
the fundamental properties of polarities synthetically, and thence to 
develop the theory of conies from von Staudt's definition, at least so 
far as to show that his definition is equivalent to Steiner's. It may 
be noted that von Staudt's definition has the advantage over Steiner's 
of including, without reference to Proposition K 2 , conies consisting 
entirely of improper points (since there exist polarities which have 
no proper self-conjugate points). The reader may in this connection 
refer to the original work of von Staudt, Die Geometric der Lage, 
Niirnberg (1847)'; or to the textbook of Enriques, Vorlesungen liber 
projective G-eometrie, Leipzig (1903). 

EXERCISES 

1. Derive the condition of Theorem 10 directly by imposing tho condition 
that the quadratic which determines the intersections of the given line with 
the conic shall have equal roots. What is the dual of this theorem ? 

2. Verify analytically the fundamental properties of poles and polars with 
respect to a conic (Theorems 13-18, Chap. V). 

3. State the dual of Theorem 11. 

4. Show how to construct the correlation between a plane of points and a 
plane 'of lines, having given the homologous pairs A, of\ B, &'; C, cf\ 7>, d'. 

5. Show that a correlation between two planes is uniquely determined if 
two pencils of points in one plane are made projective respectively with two 
pencils of lines in the other, provided that in this projectivity the point of 
intersection of the axes of the two pencils of points corresponds to the lino 
joining the two centers of the pencils of lines. 

6. Show that in our system of homogeneous point and line coSrdinaten the 
pairs of points and lines with the same coordinates are poles and polars with 
respect to the conic x * + #| + xj = 0. 

7. On a general line of a plane in which a polarity has boon defined the 
pairs of conjugate points form an involution the double points of which are 
the (proper or improper) points of intersection of the line -with the funda- 
mental conic of the polarity, 

8. A polarity in a plane is completely defined if a self-polar triangle is 
given together with one pole and polar pair of which the point is not on a 
side nor the line on a vertex of the triangle. 



f 100, 101] PAIRS OF CONICS 287 

9. Prove Theorem tt analytically. 

10. G iven a simple plane pentagon, there exists a polarity in which to each 
vertex corresponds the opposite side. 

11. The three points A', />", 6" on the sides BC, CA 9 AI)otQ, triangle that 
aro conjugate in a polarity to the vertices A> />, C respectively are collinear 
(cf. Ex. liJ, p. 125). 

12. Show that a polarity is completely determined when the two involutions 
o conjugate points on two conjugate lines are given. 

13. Construct the polarity determined by a self-polar triangle ABC and an 
involution of conjugate points on a line. 

14. Construct the polarity determined by two pole and polar pairs A, a and 
/>, // and one pair of conjugate points C, C'. 

15. If a triangle RTU is self-polar with regard to a conic C 2 , and A is any 
point of G' 2 , there are three triangles having A as a vertex which are inscribed 
to C Y2 and circumscribed to STU (ISturm, Die Lehre von den geometrischen 
Verwandtschaften, Vol. I, p. 147). 

101, Pairs of conies. If two polarities, i.e. two conies (proper or 
improper), are given, their product is a collineatiorx which leaves 
invariant any point or line which has the same polar or pole with 
regard to both conies. Moreover, any point or line which is not left 
invariant by this collineation must have different polars or poles 
with regard to the two conies. Hence the points and lines which 
have the same polars and poles with regard to two conies in the 
same plane form one of the five invariant figures of a nonidentical 
collineation. 

Type I. If the common self-polar figure of the two conies is of 
Type /, it is a self-polar triangle for both conies. Since any two conies 
are protectively equivalent (Theorem 9, Chap. VIII), the coordinate 
system may be so chosen that the equation of one of the conies, Jt a , is 

(1) itf-asJ + atf^O. 

With regard to this conic the triangle (0, 0, 1), (0, 1, 0), (1, 0, 0) is 
self-polar. The general equation of a conic with respect to which this 
triangle is self-polar is clearly 

(2) a ^ a^l + # A 2 = 0. 

An equation of the form (2) may therefore be taken as the equation 
of the other conic, # 2 , provided (1) and (2) have no other self-polar 
elements than the fxindamental triangle. Consider the set of conies 

(8) a^?- &A 2 -h o^ + XOaf- *?+) = 0. 



288 



TWO-DIMENSIONAL PRO JBCTIVITIBS [CHAP. X 



The coordinates of any point which satisfy (1) and (2) also satisfy (3), 
Hence all conies (3) pass through the points common to * and !> >a . 
For the value X = a s> (3) gives the pair of lines 
(4) (a, - a,) x* - fa - a,) v* = 0, 

which intersect in (0, 0, 1). The points of intersection of these lines 
with (1) are common to all the conies (3). 

The lines (4) are distinct, unless a^ a^ or <x 2 = <V But if ] = ,,, 
any point (xf, 0, x) on the line # 2 =0 has the polar x^ + x^ = 
both with regard to (1) and with regard to (2). The self-polar figure 
is therefore of Type IV. In order that this figure be of Type 7, the 
three numbers a v a z> a z must all be distinct. If this condition is 
satisfied, the lines (4) meet the conies (3) in four distinct points. 




FIG. 100 

The actual construction of the points is now a problem of the second 
degree. We have thus established (fig. 100) 

THEOREM 12. If two conies have a common self -polar triangle (and 
no oilier common self -polar pair of point and line), they intersect in 
four distinct points (proper or improper). Any two conies of the 
pencil determined "by these points ham the same self -polar triangle. 
Dually, two such conies have four common tangents, and any two 



wi] 



PAIRS OF CONICS 



289 



B=(010) 



eowks of tlie rttnye daterwiucd "by those common taiigents have the same 
self-j>ol(tr tri<mgk. 

CoKoiJAuy. Any pencil of comas of Type I can be represented by * 

(5) (^- afl + X (-) =0, 

4& 

the four common points being in this case (1,1, 1), (1, 1, 1), (1, 1, 1), 
1,1,1). 

//. When the 
common self-polar figure 
is of Type //, one of the 
points lies on its polar, 
and therefore, this polar is 
a tangent to eacli of the 
conies A*, jB fl . Since two 
tangents cannot intersect 
in a point of contact, the 
two lines of the self-polar 
figure are not both tan- 

gents. Hence the point JB p IGf 101 

of the self-polar figure 

which is on only one of the lines is the pole of the line I of the figure 
which is on only one of the points (fig. 101), and the line a on the two 
points is tangent to both conies at the point A which is on the two lines, 
Choose a system of coordinates with ^ = (1,0, 0), a=[0, 0, 1], 
#= (0, 1, 0), and i= [0, 1, 0]. The equation of any conic being 

2 *=* > 




the condition that A be on the conic is ^=0 ; that & then be tan- 
gent is Z> 8 = 0; that I then be the polar of J3 is ^=0. Hence the 
general equation of a conic with the given self-polar figure is 

(6) <ytf + <V* + 2 



* Equation (6) is typical for a pencil of conies of Type I, and Theorem 12 is a 
sort of converse to the developments of 47, Chap, V. The reader will note that 
if the problem of finding the points of intersection of two conies is set up directly, 
it is of the fourth degree, but that it is here reduced to a problem of the third 
degree (the determination of a common self-polar triangle) followed by two quad- 
ratic constructions. This corresponds to the well-known solution of the general 
biquadratic equation (cf. Fine, College Algebra, p. 486). For a further discussion 
of the analytic geometry of pencils of conies, cf . Clebsoh-Lindemann, Vorlesungen 
iiber Geometrie, 2d ed., Vol. I, Part I (1906), pp. 212 ff. 



290 TWO-DIMENSIONAL PROJBOTIVITIES LCHAP. x 

Since any two conies are projectively equivalent, A 1 may be chosen 
to be 

(7) +;+ 2 3^=0. 

The equation of B* then has the form (6), with the condition that 
the two conies have no other common self-polar elements. Since the 
figure in which a is polar to A and b to B can only reduce to Types 
IV or F, we must determine under what conditions each point on a 
or each point on I has the same polar with regard to (6) and (7). 
The polar of (x[, x, A- 8 ') with regard to (6) is given by 

1^3= 0- 



Hence the first case can arise only if & 2 = & 2 ; and the second only 
if a = 6 . 

> s 

Introducing the condition that a z , 8 , 5 2 are all distinct, it is then 
clear that the set of conies 



af* + a,x* + 2 l&Xi + X (xj + ^ + 2 x A ) = 
contains a line pair for X = # 2 , viz. the lines 



Hence the conies have in common the points of intersection with (7) 

of the line 

(o,-a a ) 0,+ 2(5,- ^=0. 
This gives 

THEOREM 13. If two conies have a common self -polar figure of 
Type II, tliey have three points in common and a common tangent at 
one of them. Diially, tliey have three common tangents and a common 
point of contact on one of the tangents. The two conies determine a, 
pencil and also a range of conies of Type II. 

COROLLARY. Any pencil of conies of Type II may le represented 
ly the equation x% # 8 2 + \x^ = 0. The conies of this pencil all pass 
through the points (0, 1, 1), (0, 1, 1), (1, 0, 0) and are tangent to 
# 3 = 0. 

Type III. When the common self-polar figure is of Type ///, the 
two conies evidently have a common tangent and a common point 
of contact, and only one of each. Let the common tangent be # 8 =s 0, 
its point of contact be (1, 0, 0), and let A 2 be given by 

(8) ^+ 



PAIRS OF CWICS 291 

The general equation of a conic tangent to # 3 = at (1, 0, 0) is 
( 8) <Vtt? + <v + 2 MA + 2 MA = 0, 

with regard to which the polar of any point (#/, &/, 0) on # 3 = is 
given l)y 
( 1 0) 



This will be identical with the polar of (a x^ 0) with regard to A* 
for all values of as/, ai if &,,= 2 and \*= 0. Since (1, 0, 0) only is to 
have the same polar with regard to both conies, we impose at least 
one of the conditions & 2 =tf 2 , ^=0. The line (10) will now be 
identical with the polar of (8) for any point (x[, #/, 0) satisfying the 
condition 



This quadratic equation must have only one root if the self-polar figure 
is to be of Type ///. This requires & 2 = a a , and as J 2 , a 2 cannot both 
l)e unless (9) degenerates, the equation of B* can be taken as 

(11) xl + 2 x^ + off + 2 MA = 0, (B 1 *= 0). 




A~(lQO) a=[ooi] 



The conies (8) and (11) now evidently have in common the points of 
intci'HGction of (8) with the 
line pair 

8 a; 8 fl + 2^8^8=0, 

and no other points. Since 

$*% = is a tangent, this gives 

two common points. If the 

second common point is taken 

to be (0, 0, 1), the set of 

conies which have in com- 

mon the points (0, 0, 1) and (1, 0, 0) = A and the tangent a at A 9 

and no other points, may be written (fig. 102) 

(1 2) <* + 2 fa + \% 2 x s = 0. 

THEOREM 14. If two conies have a common self-polar figure of 
Type III t thfiy have two points in common and a common tangent 
at one of them, and one other common tangent. They determine a 
pencil and a range of conies of Type III. 



292 TWO-DIMENSIONAL PROJECTIVITIES [OHAP.X 

COROLLARY. A pencil of conies of Type III can "be represented ly 
the equation x%+ 2 #3^ + \a? 2 # 3 = 0. 

Type IV. When the common self-polar figure is of Type 7F, let the 
line of fixed points be X B = and its pole be (0, 0, 1). The coordinates 
being chosen as they were for Type /, the conic A* has the equation 

X^ - XQ + #?g = U J 

and any other conic having in common with A* the self-polar tri- 
angle (1, 0, 0), (0, 1, 0), (0, 0, 1) has an equation of the form 

2== 0. 



The condition that every point on # 8 = shall have the same polar 
with regard to this conic as with regard to A* is a l = a y Hence B 
may be written 




Any conic of this form has the same tangents as A 2 at the points 
(1, 1, 0) and (1, 1, 0) (fig. 103). Hence, if X is a variable parameter, 

the last equation represents 
a pencil of conies of Type IV 
according to the classification 
previously made. 

THEOREM 1 5. If two conies 
l la w a common self-polar 
figure of Type IV, they have 

Fio 103 two points in common and 

the tangents at these points, 

They determine a pencil (which is also a range) of conies of Type IV. 
COKOLLARY. A pencil of conies of Type IV may le represented ly 

the equation 

x* 

and also by the equation 

Type V. When the common self-polar figure is of Type F, let the 
point of fixed lines be (1, 0, 0) and the line of fixed points be # 8 = 0. 
As in Type ///, let J? be given by 

(8) ^+2^3=0. 

We have seen, in the discussion of that type, that all points of x^ 
have the same polars with respect to (8) and (9), if in (9) we have 



101] PAIES OF CONICS 293 

Z> a ss a% and l^ 0. Hence, if A* and J? 2 are to have a common self- 

polar figure of Type V, the equation of & must have the form 

(13) a.(x*+2x A )+a^*~Q. 

Prom the form of equations (8) and 

(13) it is evident that the conies have 

in common only the point (1, 0, 0) and 

the tangent X B = 0, and that every point 

on # 8 = has the same polar with ro- o,= [o0lj A = (\wo) 

spect to both conies (fig. 104). Hence 

they determine a pencil of Type V. 

THEOREM 16. If two conies have a common self -polar figure of 
Type V, they have a lineal element (and no other elements) in com- 
mon and determine a pencil (which is also a range) of conies of 
Type V according to the classification already given. 

COROLLAHY. A pencil of conies of Type V can le represented ly the 

equation 

1 * 




As an immediate consequence of the corollaries of Theorems 12-16 
we have 

THEOREM 17. Any pencil of conies may le written in the form 



where f~ and #= are the equations of two conies (degenerate or 
not) of the pencil. 

EXERCISES 

1. Prove analytically that the polars of a point P with, respect to the 
conies of a pencil all pass through a point 'Q. The points P and Q are double 
points of the involution determined by the conies of the pencil on the line PQ. 
Give a linear construction for Q (cf. Ex. 3, p. 13 C). The correspondence 
obtained by letting every point P correspond to the associated point Q is a 
< < quadratic birational transformation. ' ' Determine the equations representing 
this transformation. The point Q, which is conjugate to P with regard to all 
the conies of the pencil, is called the conjugate of P with respect to the pencil. 
The locus of the conjugates of the points of a line with regard to a pencil of 
conies is a conic (cf. Ex. 81, p. 140). 

2. One an<J only one conic passes through four given points and has two 
given points as conjugate points, provided the two given points are not con- 
jugate with respect to all the conies of the pencil determined by the given 
set of four. Show how to construct this conic. 



294 TWO-DIMENSIONAL PEOJECTIVITIES [CHAP.X 

3. One conic in general, or a pencil of conies in a special case, passes 
through three given points and has two given pairs of points as conjugate 
points. Give the construction. 

4. One conic in general, or a pencil of conies in a special case, passes 
through two given points and has three pairs of given points as conjugate 
points ; or passes through a given point and has four pairs of given points as 
conjugate points ; or has five given pairs of conjugate points. Give the cor- 
responding constructions for each case. 

102. Problems of the third and fourth degrees.* The problem of 
constructing the points of intersection of two cdhics in the same 
plane is, in general, of the fourth degree according to the classifi- 
cation of geometric problems described in 83. Indeed, if one of 
the coordinates be eliminated between the equations of two conies, 
the resulting equation is, in general, an irreducible equation of the 
fourth degree. Moreover, a little consideration will show that any 
equation of the fourth degree may be obtained in this way. It 
results that every problem of the fourth degree in a plane may 
be reduced to the problem of constructing the common points (or 
by duality the common tangents) of two conies. Further, the prob- 
lem of finding the remaining intersections of two conies in a plane 
of which one point of intersection is given 1 , is readily seen to be of 
the third degree, in general; and any problem of this degree can be 
reduced to that of finding the remaining intersections of two conies 
of which one point of intersection is known. It follows that any 
problem of the third or fourth degree in a plane may be reduced 
to that of finding the common elements of two conies in the 
plane, f 

A problem of the fourth (or third) degree cannot therefore be 
solved by the methods sufficient for the solution of problems of the 
first and second degrees (straight edge and compass). | In the case 
of problems of the second degree we have seen that any such prob- 
lem could be solved by linear constructions if the intersections of 

* In this section wo have made use of Amodeo, Lezioni di Ooomctria Projettiva, 
pp. 40(5, 437. Some of the exercises are taken from the same book, pp. 448-451. 

1 Moreover, we have soon (p, 28!), footnote) that any problem of the fourth 
degree may be reduced to one of the third degree, followed by two of the floc.ond 
degree. 

| "With the usual representation of the ordinary real geometry \vo should require 
an instrument to draw conies. 



H)2] TI1IJID AND FOURTH DEGKEE PROBLEMS 295 

every line in the plane with a fixed conic in that plane were assumed 
known. Similarly, any problem of the fourth (or third) degree can 
he solved by linear and quadratic constructions if the intersections 
of every conic in the plane with a fixed conic in this plane are 
assumed known. This follows readily from the fact that any conic 
in the plane can "be transformed by linear constructions into the 
fixed conic. A problem of the third or fourth degree in a plane 
will then, in the future, be considered solved if it has been reduced 
to the finding of the intersections of two conies, combined with 
any linear or quadratic constructions. As a typical problem of the 
third degree, for example, we give the following: 

To find the double points of a nonpcrspective collineation in a plane 
'which is* determined ly four pairs of homologous points. 

Solution. When four pairs of homologous elements are given, we 
can construct linearly the point or line homologous with any given 
point or line in the plane. Let the collineation be represented by II, 
and lot A be any point of the plane which is not on an invariant 
line. Let U(A)=A r and II (A 1 )** A". The points A, A 1 , A" are then 
not collinear. The pencil of lines at 
A is protective with the pencil at 
A 1 , and these two projective pencils 
generate a conic C 2 which passes 
through all the double points of II, 
and which is tangent at A' to the 
line A r A" (fig. 105). The conic <7 2 is 
transformed by the collineation II 
into a conic C? generated by the pro- 
jective pencils of lines at A f and A". FlQ " 10 
6\ a also passes through A! and is tangent at this point- to the line 
A A 1 . The double points of II are also points of C*. The point A f 
is not a double point of II by hypothesis. It is evident, however, 
that every other point common to the two conies C 2 and C* is a 
double point. 

If (7 s and C? intersect again in three distinct points Z, H> N 9 the 
latter form a triangle and the collineation is of Type /. If <7 2 and C 
intersect in a point N, distinct from A f , and are tangent to each other 
at a third point L=zM, the collineation has M, N for double points 




296 TWO-DIMENSIOKAL PEOJECTIVITIES [CHAP.X 

and the line MN and the common tangent at M for double lines 
(fig, 106); it is then of Type //. If, finally, the two couics have 
contact of the second order at a point L = M=* N, distinct from A r , 
the collineation has the single double line which is tangent to the 
conies at this point, and is of Type III (fig. 107). 





, 106 tfw. 107 

EXERCISES 

1. Give a discussion of the problem above without making at the outset 
the hypothesis that the collineation is nonperspective. 

2. Construct the double pairs of a correlation in the piano, which is not 
a polarity. 

3. Given two polarities in a plane, construct their common polo and 
polar pairs. 

4. On a line tangent to a conic at a point A is given an involution I, and 
from any pair of conjugates P, P f of I are drawn the second tangents p, p' to 
the conic, their points of contact being Q, Qf respectively. Show that the IOCUH 
of the point pp' is a line, /, passing through the conjugate, A', of A in the invo- 
lution I ; and that the line QQ' passes through the pole of I with respect to 
the conic. 

5. Construct the conic which is tangent at two points to a given conic and 
which passes through three given points. Dualize. 

6. The lines joining pairs of homologous points of a noninvolutoric pro- 
jectivity on a conic A* are tangent to a second conic B* which is tangent to 
A 2 at two points, or which hyperosculates A*. 

7. A pencil of conies of Type // is determined by three points A, B, C 
and a line c through C. What is the locus of the points of contact of tho 
conies of the pencil with the tangents drawn from a givn point P of <?? 

8. Construct the conies which pass through a given point P and which are 
tangent at two points to each of two given conies. 

9. If /= 0, g s= 0, h = are the equations of thrcd couicH in a piano not 
belonging to the same pencil, the system of conies given by the equation 



JOiiJ THIRD AND FOURTH DEGREE PROBLEMS 297 

X, ju,, v being variable parameters, is called a bundle of conies. Through every 
point of the plane passes a pencil of conies belonging to this bundle ; through 
any two distinct points passes in general one and only one conic of the bundle. 
If the conies /, r/, h have a point in common, this point is common to all the 
conies of the bundle. Give a nonalgebraic definition o a bundle of conies. 

10. Tho set of all conies in u plane passing through the vertices of a triangle 
form a bundle. Tf the equations of the sides of this triangle are I = 0, m = 0, 
n = 0, show that the bundle may be represented by the equation 

Xwn + jjinl + vim = 0. 



What aro the degenerate conies oi this bundle ? * 

11. The not of all conies in a plane which have a given triangle as a self- 
polar triangle forms a bundle. If the equations of the sides of this triangle are 
/ = 0, m 0, n = 0, show that the bundle may be represented by the equation 

A/ 2 4- /im a + w* = 0. 

What aro the degenerate conies of this bundle? 

12. The conies of the bundle described in Ex. 11 which pass through a 
general point P of the plane pass through the other three vertices of the 
quadrangle, of which one vertex is P and of which the given triangle is the 
diagonal triangle. What happens when P is on a side of the given triangle ? 
Dualize. 

13. The reflections whose centers and axes are the vertices and opposite 
aides of a triangle form a commutative group. Any point of the plane not 
on a side of the triangle is transformed by the operations of this group into 
the other three vertices of a complete quadrangle of which the given triangle 
is the diagonal triangle. If this triangle is taken as the reference triangle, 
what are the equations of transformation ? What conies are transformed into 
themselves by the group, and how is it associated with the quadrangle- 
quadrilateral configuration ? 

14. The necessary and sufficient condition that two reflections be com- 
mutative Is that the center of each shall be on the axis of the other. 

15. The invariant figure of a collineation may be regarded as composed of 
two lineal elements, the five types corresponding to various special relations 
between the two lineal elements. 

16. A correlation which transforms a lineal element Aa into a lineal 
clement 7#; and also transforms Bl into -la is a polarity. 

17. How many collineations and correlations are in the group generated 
by the. reflections whoso centers and axes are the vertices and opposite sides 
of a triangle and a polarity with regard to which the triangle is self -polar ? 

* In connection with this and the two following exercises, cf . Castelnuovo, 
Lessioni di Geornetria Analitica e Projettiva, Vol. I, p. 395. 



CHAPTER XI* 



FAMILIES OF LINES 

103- The regulus. The following theorem, on which depends the 
existence of the figures to be studied in this chapter, IKS logically 

equivalent (in the presence of Assump- 
tions A and E) to Assumption P. 1 1 
might have been used to replace that 
assumption. 

THEOREM 1. If l v 1 S9 1 9 are three, 
mutally skew lines, and if m v w 2 , w 8 , 
w 4 are four lines each of which meets 
each of the lines l v /, 7 3 , then any line l^ 
which meets three of the lines m v m^ 
m s , m^ also meets the fourth. 

Proof. The four planes l$n v Ijn^ 
Z x m 3 , lp\ of the pencil with axis l l are 
perspective through the pencil of points 
on 1 3 with the four planes l z m v / a w fi , 
fy j h> ^2 m 4 ^ ^ P enc il with axis / a 
(fig. 108). For, by hypothesis, the lines 
of intersection m v w 2 , m^ m^ of the 
pairs of homologous planes all meet / 8 . 
The set of four points in which the four planes of the pencil on ^ 
meet Z 4 is therefore projective with the set of four points in which 
the four planes of the pencil on 2 meet Z 4 . But l meets three of the 
pairs of homologous planes in points of their lines of intersection, 
since, by hypothesis, it meets three of the lines m v m 2 , w 3 , w 4 . Hence 
in the projectivity on l there are three invariant points, and hence 
(Assumption P) every point is invariant. Hence l meets the remain- 
ing line of the set m v m^ m v m. 

* AH the developments of this chapter are on the basis of Assumptions A, E, P, HO. 
But see the exercise oa page 201. 

208 




108 



SW] THE EEG-ULUS 299 

DEFINITION. If l v Z 2 , l s are three lines no two of which are in the 
Banie plane, the set of all lines which meet each of the three given lines 
is called a regulus. The lines l v ,, 1% are called directrices of this regulus. 

It is clear that no two lines of a regains can intersect, for other- 
wise two of the directrices would lie in a plane. The next theorem 
follows at once from the last. 

THEOREM 2. If l v / 2 , Z 8 are three lines of a regulus of which 
m v v/6 a , w 8 are directrices, m v m 2 , m 3 are lines of the reyidus of which 
l v l, / 8 arc directrices. 

It follows that any three lines no two of which lie in a plane are 
directrices of one and only one regulus and are lines of one and only 
one regulus. 

DEFINITION. Two reguli which are such that every line of one 
meets all the lines of the other are said to be conjugate. The lines of 
a regulus are called its generators or rulers ; the lines of a conjugate 
regulus are called the directrices of the given regulus. 

THEOREM 3, Every regulus has one and only one conjugate regulus. 

This follows immediately from the preceding. Also from the proof 
of Theorem 1 we have 

THEOREM 4. Tlie correspondence THEOREM 4'. The correspond- 
established ly the lines of a regu- ence established "by the lines of a 
lux between the points of two lines regulus between the planes on any 
of its conjugate regulus is projec- two lines of its conjugate regulus 
tivc. is projectile. 

THEOR*^ 5. The set of all lines THEOREM 5'. The set of all lines of 
joining pairs of homologous points intersection of pairs of homologous 
of two projectioe pencils of points planes of two projectile pencils of 
on #7iYW lines is a regulus. planes on skew lines is a regulus. 

Proof. We may confine ourselves to the proof of the theorem on 
the left. By Theorem 6, Chap. Ill, the two pencils of points are 
perspective through a pencil of planes. Every line joining a pair of 
homologous points of these two pencils, therefore, meets the axis of 
the pencil of planes. Hence all these lines meet three (necessarily 
skew) lines, namely the axes of the two pencils of points and of the 
pencil of planes, and therefore satisfy the definition of a regulus. 
Moreover, every line which meets these three lines joins a pair of 
homologous points of the two pencils of points. 



300 FAMILIES OF LINES LOUAP. XI 

THEOREM 6. If [p] are tlie lines of a rcyidus tt,nd q in u* dirwkrir, 
of tlie regulus, the pencil of points gfj?] 'is projcvtire with, tlie, pMwil 
of planes q [p]. 

Proof. Let g[ be any other directrix. By Theorem 4 the pencil of 
points q[p] is perspective with the pencil of points j'jjp]. But each 
of the points of this pencil lies on the corresponding piano <jj>. 
Hence the pencil of points < f [p] is also perspective with the pencil 
of planes q [p]. 

EXERCISES 

1. Every point which is on a line of a rogulus is also on a line of its 
conjugate regains. 

2. A plane which contains one line of a regnlus contains also a line of its 
conjugate regains. 

3. Show that a regains is uniquely denned by two of its lines and throe 
of its points,* provided no two of the latter are coplanar with either of the 
given lines. 

4. If four lines of a regains cut any line of the conjugate regains in points 
of a harmonic set, they are cut by every such line in points of a harmonic 
set. Hence give a construction for the harmonic conjugate of a line oi! a 
regains with respect to two other lines of the regains. 

5. Two distinct reguli can have in common at most two distinct linen. 

6. Show how to construct a regnlus having in common with a given 
regains one and but one ruler. 

104. The polar system of a regulus. A plane meets every line of 
a regulus in a point, unless it contains a line of the regulus, in which 
case it meets all the other lines in points that are collinear. Since 
the regulus may be thought of as the lines of intersection of pairs of 
homologous planes of two projective axial pencils (Theorem fi'), the 
section by a plane consists of the points of intersection of pairs of 
homologous lines of two projective flat pencils. Hence the section 
of a regulus by a plane is a point conic, and the conjugate regulun 
has the same section. By duality the projection of a regulus and its 
conjugate from any point is a cone of planes. 

The last remark implies that a line conic is the "picture " in a piano of 
a regulus and its conjugate. For such a picture in clearly a plant) Hoctiou of 
the projection of the object depicted from the eye of an observer. Fig. 108 
illustrates this fact. 

* By a point of a regains is meant any point on a line of the regains. 



iw] THE REGULUS 301 

The section of a regiilus by a plane containing a line of the regu- 
lus is a degenerate conic of two lines. The plane section can never 
degenerate into two coincident lines because the lines of a regulus 
and its conjugate are distinct from each other. In like manner, the 
projection from a point on a line of the regulus is a degenerate cone 
of planes consisting of two pencils of planes whose axes are a ruler 
and a directrix of the regulus. 

DEFINITION. The class of all points on the lines of a regulus is 
called a surface of the second order or a guadric surface. The planes 
on the lines of the regulus are called the tangent planes of the sur- 
face or of the regulus. The point of intersection of the two lines of 
the regulus and its conjugate in a tangent plane is called the point 
of contact of the plane. The lines through the point of contact in a 
tangent plane are called tangent lines, and the point of contact of the 
plane is also the point of contact of any tangent line. 

The tangent lines at a point of a quadric surface include the lines 
of the two conjugate reguli through this point and all other lines 
through this point which meet the surface in no other point Any 
other line, of course, meets the surface in two or no points, since a 
plane through the line meets the surface in a conic. The tangent 
lines are, by duality, also the lines through each of which passes only 
cue tangent plane to the surface. 

THEOREM 7. TJw tangent planes at the points of a plane section of 
a quadric surface pass through a point and constitute a cone of planes. 
Dually, the points of contact of the cone of tangent planes through a 
point are coplanar and form a point conic. 

Proof. It will suffice to prove the latter of these two dual theorems. 
Let the vertex P of the cone of tangent planes be not a point of the 
surface. Consider three tangent planes through P, and their points of 
contact. The three lines from these points of contact to P are tan- 
gent lines o the surface and hence there is only one tangent plane 
through each of them. Hence they are lines of the cone of lines asso- 
ciated with the coue of tangent planes. Let TT be the plane through 
their points of contact. The section by TT of the cone of planes through 
P is therefore the conic determined by the three points of contact 
and the two tangent lines in which two of the tangent planes meet 
TT. The plane TT, however, meets the regulus in a conic of which the 
three points of contact are points. The two lines of intersection with 



302 FAMILIES OF LINES LULU-. XI 

TT of two of the tangent planes througli P are tangents to this conic, 
because they cannot meet it more than one point each. The section 
of the surface and the section of the cone ot planes then have three 
points and the tangents through two of them in common. Hence these 
sections are identical, which proves the theorem when P IN not on 
the surface. 

If P is on the surface, the cone of planes degenerates into two lines 
of the regulus (or the pencils of planes on these lines), and the points 
of contact of these planes are all on the same two lines. Hence the 
theorem is true also in this case. 

DEFINITION. If a point P and a plane TT are so related to a regulus 
that all the tangent planes to the regulus at points of its section 
by TT pass through P (and hence all the points of contact of tangent 
planes through P are on TT), then P is called the pole of vr and TT the 
polar of P with respect to the regulus. 

COROLLARY. A tangent plane to a regulus is the polar of its point 
of contact. 

THEOREM 8. The polar of a point P not on a regidus contains all 
points P 1 such that the line PP 1 meets the siwface in two points which 
are harmonic conjugates with respect to P, PJ 

Proof. Consider a plane, a, through PP' and containing two lines 
a, Z> of the cone of tangent lines through P. This plane meets the 
surface in a conic <7 2 , to which the lines a f I are tangent. As the polar 
plane of P contains the points of contact of a and ft, its section by a 
is the polar of P with respect to <7 2 . Hence the theorem follows 
as a consequence of Theorem 13, Chap. V, 

THEOREM 9. The polar of a point of a plane TT with respect to a 
regulus meets TT in the 'polcvr line of this point with regard to the conic 
which is the section of the regulus ly TT. 

Proof. By Theorem 8 the line in which the polar plane meets TT 
has the characteristic property of the polar line with respect to a conic 
(Theorem 13, Chap. V). This argument applies equally well if the 
conic is degenerate. In this case the theorem reduces to the following 

COROLLARY. The tangent lines of a regulus at a point on it, are 
paired in an involution the double lines of which are the ruhr (tn,d 
directrix through that point. Each line of a pair contains the polar 
points of all the planes on the other line. 



w*J THE EEGULUS 303 

THKOKBM 10. The polar s with regard to a regulus of the points of 
a line I are an axial pencil of planes projeetive with the pencil of 
points on L 

Proof. In case the given line is a line of the regulus this reduces 
to Theorem 6. In any other case consider two planes through I. In 
each plane the polars of the points of I determine a pencil of lines 
projeetive with the range on L Hence the polars must all meet the 
line joining the centers of these two pencils of lines, and, being per- 
spective with either of these pencils of lines, are projeetive with the 
range on /. 

DMFINITTON. A line V is polar to a line I if the polar planes of the 
points of I meet on V. A line is conjugate to I if it meets V. A point 
JP r is conjugate to a point P if it is on the polar of P. A line p is 
conjugate to P if it is on the polar of P. A plane TT' is conjugate to 
a plane TT if TT' is on the pole of ir. A line p is conjugate to TT if it 
is on the pole of TT. 

EXERCISES 

Polar points and planes with respect to a regulm are denoted "by corresponding 
capital Roman and small Greek letters. Conjugate elements of the same kind are 
denoted by the same letters wilh primes. 

1. If TT is on It, then 7* is on g. 

2. If Ms polar to /, then /is polar to /. 

3. If one clement (point, line, or plane) is conjugate to a second element, 
then the second element is conjugate to the first. 

4. If two lines intersect, their two polar lines intersect. 

5. A ruler or a directrix of a regulus is polar to itself. A tangent line is 
polar to its harmonic conjugate with regard to the ruler and directrix through 
its point of contact. Any other line is skew to its polar. 

6. The points of two polar lines are conjugate. 

7. The pairs of conjugate points (or planes) on any line form an involu- 
tion the double points (planes) of which (if existent) are on the regulus. 

8. The conjugate lines in a flat pencil of which neither the center nor the 
plane is on the regulus form an involution. 

9. The line of intersection of two tangent planes is polar to the line 
joining the two points of contact. 

10. A line of! the regulus which meets one of two polar lines meets the other. 

11. Two one- or two-dimensional forms whose bases are not conjugate or 
polar are projeetive if conjugate elements correspond. 

12. A line / is conjugate to I' if and only if some plane on / is polar to 
some point on /'. 



304 FAMILIES OF LINES LO.IAP. xi 

13. Show that there are two (proper or improper) lines ?, ,s nutting two 
given linos and conjugate to them "both. Show also that r is the polar of . 

14. Jf , b, c are three generators of a regains and ', //, <' three of the con- 
jugate rcgulus, then the three diagonal lines joining the points 

(?;<?') and (Zi'e), 
(c'a) and (c'), 
(&') and (a'fi) 

meet in a point tf which is the pole of a plane containing the lines of intersec- 
tion of the pairs of tangent planes at the same vertices. 

15. The six lines a, &, c, a', ft', c' of Ex. 14 determine the following trios 
of simple hexagons 

(bc'aVca') , (Aa'aeW) , (W/aa'cc") , 
(&c'aa'cft') , (bb'ac'ca') , (6a W/cc') . 

The points S determined by each trio of hexagons are collinear, and the two 
lines on which they lie are polar with regard to the quadric surface.* 

16. The section of the figure of Ex. 14 by a plane leads to the Pascal 
and Brianchon theorems ; and, in like manner, Ex. 15 leads to the theorem 
that the 60 Pascal lines corresponding to the GO simple hexagons formed 
from 6 points of a conic meet by threes in 20 points which constitute 10 
pairs of points conjugate with regard to the conic (cf. Ex. 10, p. 



105. Protective conies. Consider two sections of a regains by 
planes which are not tangent to it. These two conies are both per- 
spective with any axial pencil of a pair of axial pencils which generate 
the regulus (cf. 76, Chap. VIII). The correspondence established 
between the conies by letting correspond pairs of points which lie on 
the same ruler is therefore projective. On the line of intersection, I, 
of the two planes, if it is not a tangent line, the two conies determine 
the same involution I of conjugate points. Hence, if one of them inter- 
sects this line in two points, they have these two points in common. 
If one is tangent, they have one common point and one common 
tangent. The projectivity between the two conies fully determines a 
projectivity between their planes in which the HBO I is transformed 
into itself. The involution I belongs to the projectivity thus deter- 
mined on I The converse of these statements leads to a theorem 
which -is exemplified in the familiar string models: 

THEOREM 11. The lines joining corresponding points of 9 two pro- 
jective conies in different planes form a reyul'U$ t provided GM two 
conies determine the same involution, T, of conjugate points on th.Q 

* Cf. Saimia, Lezioni di Geometria Projettiva (Naples, 1805), pp. 



10SJ 



PKOJECTIVE CONICS 



305 



lin,e of intersection, I, of the two planes; and provided the collineation 
between the two pkmes determined ly the correspondence of the conies 
transforms I into itself ly a projectiviky to which I belongs (in par- 
ticular, 'if the conies meet in two points which are self -corresponding 
in> the. 



Proof. Let L be the pole with regard to one conic of the line of 
intersection, /, of the two planes (fig. 109). Let A and B be two 




points of this conic collinear with L and not on L The conic is gen- 
erated by the two pencils A [P] and B [P r ] where P and P f are con- 
jugates in the involution I on I (cf. Ex. 1, p. 137). Let A and 
B be the points homologous to A and B on the second conic, and let 
2 be the point in which the second conic is_ met by the plane con- 
taining A, A> and the tangent at A\ and let B be the point in which 
the second conic is met by the plane of B, B, and the tangent at B. 

The line AB contains the pole of I with regard to the second conic 
because this line is projective with --4-??. Since the tangents to the 
first conic at A and B meet on Z, the complete quadrangle AABB has 
one diagonal point, the intersection of A A and BB, on I ; hence the 



306 FAMILIES OF LINES l<!iui.xi 

opposite side of the diagonal triangle pauses through the polo of /, 
Hence it intersects AB in the pole of /. But the intersection of All 
with An is on this diagonal line. Hence All uieetH AB in the pole 
of /. Hence the pencils A[P] and B[P f "\ generate the second conic. 
Hence, denoting by a and 6 the lines A A and BB, the pencils of planes 
a [P] and I [P f ] are protective and generate a regulus of which the 
two conies are sections. 

The projectivity between the planes of the two conies established 
by this regulus transforms the line I into itself by a projectivity to 
which the involution I belongs and makes the point A correspond 
to A. The projectivity between two conies is fully determined by 
these conditions (cf. Theorem 12, Cor. 1, Chap. VIII). Hence the 
lines of the regulus constructed above join homologous points in the 
given projectivity. Q.E.D. 

It should be observed that if the two conies are tangent to I, the 
projectivity on I fully determines the projectivity between the two 
conies. For if a point P of I corresponds to a point Q of /, the unique 
tangent other than I through P to the first conic must correspond to 
the tangent to the second conic from Q. If the projectivity between 
the two conies is to generate a regulus, the projectivity on I must be 
parabolic with the double point at the point of contact of the conies 
with L For if another point D is a double point of the projectivity 
on I, the plane of the tangents other than I, through D to the two 
conies meets each conic in one and only one point, and, as these 
points are homologous, contains a straight line of the locus generated. 
As this plane contains only one point on either conic, it meets the 
locus in only one line, whereas a plane meeting a regulus in one 
line meets it also in another distinct line. 

Since the parabolic projectivity on I is fully determined by the 
double point and one pair of homologous points, the projectivity be- 
tween the two conies is fully determined by the correspondent of one 
point, not on Z, of the first conic. 

To show that a projectivity between the two conies which is para- 
bolic on I does generate a regulus, let A be any point of the first 
conic and A J its correspondent on the second (fig. 110). Let the 
plane of A 1 and the tangent at A meet the second conic in A n . 
Denote the common point of the two conies by , and consider the 



> 105 J 



PEOJECTIVE CONICS 



307 




FIG. 110 



two conies as generated by the flat pencils at A and B and at A" 
and 11. The correspondence established between the two flat pencils 
at 7> by letting correspond lines joining B to homologous points of 
the two conies is perspective because the line I corresponds to itself. 
Hence there is a pencil of 
planes whose axis, Z>, passes 
through 7* and whose planes 
contain homologous pairs 
of lines of the flat pencils 
at 7A The correspondence 
established in like manner 
between the flat pencil at A 
and the flat pencil at ./I" may 
bo regarded as the product 
of the projectivity between 
the two planes, which car- 
ries the pencil at A to the 
pencil at A', followed by 
the projectivity between the 
pencils at A' and A n generated by the second conic. Both of these 
projectivities determine parabolic projectivities on I with B as inva- 
riant point. Hence their product determines on I either a parabolic 
projectivity with B as invariant point or the identity. This product 
transforms the tangent at A into the line A n A*. As these lines meet 
/ in the same point, the projectivity determined on I is the identity. 
Hence corresponding lines of the projective pencils at A and A n meet 
on Z, and hence they determine a pencil of planes whose axis is a=AA lf . 
The axial pencils on a and & are projective and hence generate a 
regulus the lines of which, by construction, pass through homologous 
points of the two conies. We are therefore able to supplement 
Theorem 11 by the following 

COROLLARY 1. The lines joining corresponding points of two pro- 
jectwe conies in different planes form a regulus, if the two conies 
have a common tangent and point of contact and the projectivity 
determined between the two planes ly the projectivity of the conies 
transforms their common tangent into itself and has the common 
point of the two conies as its only fixed point. 



808 FAMILIES OB 1 LINES [CHAP. XI 

Tlu. generation of a ivguiuH by projective ranges of points on skew 
linen may bo regarded UH a degenerate case of this theorem and eor- 
olluty, A further degenerate cane is stated in the first exercise, 

The proof of Theorem 11 given above is more complicated than it would 
lutve been if, under Proposition K 2 , we had made use of the points of intor- 
Hwtioii of the lino I with the two conies. But since the discussion of linear 
families of linen in the following section employs only proper elements and 
depends in part on this theorem, it seems more satisfactory to prove this 
theorem as we have done. It is of course evident that any theorem relating 
entirely to proper elements of space which is proved with the aid of Proposi- 
tion, K n can also be proved Ly an argument employing only proper elements. 
The latter form of proof is often much more difficult than the former, but it 
often yields more information as to the constructions related to the theorem. 

These results may be applied to the problem of passing a quadric 
surface through a given set of points in space. Proposition K 2 will be 
used in this discussion so as to allow the possibility that the two con- 
jugate reguli may be improper though intersecting in proper points. 

COROLLARY 2. If three planes a, fi, j meet in three lines a = /3% 
6 = yor, e = a/3 and contain three conies A 2 , *, C 2 , of which Jf and C' 2 
meet in two points P, P r of a, (? and A 2 meet in two points Q, Q r of I, 
and A 2 and B* meet in two points E, It' of e, then there is one and lut 
one qiiadric surface * containing the points of the three conic,*. 

Proof. let M be any point of <7 2 . The conic # 2 is projected from 
M by a cone which meets the plane a in a conic which intersects A* 
in two points, proper or improper or coincident, other than H and Ji 1 . 
Hence there are two lines m, m f , proper or improper or coincident, 
through M which meet both A* and B*. The projectivity determined 
between A* and JB* by either of these lines generates a regulus, or, 
in a special case, a cone of lines, the lines of which must pass through 
all points of <7 2 because they pass through P, P', Q r , Q, and M t all of 
which are points of <7 2 . 

The conjugate of such a regulus also contains a line through M 
which meets both A* and B*. Hence the lines m and m' determine 
conjugate reguli if they are distinct. If coincident they evidently de- 
termine a cone. The three conies being proper, the quadrio miittt con- 
tain proper points even though the lines rn, m f are improper. 

* In this corollary and in Theorem 12 the term quadric surface must be taken 
to include the points on a cone of lines as a special case. 



1031 



QUADEJU THROUGH NINE POINTS 



309 



9 

If six points 1, 2, , 4, 5, 6 are given, no four of which are col- 
linear,* there evidently exist two planes, cc and /?, each containing 
three of the points and having none on their line of intersection. 




FIG. ill 



Assign the notation so that 1, 2, 3 are in #. A quadric surface which 
contains the six points must meet the two planes in two conies A 2 , 
If which meet the line #/3=c in a common point-pair or point of 
contact ; and every point-pair, proper or improper or coincident, of c 
determines such a pair of conies. 

Let us consider the problem .of determining the polar plane & of 
an arbitrary point on the line c. The polar lines of with regard 
to a pair of conies JL and B* meet c in the same point and hence 
determine o>. If no two of the points 1, 2, 3, 4, 5, 6 are collinear 
with 0, any line I in the plane a determines a unique conic A 2 with 
regard to which it is polar to 0, and which passes through 1, 2, 3. 
A* determines a unique conic 2? a which passes through 4, 5, 6 and 
meets c in the same points as -4 2 ; and with regard to this conic 

* The construction of a quadric surface through nine points by the method used 
in the text is given in Rohn'and Papperitz, Darstellende Geometric, Vol. II 
(Leipzig, 1806), 676, 677. 



310 FAMILIES OF LINES [CHAP. XI 



has a polar line m. Thus there is established a one-to-one corre- 
spondence II between the lines of a and the lines of fi. This corre- 
spondence is a collineation. ' For consider a pencil of linen [I] in a. 
The conies J 2 determined by it form a pencil. Hence the point-pairs 
in which they meet c are an involution. Hence the conies 7> a deter- 
mined by the point-pairs form a pencil, and hence the lines [w] form 
a pencil. Since every line I meets its corresponding line m on c, the 
correspondence II is not only a collineation but is a pernpectivity, 
of which let the center be (7. Any two corresponding lines I and m 
are coplanar with C. Hence the polar planes of wiih regard to 
qiiadrics through 1, 2, 3, 4, 5, 6 are the planes on G. 

This was on the assumption that no two of the points 1, 2, 3, 4, 5, 6 
are collinear with 0. If two are collinear with 0, every polar plane 
of must pass through the harmonic conjugate of with regard to 
them. This harmonic conjugate may be taken as the point C. 

Now if nine points are given, no four being in the same plane, the 
notation may be assigned so that the planes a 123, /? = 456, 7 = 780 
are such that none of their lines of intersection a = 0% "b ~ yet, c = a/3 
contains one of the nine points. Let be the point <xj3y (or a point 
on the line aft if a, ft, and 7 are in the same pencil). By the argu- 
ment above the polars of with regard to all quadrics through the 
six points in a and yS must meet in a point C. The polars of with 
regard to all quadrics through the six points in ft and 7 must simi- 
larly pass through a point A, and the polars with regard to all quad- 
rics through the six points in 7 and a must pass through a point /?. 

If A, B, and G are not collinear, the plane <o=xAnC must be the 
polar of with regard to any quadric through the nine points. The 
plane <a> meets a, ft, and 7 each in a line which must be polar to 
with regard to the section of any sucli quadric. But this determines 
three conies A* in a, B* in y8, and (7 2 in 7, which meet by pairs in 
three point-pairs on the lines a, 5, c. Hence if a, /3, 7 are not in the 
same pencil, it follows, by Corollary 2, that there is a unique quadric 
through the nine points. If a, /9, 7 have a line in common, the three 
conies A 2 , z , C* meet this line in the same point-pair. Consider a 
plane <r through which meets the conies A*, I? 2 , C 2 in three point- 
pairs. These point-pairs are separated harmonically by and the 
trace, $, oh <r of the plane G>. Hence they lie on a conic 7> a , which, 
with A 2 and P 2 , determines a unique quadric. The section of this 



ion, iixij LINEAH DEPENDENCE 0V LINES 311 

quadrie by the plane 7 lias in common with # 2 two point-pairs and 
the polar pair 0, a. Hence the quadric has G 1 as its section by 7. 

Tu case A, 7?, and G are collinear, there is a pencil of planes CD which 
meet them. There is thus determined a family of quadrics which is 
called a pwml and is analogous to a pencil of conies. In case A, B, 
and G coincide, there is a bundle of possible planes a> and a quadric is 
determined for each one. This family of quadrics is called a bundle. 
Without inquiring at present under what conditions on the points 
1, 2, , 9 these cases can arise, we may state the following theorem : 

THEOREM 1 2. Through nine points no four of which are collinear 
there JMMMS one quadric surface or a pencil of yuadrics or a bundle 
of qit,adric8. 

EXERCISES 

1. The lines joining homologous points of a projective conic and straight 
lino form a rcguliiH, provided the line meets the conic and is not coplanar 
with it, and their point of intersection is self -corresponding. 

2. State the duals of Theorems 11 and 12. 

3. Show that two (proper or improper) 'conjugate reguli pass through two 
(ionics in different planes having two points (proper or improper or coincident) 
in common and through a point not in the plane of either conic. Two such 
conies and a point not in either plane thus detenninc one quadric surface. 

4. Show how to construct a regulus passing through six given points 
and a given line. 

106. Linear dependence of liaes. DEFINITION. If two lines are co- 
plauar, the lines of the flat pencil containing them both are said to 
be linearly dependent on them. If two lines are skew, the only lines 
linearly dependent on them are the two lines themselves. On three 
skew lines are linearly dependent the lines of the regulus, of which 
they are rulers. If l v l^ - , l n are any number of lines and m v m z ,--, m k 
are lines such that ^\ is linearly dependent on two or three of l v Z 2 , - - - , Z n , 
and w 2 is linearly dependent on two or three of l v 1 2 , , l n , m v and 
so on, m k being linearly dependent on two or three of l v l v , l n , m lt 
m< 2 , - , tn, kmml , then m k is said to be linearly dependent on l v l%> , l^ 
A set of n lines no one of which is linearly dependent on the n~l 
others is said to be linearly independent. 

As examples of these definitions there arise the following cases of 
linear dependence of lines on three linearly independent lines which 
may be regarded as degenerate cases of the regulus. (1) If lines a 



312 FAMILIES OF LINES . [OIIAP.XI 

and I intersect in a point P, and a line c skew to both of them meets 
their plane in a point Q, then in the first place all lines of the pencil 
ctb are linearly dependent on a, &, and e ; since the line QP is in this 
pencil, all lines of the pencil determined by QP and c are in the sot. 
As these pencils have in common only the line QP and do not con- 
tain three mutually skew lines, the set contains no other linen. 
Hence in this case the lines linearly dependent on a, "b, c are the flat 
pencil ab and the flat pencil (c, QP). (2) If one of the lines, as a, meets 
both of the others, which, however, are skew to each other, the set of 
linearly dependent lines consists of the flat pencils ab and UG. This 
is the same as case (1). (3) If every two intersect but not all in the 
same point, the three lines are coplanar and all lines of their plane 
are linearly dependent on them. (4) If all three intersect in the same 
point and are not coplanar, th bundle of lines through their common 
point is linearly dependent on them. The case where all three are 
concurrent and coplanar does not arise because three such lines are 
not independent. 

This enumeration of cases may be summarized as follows : 

THEOREM 13. DEFINITION. The set of all lines linearly dependent 
on three linearly independent lines is either a regidus, or a "bundle, of 
lines, or a plane of lines, or two flat pencils having diffc-rcnt centers 
and planes hot a common line. The last three sets of lines are called 
degenerate reguli. 

DEFINITION. The set of all lines linearly dependent on four linearly 
independent lines is called a linear congruence. The set of all linen 
linearly dependent on five linearly independent lines is called a linear 
complex* 

107. The linear congruence. Of the four lines a, ?;, 0, d upon 
which the lines of the congruence are linearly dependent, ft, v, d 
determine, as we have just seen, either a regulus, or two Hat pencils 
with different centers and planes but with one common lino, or a 
bundle of lines, or a plane of lines. The lines 7>, 0, d can of courso be 
replaced by any three which determine the same rcgulus or degen- 
erate regulus as &, c, d. 

* The terms congruence and complex are general terms to denote two- and three- 
parameter families of lines respectively, For example, all lines mooting a curve or 
all tangents to a surface form a complex, while all lines meeting two curves or all 
common tangents of two surfaces are a congruence. 



i<>7] THE LINEAll CONGRUENCE 313 

So in case 5, c, d determine a nondegenerate regulus of which a is 
not a directrix, the congruence can be regarded as determined by four 
mutually skew lines. In case a is a directrix, the lines linearly de- 
pendent on a, ?>, c, d clearly include all tangent lines to the regulus 
bud, whose points of contact are on a. But as a is in a flat pencil 
with any tangent whose point of contact is on a and one of the 
rulers, the family of lines dependent on a, 5, c, d is the family de- 
pendent on ft, c, d and a tangent line which does not meet &, c, d. Hence 
in either case the congruence is determined by four skew lines. 

If one of the four skew lines meets the regulus determined by the 
other three in two distinct points, JP, Q, the two directrices p, q 
through these points meet all four lines. The line not in the regulus 
determines with the rulers through P and Q, two flat pencils of lines 
which join P to all the points of q, and Q to all the points of p. 
From this it is evident that all lines meeting loth p and q are linearly 
dtyenflvit on Ike given four. For if 1\ is any point on p, the line 
I\Q and the ruler through P l determine a flat pencil joining P to 
all the points of q; similarly, for any point of q. No other lines 
can be dependent on them, because if three lines of any regulus 
meet p and q, so do all the lines. 

If one of the four skew lines is tangent to the regulus determined 
by the other three in a point P, the family of dependent lines in- 
cludes the regulus and all lines of the flat pencil of tangents at P. 
Hence it includes the directrix p through P and hence all the tangent 
lines whose points of contact are on p. By Theorem 6 this family 
of lines can be described as the set of all lines on homologous pairs 
in a certain projectivity II between the points and planes of p. Any 
two lines in this set, if they intersect, determine a flat pencil of lines 
in the set. Any regulus determined by three skew lines I, m, n of 
the set determines a projectivity between the points and planes on p, 
but this projectivity sets up the same correspondence as II for the 
three points and planes determined by I, m, and n. Hence by the 
fundamental theorem (Theorem 17, Chap. IV) the projectivity deter- 
mined by the regulus Imn is the same as H, and all lines of the 
regulus are in the set. Hence, when one of four skew lines is tangent 
to the rqjulus of the other three, the family of depende?it lines consists 
of a regulus and all lines tangent to it at points of a directrix. The 
directrix is itself in the family. 



314 



FAMILIES OF LINES 



F. XI 



If no one of the four skew lines meets the regulas of the other 
three in a proper point, we have a case studied more fully below. 

In case 5, c } d determine two flat pencils with a common line, u. 
may meet the center A of one of the pencils. The linearly dependent 
lines, therefore, include the bundle whose center is A. The plane of 
the other flat pencil passes through A and contains throe noucon- 
current lines dependent on a, &, e, d. Hence the family of lines also 
includes all lines of this plane. The family of all lines through a 
point and all lines in a plane containing this point has evidently 
no further lines dependent on it. This is a degenerate case of a con- 
gruence. If a is in the plane of one of the flat pencils, we have, by 
duality, the case just considered. If a meets the common line of the 
two flat pencils in a point distinct from the centers, the two flat 
pencils may be regarded as determined by their common line d r and 
by lines V and c f , one from each pencil, not meeting a. Hence the 
family of lines includes those dependent on the regulus aW and its 
directrix d f . This case has already been seen to yield the family of all 
lines of the regulus aW and all lines tangent to it at points of d( 




Fia. 112 

If a does not meet the common line, it meets the plants of tho 
two pencils in points and D. Call the centers of the pencils A and 
B (fig. 112). The first pencil consists of the lines dependent on AD 
and AB, the second of those dependent on An and BC. AH (ID is 
the line a, the family of lines is seen to consist of the lines which 
are linearly dependent on AB, BC, CD, DA. Since any point of Bl) 
is joined by lines of the family to A and C, it is joined by lines of 



107 J THE LINEAll CONGRUENCE 315 

tho family to every point of AC. Hence this case gives the family 
of all lines meeting both AC and BD. 

li\ caiso ft, 0, d determine a bundle of lines, a, being independent of 
them, does not pass through the center of the bundle. Hence the 
family of dependent lines includes all lines of the plane of a and the 
center of the bundle as well as the bundle itself. 

Lastly, if &, c, d are coplanar, we have, by duality, the same case as 
if I, e, d were concurrent. We have thus proved 

THEOREM 14. A linear congruence is cither (1) a set of lines 
linearly dependent on four linearly independent skew lines, such that 
no otic of them meets tlic rcgulus containing the other three in a proper 
point; or (2) it is the set of all lines meeting two skew lines; or (3) 
it is the set of all rulers and tangent lines of a given regulus which 
meet a Ji<ml directrix of the regulus ; or (4) it consists of a bundle 
of lines and a plane of lines, the center of the bundle being on the 
plane. 

DEFINITION. A congruence of the first kind is called elliptic; of the 
second kind, hyperbolic ; of the third kind, parabolic ; of the fourth 
kind, degenerate. A line which has points in common with all lines 
of a congruence is called a directrix of the congruence. 

COHOLLAKY. A parabolic congruence consists of all lines on corre- 
sponding points and planes in a projectivity between the points and 
planes on a line. The directrix is a line of the congruence. 

To study the general nondegenerate case, let us denote four linearly 
independent and mutually skew lines on which the other lines of the 
congruence depend by a, b, c, d, and let TT I and 7r 2 be two planes in- 
tersecting in a. Let the points of intersection with ^ and ?r 2 of b, c, 
and d be 7^, C v and D l and J? 2 , O v and > z respectively. By letting 
the complete quadrilateral a, B&, G^ D^ correspond to the 
complete quadrilateral a, J? 2 (7 2 , 2 Z> 2 , Z> 2 # 2 , there is established a 
protective collineation II between the planes ir^ and TT Z in which 
the lines &, a, d join homologous points (fig. 113). 

Among the lines dependent on a, b, c, d are the lines of the reguli 
ale, acd, adb, and all reguli containing a and two Knes from any 
of these three reguli. But all such reguli meet ^ and TT Z in lines 
(e.g. B^D V ^ 2 # 2 ) because they have a in common with ^ and 
7T 2 , Furthermore, the lines of the fundamental reguli join points 



316 



FAMILIES OF LIKES 



which correspond in II (Theorem 5 of this chapter and Theorem 18, 
Chap. IV), Hence the reguli which contain a and linos nhown by 
means of such reguli to be dependent on a, i, c, d arc those gen- 
erated by the projectivities determined by II between lines of TT JL 
and 7r 2 . 

d fi* 



\ 




rio. 113 

Now consider reguli containing triples of the lines already shown 
to be in the congruence, but not containing a. Three such linos, /, 
m, n, join three noncollinear points L v M lt ^ of ir l to the points 
L 2 , M ZJ jV 2 of 7r 3 which correspond to them in the collineation II. The 
regulus containing Z, m> and n meets TT^ and TT S in two conies which 
are protective in such a way that L v M v N^ correspond to &, Jf 2 , JV 9 . 
The projectivity between the conies determines a projectivity between 
the planes, and as this projectivity has the same effect an II on the 
quadrilateral composed of the sides of the triangle L^M^ and the 
line a, it is identical with II. Hence the lines of the regulus I, m, n 
join points of 7r t and 7r 2 which are homologous under II and are 
therefore among the lines already constructed. 

Among the lines linearly dependent on the family thus far con- 
structed are also such as appear in flat pencils containing two inter- 
secting lines of the family. If one of the two lines is a, the other 
must meet a in a double point of the projectivity determined on a by 
II. If neither of the two lines is a, they must meet 7r a and 7T 3 , the 
first in points JFJ, JEJ and the second in points Q v <? 2 , and these four 



107J THE LINEAK CONGRUENCE 317 

points are clearly distinct from one another. But as the given lines of 
the congruence, JJJJ and Q^, intersect, so must also the lines P^ and 
Wa ' rrr \ an( l ^a intersect, and the projectivity determined between 
l[Qi and Jj() a by II is a perspectivity. Hence the common point of 
./^ and JJ(J a IH a point of # and is transformed into itself by II. 
Hence, if lines of the family intersect, II has at least one double point 
on a, which means, by 105,* that the line a meets the regulus led 
and the congruence has one or two directrices. Thus two lines of a 
nondcgenerate congruence intersect only in the parabolic and hyper- 
bolic cases ; and from our previous study of these cases we know that 
the lines of a congruence through a point of intersection of two lines 
form a flat pencil. 

We have thus shown that all the lines linearly dependent on 
a, I, c, d, with the exception of a flat pencil at each double point of 
the projectivity on a, are obtained by joining the points of TT X and 7r 2 
which are homologous under II. From this it is evident that any four 
linearly independent lines of the congruence could have been taken 
as the fundamental lines instead of a, 6, c, d. These two results are 
summarized as follows : 

THKOKKM 15. All the lines of a linear congruence are linearly 
dependent on any linearly independent four of its lines. No lines not 
in the congruence are linearly dependent on four such lines. 

THEOREM 16. If two planes meet in a line of a linear congruence 
and neither contains a directrix, the other lines of the congruence meet 
the planes in homologous points of a projectivity. Conversely, if two 
planes are protective in such a way that their line of intersection cor- 
responds to itself, the lines joining homologous points are in the same 
linear congruence. 

* If there are two double points, J0, JP, on a, the conic BiCiDiEF must be trans- 
formed by H into -the conic BaOtDtEFi and the lines joining corresponding points of 
these conies must form a regulus contained in the congruence. As JS? and 3? are 
on lines of the regulus 6c<Z, there are two directrices jp, q of this regulus which 
meet *$ and JP respectively. The lines p and g meet all four of the lines a, &, c, d. 
Hence they meet all lines linearly dependent on a, 6, c, d. 

In the parabolic case the regulus bed must be met by a in the single invariant 
point IT of the parabolic projectivity on a, because the conic tangent to a at IT and 
passing through BtCiDi must be transformed by II into the conic tangent to a at JET 
and passing through JfeCiD* ; and the'lines joining homologous points of these conies 
must form a regulus contained in the congruence. As JET, a point of a, is on a line 
of the regulus &ctf, there is one and only one directrix p of this regulus which meets 
all four of a, 6, c, d and hence meets all lines of the congruence. 



318 FAMILIES OF LINES [CHAIN XI 

The dual of Theorem 16 may be stated in the following form: 
THEOREM 17. From two points* on the stmw line, of a, linear congru- 
ence the latter is projected Ity two projecMoe "biuutlM of pitmen, (^in- 
versely, two bundles of planes projectile in such a wj/ Unit the. line, 
joining their centers is self -corresponding, generate a Iw.ar MnyruMiec. 

DEFINITION. A regulus all of whose rulers are in, a congruence is 
called a regulus of-the congruence and is said to le in or to le con- 
tained in the congruence. 

COROLLARY. If three lines of a regulus are in a congruence, the 
regulus is in the congruence. 

In the hyperbolic (or parabolic) case the regulus led (in the notation 
already used) is met by a in two points (or one point), its points of 
intersection with the directrices (or directrix). In the elliptic case the 
regulus bed cannot be met by a in proper points, because if it were, 
the projectivity II, between TT X and 7r 2 , would have these points as 
double points. Hence no line of the congruence meets a regulus of 
the congruence without being itself a generator. Hence through each 
point of space, without exception, there is one and only one line of 
the congruence. The involution of conjugate, points of the regulus 
led on the line a is transformed into itself by II, and 'the same must 
be true of any other regulus of the congruence, if it does not con- 
tain a. Since there is but one involution transformed into itself by a 
noninvolutoric projectivity on a line (Theorem 20, Chap. VIII), we 
have that the same involution of conjugate points is determined on 
any line of the congruence by all reguli of the congruence whicli do 
not contain the given line. This is entirely analogous to the hyper- 
bolic case, and can be used to gain a representation in terms of proper 
elements of the improper directrices of an elliptic congruence* 

The three kinds of congruences may be characterized as follows : 

THEOREM 18. In a parabolic linear congruence each line is tangent 
at a fixed one of its points to all reguli of the congruence of which it u 
not a ruler. On each line of a hyperbolic or elliptic congruence all reguli 
of the congruence not containmg the given line determine the same 
involution of conjugate points. Through each point of space there is 
one and only one line of an elliptic congruence, for hyperbolic and 
parabolic congruences this statement i$ true except for points on a 
directrix. 



i<>7, iosj TJIE LINEAR COMPLEX 319 

EXERCISES 

1. All lines of a congruence can bo constructed from four lines by means 
of reguli all of which have two given lines in counnon. 

2. (liven two involutions (both having or both not having double points) 
ou two skew line,H. Through each point of apace there are two and only two 
lines which are axes of pern] activity projecting one involution into the other, 
i.e., Much that two planes through conjugate pairs of the first involution pass 
through a conjugate pair of the second involution. These lines constitute 

tWO COngrUetKHiH. 

3. All lines of a congruence meeting a line not in the congruence form 
a regular, 

4. A linear congruence i self-polar with regard to any regulus of the 
congruence. 

5. A degenerate linear congruence consists of all linos meeting two inter- 
secting linos. 

108* The linear complex. THTCOKEM 19. A linear complex con- 
mix of ail linw linearly dependent on the edges of a simple skew 



Proof, By definition ( 106) the complex consists of all lines 
linearly dependent on five independent lines. Let a be one of these 
which does not meet the other four, V, c f , $, e f . The complex consists of 
all linos dependent on # and the congruence Vctdld. If this con- 
gruence is degenerate, it consists of all lines dependent on three sides 
o a triangle ode and a line I not in the plane of the triangle 
(Theorems 14, 15). As 6 may be any line of a bundle, it may be 
chosen so as to meet a c may be chosen so as to meet I, and e may 
be so chosen as to meet a. Thus in this case the complex depends 
on five lines a, J, c, d, e not all coplanar, forming the edges of a simple 
pentagon. 

If the congruence is not degenerate, the four lines 6", c rr , d", e" upon 
which it depends may (Theorem 15) be chosen so that no two of 
them intersect, but so that two and only two of them, W and e n , 
meet a. Thus the complex consists of all lines linearly dependent 
on the two flat pencils aW and ae" and the two lines c" and d". Let 
& and e be the lines of these pencils (necessarily distinct from each 
other and from a) which meet c n and d" respectively. The complex 
then consists of all lines dependent on the flat pencils e&&, "be", ae, ed". 

* The edges of a simple skew pentagon are five lines in a given order, not all 
coplanar, each line intersecting its predecessor and the last meeting the first. 



20 FAMILIES OF LINKS [(iiur. xi 

.Finally, let c and d be two intersecting linos distinct from // ami f, 
which are in the pencils bo n and #F'. The complex consists of all lines 
linearly dependent on the flat pencils (tf> 9 be-, cd, <l< l t w. Not all tlio 
vertices of the pentagon dbcdc can be ooplanar, because then all the 
lines would be in the same degenerate congruence. 

THEOREM 20. DEFINITION. There are two dawn of wm>j*lc t m #m'h 
that all contylettes of cither clans are projecti'veli/ MjKiwtleuf. A c<nn, 
plc& of one class consists of a line and all linw of xptw, which, meet 
it. These are called special complexes. A complex of the, ofhe.r cJtttttt 
is catted general. No four vertices of a pentagon which determines it 
are coplanar. 

Proof. Given any complex, by the last theorem there is at least 
one skew pentagon abode which determines it. If there is a line / 
meeting the five edges of this pentagon, this line must meet all lines 
of the complex, because any line meeting three linearly independent 
lines of a regains (degenerate or not) meets all lines of it. Moreover, 
if the line I meets a and I as well as c and rf, it must either join 
their two points of intersection or be the line of intersection of their 
common planes. If I meets e also, it follows in either case thai four 
of the vertices of the pentagon are coplanar, two of them being on c. 
(That all five cannot be coplanar was explained at the end of the 
last proof.) Conversely, if four of the five vertices of the skew 
pentagon are coplanar, two and only two of its edges are not in this 
plane, and the line of intersection of the plane of the two edges with 
the plane of the other three meets all five edges. 

Hence, if and only if four of the five vertices are wplwmr, there. r,r- 
ists a line meeting the Jive linen. Since any two skew pentagons are pro- 
jectively equivalent, if no four vertices are coplanar (Theorem 1U, 
Chap. Ill), any two complexes determined by such pentagons are 
protectively equivalent. Two simple pentagons are also equivalent 
if four vertices, but not five, of each are coplanar, because any simple 
planar four-point can be transformed by a collineation of space into 
any other, and then there exists a collineation holding the plane 
of the second four-point pointwise invariant and transforming any 
point not on the plane into any other point not on the plane. There- 
fore all complexes determined by pentagons of this kind are pro jec- 
tively equivalent. But these are the only two kinds of skew pentagons. 
Hence there are two and only two kinds of complexes. 



f 



THE LINE All COMPLEX 



321 



hi ease four vertices of the pentagon are coplanar, we have seen 
that there is a line / meeting all its edges. Since this line was 
detiH'mmed as the intersection of the plane of two adjacent edges 
with the plane of the other three, it contains at least two vertices. 
It ('.winot contain three vertices because then all five would be 
<*.o]>luiwr. AH one of the two planes meeting on I contains three 
independent lines, all lines of that plane are lines of the complex. 
Tim line / ilnel is therefore in the complex as well as the two lines 
of the other plane. Hence all lines of both planes are in the complex. 
II once all lines meeting / are in the complex. But as any regulus 
three of whose lines meet / has all its lines meeting Z, the complex 
IioM the requirements stated in the theorem for a special complex. 




Fi<*. 114 



A more definite idea of the general complex may be formed as 
follows. Let piPiPdW* (% 11 ' i ) ke a simple pentagon upon whose 
edges all lines of the complex are linearly dependent. Let q be the 
line of the flat pencil jp 8 jp 4 which meets p v and let ft be the point of 
intersection of q and p r Denote the vertices of the pentagon by P 12 , 
& ^H> / ?&i> ^ ie subscripts indicating the edges which meet in a 
given vertex. 

The four independent lines p^^y determine a congruence of lines 
all of which are in the complex and whose directrices are # = jftj^ 8 
and a f a* 7f 2 / 8 V Iu like manner, gp^p 6 p^ determine a congruence whose 
directrices are l RP^ and V^=P M P 6l . The complex consists of all 
lines linearly dependent on the lines of these two congruences. The 



322 



FAMILIES OF LINES 



[('HAP. XI 



directrices of the two congruences intersect at R and /,5 t respectively 
and determine two planes, al = p and aW^vr, which meet OH <y. 

Through any point P of space not on /> or TT then 1 , are two linos 
1 9 m, the first meeting a and a r , and the second meeting h and // 
(fig. 115). All lines in the ilat pencil Ini are in the complex by deii- 
nition. This fiat pencil meets p and TT in two perspective ranges of 




FIG. 115 



points and thus determines a projectivity between the flat pencil a<b 
and the flat pencil a'V 9 in which a and a f , & and V correspond and q 
corresponds to itself. The projectivity thus determined between the 
pencils ab and aV is the same for all points P> because a< 9 ft, q always 
correspond to a r , V, q f . Hence the complex contain,* all lines in the 
flat pencils of lines which meet homologous lines in the proJMtiwity 

determined ly 

alq -ft a'Vq. 

Denote this set of lines ly S. We have seen that it has the properly 
that all its lines through a point not on p or TT are coplanar. If a 
point P is on p but not on q, the line Pit has a corresponding linojp/ 
in the pencil a'V and hence S contains all lines joining /* to points 
of p f . Similarly, for points on TT bxit not on q. By duality every plane 
not on q contains a flat pencil of lines of S. 

Each of the flat pencils not on q has one line meeting q. Hence 
each plane of space not on q contains one and only one line of S 
meeting j. Applying this to the planes through P H not contain- 
ing 2, we have that any line through P^ and not on p is not in the 



i 108 J THE LINEAR COMPLEX 323 

sot S. Let I be any such line. All lines of S in each plane through 
/ form a flat pencil P, and the centers of all these pencils lie on a line 
/', because all lines through two points of I form two flat pencils each 
of which contains a line from each pencil P. Hence the lines of S 
meeting I form a congruence whose other directrix V evidently lies on 
p. The point of intersection of V with q is the center of a flat pencil 
of lines of S all meeting L Hence all lines of the plane lq form a flat 
pencil. Since I is any line on P 34 and not on TT, this establishes that 
uaeh plane and, by duality, each point on q, as well as not on q, con- 
tains a Hat pencil of lines of S. 

We can now prove that S contains no lines not in the complex. 
To do 80 we have to show that all lines linearly dependent on lines 
of S are in S. If two lines of S intersect, the flat pencil they deter- 
mine is by definition in S. If three lines m v m 29 m B of S are skew to 
one another, not more than two of the directrices of the regulus con- 
taining them are in S. For if three directrices were in S, all the tan- 
gent line** at points of these three lines would be in S, and hence any 
piano would contain three nonconcurrent lines of S. Let I be a 
directrix of the regulus m^m^m^ which is not in S. By the argu- 
ment made in the last paragraph all lines of S meeting I form a con- 
gruence. But this congruence contains all lines of the regulus mjn^n^ 
and hence all lines of this regulus are in S. Hence the set of lines ,S 
is identical with the complex. 

TIIKOREM 21 (SYLVESTER'S THEOREM*). If two protective flat pencils 
with different centers and planes have a line q in common which is 
Mlfaorretipontliny, all lines meeting homologous pairs of lines in these 
two pencils are in the same linear complex. This complex consists of 
these lines together with a parabolic congruence whose directrix is q. 

Proof. This has all been proved in the paragraphs above, with the 
exception of the statement that q and the lines meeting q form a 
linear congruence. Take three skew lines of the complex meeting q ; 
they determine with q a congruence C all of whose lines are in the 
complex. There cannot be any other lines of the complex meeting q, 
because there would be dependent on such lines and on the congru- 
ence C all lines meeting q, and hence all lines meeting q would be in 
the given complex, contrary to what has been proved above. 

* Cf. Coinptes Kendus, Vol. LH (1861), p. 741. 



324 FAMILIES OF LINKS [CUAI>.XI 

Another theorem proved in the discussion above is: 

THEOREM 22. DEFINITION OF NULL MYWTKM. All the lhw& of a 
linear complex which pans through a point, /* lie, in <i plans. TT, and all 
the lines which lie in a plane TT jprwwf tlirowgh a 'point /*. In MM, of 
a special complex, exception must be made of the points and pht.nw on 
the directrix. The point P is called the null point of the plane TT and 
TT is called the null plane of P with regard to the wm-pliw. The cor- 
res^^ondence between the points and planes of space thus established is 
called a null system or null polarity. 

Another direct consequence, remembering that there are only two 
kinds of complexes, is the following : 

THEOREM 23. Any five linearly independent lines are in one and 
only one complex. If the edges of a simple pentagon are in a given 
complex, the pentagon is skew and its edges linearly independent. If 
the complex is general, no four vertices of a simple pentagon of its 
lines are eoplanar. 

THEOREM 24. Any set of lines, K, in space such that the lines of the 
set on each point of space constitute a flat pencil is a linear complex. 

Proof, (a] If two lines of the set K intersect, the set contains (ill 
lines linearly dependent on them, by definition. 

(6) Consider any line a not in the given set K. Two points A, B on 
a have flat pencils of lines of K on different planes ; for if the planes 
coincided, every line of the plane would, by (a), be a line of K. Hence 
the lines of K through A and B all meet a line a 1 skew to a. Prom 
this it follows that all the lines of the congruence whoso directrices 
are a, a' are ii^ K. Similarly, if & is any other line not in K but meet- 
ing a, all lines of K which meet & also meet another line ?/. More- 
over, since any line meeting a, &,and V is in K and hence also meets 
a 1 , the four lines a, a 1 , &, V lie on a degenerate regains consisting of the 
flat pencils aa r and W (Theorem 13). Let q (fig. 115) bo the common 
line of the pencils al and a!V. Through any point o sjwce not on one 
of the planes ab and a'U there are three eoplanar lines of K which 
meet q and the pairs aa r and W. Hence K consists of lines mooting 
homologous lines in the projectivity 



and therefore is a complex by Theorem 21. 



THE LINEAR COMPLEX 325 

OOROLLAKY. Any (1,1) correspondence between the points and the 
planes of space such that each point lies on its corresponding plane 
in a null system. 

THEOREM 25. Two linear complexes have in common a linear 
congruence. 

Proof, At any point of space the two flat pencils belonging to the 
two complexes have a line in common. Obviously, then, there are 
three linearly independent lines l v L 2 , Z 3 common to the complexes. 
All lines in the regulus / i y a are, by definition, in each complex. But 
as there are points or planes of space not on the regulus, there is a 
line 1 4 common to the two complexes and not belonging to this regulus. 
All lines linearly dependent on l v Z 2 , Z 3 , 1 4 are, by definition, common 
to the complexes and form a congruence. No further line could be 
common or, by Theorem 23, the two complexes would be identical 

COROLLARY 1. The lines of a complex meeting a line I not in the 
complex form a hyperbolic congruence. 

Proof. The line is the directrix of a special complex which, by the 
theorem, has a congruence in common with the given complex. The 
common congruence cannot be parabolic because the lines of the first 
complex in a plane on I form a flat pencil whose center is not on I, 
since I is not in the complex. 

COROLLARY 2. The lines of a complex meeting a line I of the com- 
plex form a parabolic congruence. 

Proof. The centers of all pencils of lines in this congruence must 
be on I because I is itself a line of each pencil. 

DEFINITION. A line I is a polar to a line V with regard to a 
complex or null system, if all lines of the complex meeting I also 
meet V. 

COROLLARY 3. If I is polar to I', V is polar to I A line is polar 
to itself, if and only if it is a line of the complex. 

THEOREM 26. A null system is a protective correspondence "between 
the points and planes of space. 

Proof. The points on a line I correspond to the planes on a line V 
by Corollaries 1 and 2 of the last theorem. If I and V are distinct, 
the correspondence between the points of I and planes of V is a per- 
spectivity. If l*=V, the correspondence is projective by the corollary 
of Theorem 14. 



320 FAMILIES OF LINES l<'n.u>. XI 

EXERCISES 

1. If a point P is on a plane p, the null piano TT of /'is on the null point It of p. 

2. Two pairs of lines polar with regard to the same, null system are uhvuys in 
the same regulus (degenerate, if a line of one, pair meets a lino of the other pair). 

3. If a lino / meets a line m, the polar of I moots the polar of nt. 

4. Pairs of lines of the, regains in Ex. "2 which are, polar with regard to 
the complex are met by any directrix of the regains in pairs of points of an 
involution. Tims the complex determines an involution among the linos of 
the regnlus. 

5. Conversely (Theorem of Chasie,s), the lines meeting conjugate pairs of 
lines in an involution on a regulus are in the same complex. Show that 
Theorem 21 is a special case of this. 

6. Find the lines common to a linear complex and a regulus not in the 
complex. 

7. Three skew lines , I, m determine one and only one complex contain- 
ing k and having I and m as polars of each other. 

8. If the number of points on a line is + 1, how many reguli, how many 
congraences, how many complexes are there in space? How many lines are 
there in each kind of regulus, congruence, complex ? 

9. Given any general complex and any tetrahedron whose faces are not 
null planes to its vertices. The null planes of the vertices constitute a second 
tetrahedron whose vertices lie on the planes of the first tetrahedron. The 
two tetrahedra are mutually inscribed and circumscribed each to the, other* 
(cf. Ex. G, p. 105). 

10. A null system is fully determined by associating with the tlirco vertices 
of a triangle three planes through these vertices and having their ono common 
point in the plane of the triangle but not on one of its sides. 

11. A tetrahedron is self-polar with regard to a null system if two opposite 
edges are polar. 

12. Every line of the complex determined by a pair of Mobius tolrahodra 
meets their faces and projects their vertices in projectivo throws of points and 
planes. 

13. If a tetrahedron T is inscribed and circumscribed to 7\ and also to 7^, 
the lines joining corresponding vertices of !Z\ and T, 2 and the linos of intersec- 
tion of their corresponding planes are all in the same complex. 

14. A null system is determined by the condition that two pairs of lines 
of a regulus shall be polar. 

15. A linear complex is self-polar with regard to a regnlus all of whoso 
lines are in the complex. 

16. The lines from which two projective pencils of points on skew linos 
are projected by involutions of planes are all in the same complex* Dunllxo. 

* This configuration was discovered by Mobius, Journal fur Mathomatik, Vol. HI 
(1828), p. 273. Two tetrahedra in this relation are known as Mobius tctrahwlra, 



iw] LINE COORDINATES 327 

109. The Pliicker line coordinates. Two points whose coordinates are 



determine a line L The coordinates of the two points determine six 
numbers 



which are known as the Pluckw coordinates of the line. Since the 
coordinates of the two points are homogeneous, the ratios only of the 
numbers p lf are determined. Any other two points of the line deter- 
mine the same set of line coordinates, since the ratios of the jp y 's are 
evidently unchanged if (& v x, # 3 , # 4 ) is replaced by (x l + \y v o? 2 - 
The six numbers satisfy the equation* 

0. 



Thin is evident on expanding in terms of two-rowed minors the 
identity 



x l 



2/4 



= o. 



(Jbnversely, if any six numbers, jp , are given which satisfy Equa- 
tion (1), then two points P^fa, %%> 8 , 0), Q~(y v 0, y 8 , y 4 ) can be 
determined such that the numbers p tf are the coordinates of the line 
PQ. To do this it is simply necessary to solve the equations 



which are easily seen to be consistent if and only if 



Hence we have 

THEOREM 27. Every line of space determines and is determined 
ly the ratios of six niwibers jp u , jp 18 , jp 14> p$v p&, p 2S subject to the 



* Notice that in Equation (1) the number of inversions in the four subscripts of 
any term is always even. 



328 

condition 



FAMILIES OF LINES [<',HAI>, XI 

! 0, Mtch thttt. if (,','p ,-r5 a , .", , 4 ) <r/f 



it; 8 ^'4 

^ y/4 



i'w 

^=~" 



J'u- 



'*'4 



?/,! 



OOROLLAEY. Four independent coordinates determine, a line. 
In precisely similar manner two planes (u, v n^ w s , v^) and (v\, 7? a , v rt , v 4 
determine six numbers such that 



2 12 == 



M, 



The quantities 2 V . satisfy a theorem dual to the one just proved for 
the jp v 's. 

THEOREM 28. I7^e jp and % coordinates of a line are connected 1>y 
the equations jp 12 : _p ls : jp u : p 34 : ^ 42 : ^ 23 = J M : <7, 2 : JJT W : y la : </ u : y w . 

Proof. Let the j? coordinates he determined l)y the two points 
(x v XM x 3 , x^j (y v y z , y^ y 4 ), and the q coordinates by the two planes 
(v, v u# u a , u 4 ), (v v v> 2> -z; 8 , i? 4 ). These coordinates satisfy the four equations 

u^ + Ufa + Ufa + Ufa = 0, 
V& + vfa + vfa + v& = 0, 

^2/1 + ^2/2 + ^ 8 + ^4 = 0, 
^1^1 + ^8 + ^#8 + '^4^4 = ' 



Multiplying the first equation by # x and the second by W L and adding, 
we obtain 



In like manner, from the third and fourth equations wo obtain 



Combining the last two equations similarly, we obtain 



or, * = *. 

<?14 Pi* 

By similar combinations of the first four equations we find 



i* n u 



: Pu : 



: Si* : 



io, noj 



LINE COORDINATES 



329 



EXERCISE 

Given tlie tetrahedron of reference, the point (1, 1, 1, 1), and a line I, 
determine six sets of four points each, whose cross ratios are the coordinates 
of/. 

110. Linear families of lines. THEOREM 29. The necessary and 
tni,f/i('.ieti,t condition that two lines p and p f intersect, and hence are 
coplanar, is 



where p (j are the coordinates of p and p!j of p 1 . 

Proof. If the first line contains two points x and y, and the second 
two points x 1 and y r , the lines will intersect if and only if these four 
points are coplanar ; that is to say, if and only if 



= 



ffi ft ?/ 

./ ,/ ,/ 

'''i '^a '*'8 

';/' <?/' <?/' 

yi 2^2 ya 



uP* + J'sJ'u 



TIIEOHKM 30. Aflat pencil of lines consists of the lines whose coordi- 
nates are Xp,. ; +/*p^> if p und p f are two lines of the pencil. 

Proof, The lines p and p f intersect in a point A and are perspec- 
tive with a range of points <7+XZ>. Hence their coordinates may be 
written 



which may be expanded in the form 



THEOREM 31. The lines whose coordinates satisfy one linear 
eqitation 



form a linear complex. Those whose coordinates satisfy two independ- 
ent linear equations form a linear congruence, and those satisfying 
three independent linear equations form a regulus. Four independent 
linear equations are satisfied ly two (distinct or coincident] Iines 9 
which may 'be improper. 



330 



FAMILIES OF LINES 



[CHAP. XI 



Proof. If (\, Z^, J 8 , 6 4 ) is any point of space, the points (M V w a , ,f,,, ,f 4 ) 
which lie on lines through 1 19 J a , 6 8 , 6 4 satisfying (1) must satisfy 



KX.X^^X,^ 



( '4 '"'2 



= 0, 



or 



(2) 



^ A) t/; a 



4 



which is the equation of a plane. Hence the family of lines repre- 
sented by (1) has a flat pencil of lines at every point of space, and so, 
by Theorem 24, is a linear complex. 

Since two complexes have a congruence of common lines, two linear 
equations determine a congruence. Since a congruence and a complex 
have a regulus in common, three linear equations determine a regulus. 

If the four equations 



are independent, one of the four-rowed determinants of their cooili- 
cients is different from zero, and the equations have solutions of the 



If one of these solutions is to represent the coordinates of' a line, it 
must satisfy the condition 



which gives a quadratic equation to determine X//A. Hence, by Propo- 
sition K 2 , there are two (proper, improper, or coincident) lines whose 
coordinates satisfy four linear equations. 

COKOLLARY 1. TJie lines of a regulus are of the form 



where p r , p fr , p nr are lines of the reyulus. In like manner, (lie Un&x of 
a congruence are of the form 



* Cf . Bocher, Introduction to Higher Algebra, Chap. IV. 



>, m] LINE COORDINATES 331 

of tt complex of the form 



All of these formulas must le taken in connection with 



CouoLLAiiY 2. As a transformation from points to planes the null 
determined "by the complex whose equation is 

= 







/ -y xv m I C\ 

The first of these corollaries simply states the form of the solu- 
tions of systems of homogeneous linear equations in six variables. 
The second corollary is obtained by inspection of Equation (2) the 
coefficients of which are the coordinates of the null plane of the 
point (l v & 2 , 5 8 , & 4 ). 

Corollary 1 shows that the geometric definition of linear dependence of 
lim*H givon in thin chapter corresponds to the conventional analytic concep- 
tion of linear dependence. 

111. Interpretation of line coordinates as point coordinates in S 5 . 

It may be shown without difficulty that the method of introducing 
homogeneous coordinates in Chap. VII is extensible to space of any 
number of dimensions (cf. Chap. I, 12). 'Therefore the set of all sets 
of six numbers 

can be regarded as homogeneous point coordinates in a space of five 
dimensions, S 6 . Since the coordinates of a line in S s satisfy the 
quadratic condition 



they may be regarded as forming the points of a quadratic locus or 
spread,* L 4 a , in S 6 . The lines of a linear complex correspond to the 
points of intersection with this spread of an S 4 that is determined by 
one linear equation. The lines of a congruence correspond, therefore, 
to the intersection with L 4 2 of an S 3 , the lines of a regulus to the 

* This is a generalization of a conic section. 



332 FAMILIES OF LINES [<hui>. xi 

intersection with Lj 2 of au S 2 , and any pair of linos to the intersee- 
tion with L* of an S r 

Any point (^ p(. l9 p{, |4, }>&,}>&) of S c has as its polar*' S 4 , with 
regard to L 4 2 , 

(2) l*Ll\i + pJaPu + 2>*2>u + P/a^M + 1>l* Vi* + u ?' = (] > 



.which is the equation of a linear complex in the original S. r 
any point in S 5 cttn be thought of tin re.preMnting the, winylcs, of I 
represented Inj the points of S 5 in whitfi its poltir S 4 meets L.j*. 

Since a line is represented by a point on L 4 a , a special complex is 
represented by a point on L^ 2 , and all the lines of the special complex 
by the points in which a tangent S 4 meets L*. 

The points of a line, # + X&, in S 5 represent a set of complexes 
whose equations are 

(3) ( 



and all these complexes have in common the congruence common to 
the complexes a and &. Their congruence, of course, consists of the 
lines of the original S 3 represented by the points in which L* is met 
by the polar S 3 of the line a + X&. 

A system of complexes, a + \l, is called a pencil of ctMi'pliWx, and 
their common congruence is called its Iww or laml MMfrwnw. It 
evidently has the property that the null planes of any point with 
regard to the complexes of the pencil form an axial pencil whose 
axis is a line of the basal congruence. Dually, the null points of 
any plane with regard to the complexes of the pencil form a rnngo 
of points on a line of the basal congruence. 

The cross ratio of four complexes of a pencil may be defined as 
the cross ratio of their representative points in S 6 . From the. form of 
Equation (3) this is evidently the cross ratio of the four null planes 
of any point with regard to the four complexes. 

A pencil of complexes evidently contains the special complexes 
whose directrices are the directrices of the basal congruence. Ilenco 

* Equation (2) may be taken as the definition of a polar S 4 of a point with 
regard to L|. Two points are conjugate with regard to L* if the polar S 4 of one 
contains the other. The polar S 4 's of the points of an S t - (i =s 1, 2, &, 4) all have an 
S 4 _, in common which is called the polar S 4 ^ t - of the S;. These and other obvioiw 
generalizations of the polar theory of a conic or a regains we take for granted 
without further proof. 



1J J LIKE COOEDINATES 333 

there are two improper, two proper, one, or a flat pencil of lines which 
are the directrices of special complexes of the pencil. These cases 
arise as the representative line a + \l meets L 4 2 in two improper 
points, two proper points, or one point, or lies wholly on L 4 2 . Two 
points in which a representative line meets L 2 are the double points 
of an involution the pairs of which are conjugate with regard to L 2 . 

Two complexes p, p r whose representative points are conjugate 
with regard to L.* are said to be conjugate or in involution. They 
evidently satisfy Equation (2) and have the property that the null 
points of any plane with regard to them are harmonically conjugate 
with regard to the directrices of their common congruence. Any 
complex a is in involution with all the special complexes whose 
directrices are lines of a. 

Let ff t be an arbitrary complex and & 2 any complex conjugate to 
(in involution with) it. Then any representative point in the polar S 8 
with regard to L* of the representative line a^a 2 represents a complex 
conjugate to a l and a y Let cr, B be any such complex. The represent- 
ative points of a v a a , a s form a self-conjugate triangle of L 4 2 . Any 
point of the representative plane polar to the plane a^a^ with 
regard to L 4 2 is conjugate to a^a^ Let such a point be 4 . In like 
manner, a. $ and & can be determined, forming a self-polar 6-point of 
LJ*, the generalization of a self-polar triangle of a conic section. The 
Hix points are the representatives of six complexes, each pair of which 
in in involution. 

It can be proved that by a proper choice of the six points of refer- 
ence in the representative S e , the equation of L 4 2 may be taken as any 
quadratic relation among six variables. Hence the lines of a three- 
space may be represented analytically by six homogeneous coordinates 
subject to any quadratic relation. In particular they may be repre- 
sented by (#i, a? 9 ,- , # 6 ), where 

= 0.* 



In this case, the six-point of reference being self-polar with regard 
to L*, its vertices represent complexes which are two by two in 
involution. 

* Those are known as Klein's coordinates. Most of the ideas in the present sec- 
tion are to be found in F. Klein, Zur Theorie der Liniencomplexe des ersten 
zwoiten Grades, Mathematische Annalen, Vol. II (1870), p. 198. 



334 FAMILIES OF LINKS Loiur. xi 

EXERCISES 

1. If a pencil of complcxe-H contains two special complexes, the. basal con- 
gruence of the pencil is hyperbolic or elliptic, according as the special com- 
plexes are proper or improper. 

2. If a pencil of linear complexes contains only a single special complex, 
the basal congruence is parabolic. 

3. If all the complexes of a pencil of linear complexes are special, tlw 
basal congruence is degenerate, 

4j. Define a pencil of complexes as the system of all complexes having a 
common congruence of lines and derive its properties synthetically. 

5. The polars of a line with regard to the complexes of a pencil form 
a regulus. 

6. The null points of two planes with regard to the complexes of a pencil 
generate two projective pencils of points. 

7. If C = 0, C'= 0, C"' = are the equations of three linear complexes 
which do not have a congruence in common, the equation C + AC" + i**C" = 
is said to represent a "bundle of complexes. The lines common to the three 
fundamental complexes C, C", C" of the bundle form a regulus, the con- 
jugate regulus of which consists of all the directrices of the special com- 
plexes of the bundle. 

8. Two linear complexes Sa, 7j fr/ = and Sfyjft/ = are in invohition if and 
only if we have 

+ G * As 



9. Using Klein's coordinates, any two complexes are given by 
and SMi 0. These two are in involution if 2a^ = 0. 

10. The six fundamental complexes of a system of Klein's coordinates 
intersect in pairs in fifteen linear congruences all of whose directrices are dis- 
tinct. The directrices of one of these congruences are lines of tho remaining 
four fundamental complexes, and meet, therefore, the twelve directrices of 
the six congruences determined by these four complexes. 



INDEX 



The numbers refer to pages 



Abolian group, 67 

Absoksa, 170 

Abstract science-, 2 

Addition, of points, 142, 231 ; theorems 
on, 142-144 ; other definitions of, 1C7, 
Em 3, 4 

Adjacent sides or vertices of simple 
ft-lino, 87 

Algebraic curve, 259 

Algebraic problem, 288 

Algebraic surface, 259 

Alignment, assumptions of, 10 ; consist- 
ency of assumptions of, 17; theorems 
of, for the, plane, 17-20 ; theorems of, 
for #-space, 20-24 ; theorems of, for 
4-space, 25, Ex. 4; theorems of, for 
n-spaco, 29-33 

Amodeo, E,, 120, 294 

Anharmonic ratio, 159 

Apollonius, 286 

Associative law, for correspondences, 
06 ; for addition of points, 143 ; for 
null tit >li cation of points, 146 

Assumption, H , 45; H , rOlo of, 81, 
261; of projectivity, 95; of projec- 
tivity, alternative forms of, 105, 106, 
Exs. 10-12 ; 298 

Assumptions, are necessary, 2; exam- 
ples of, for a mathematical science, 
2; consistency of, 3; independence 
of, 6 ; categoricalnoss of, 6 j of align- 
ment, 16; of alignment, consistency 
of, 17; of extension, 18, 24; of clo- 
sure, 24 ; for an n-space, 33 

Axial pencil, 55 

Axial perspectivity, 57 

Axis, of perspectivity, 36; of pencil, 
55 ; of perspective collineation, 72 ; of 
homology, 104; of coSrdinates, 169, 
191; of projectivity on conic, 218 

Bawo, of piano of points or lines, 55 ; of 
pencil of complexes, 332 

Bilinear equation, "binary, represents 
projectivity on a line, 156 ; ternary, 
represents correlation in a plane, 267 

Binary form, 251, 252, 254 

B&cuor, M., 156, 272, 289, 330 

Braikenriclge, 119 

Brian chon point, 111 

Brianchon's theorem, 111 



Bundle, of planes or lines, 27, 55; of 
conies, 297, Exs. 9-12; of quadrics, 
311 ; of complexes, 334, Ex. 7 

Burnside, W., 150 

Bussey, W. H,, 202 

Canonical forms, of collineations in 
plane, 274-276; of correlations in a 
plane, 281 ; of pencils of conies, 287- 
293 

Castelnuovo, G., 139, 140, 237, 297 

Categorical set of assumptions, 6 

Cayley, A., 52, 140 

Center, of perspectivity, 36 ; of flat pen- 
cil, 55 ; of bundle, 55 ; of perspective 
collineation in plane, 72 ; of perspec- 
tive collineation in space, 75; of 
homolpgy, 104; of coordinates, 170; 
of projectivity on conic, 218 

Central perspectivity, 57 

Characteristic constant of parabolic 
projectivity, 207 

Characteristic equation of matrix, 165 

Characteristic throw and cross ratio, of 
one-dimensional projectivity, 205, 211, 
Exs. 2, 3, 4 ; 212, Exs. 5, 7 ; of involu- 
tion, 206; of parabolic projectivity,, 
206 

Chasles, 125 

Class, notion of, 2 ; elements of, 2 ; re- 
lation of belonging to a, 2 ; subclass of 
a, 2 ; undefined, 15 ; notation for, 57 

Clebsch, A., 289 

Cogredient n-line, 84, Ex. 13 

Cogredient triangle, 84, Exs. 7, 10 

Collineation, defined, 71 ; perspective, in 
plane, 72 ; perspective, in space, 75 ; 
transforming a quadrangle into a 
quadrangle, 74 ; transforming a five- 
point into a five-point, 77 ; transform- 
ing a conic into a conic, 182 ; in plane, 
analytic form of, 189, 190, 268 ; be- 
tween two planes, analytic form of, 
190 ; in space, analytic form of, 200 ; 
leaving conic invariant, 214, 220, 235, 
Ex. 2; is the product of two polar- 
ities, 265; which is the product of 
two reflections, 282, Ex. 5 ; double ele- 
ments of, in plane, 271 ; character- 
istic equation of, 272 ; invariant figure 
of, is self-dual, 272 



335 



336 



INDEX 



Collineationw, types of, in plants 100, 
278; associated with two conies oC 
a pencil, i;U, EXH. 2, 4, 0; 135, 
Kx. 2 ; 130, Ex. 2 ; group of, in piano, 
208; represented by matriees, 208- 
270; two, not in general commuta- 
tive, 208 ; canonical forms of, 274- 
276 

Commutative C'orrospondoncc, 00 

Commutative group, 07, 70, Kx. 1; 228 

Commutative law of multiplication, 
148 

Commutative pro jectivi ties, 70, 210, 228 

Compass, constructions with, 240 

Complete Wrline, in plane, 37; on point, 38 

Complete w-plane, in space, 37 ; on point, 
38 

Complete n-point, in space, 30 ; in piano, 
37 

Complete quadrangle and quadrilat- 
eral, 44 

Complex, linear, 312; determined by 
skew pentagon, 319 ; general and spe- 
cial, 320 ; determined by two projec- 
tive flat pencils, 323 ; determined by 
live independent lines, 324; deter- 
mined by correspondence between 
points and planes of space, 324 ; null 
system of, 324 ; generated by involu- 
tion on regains, 320, Ex. 6 ; equation 
of, 329, 331 

Complexes, pencil of, 332; in involu- 
tion, 333 ; bundle of, 334, Ex. 7 

Concrete representation or application 
of an abstract science, 2 

Concurrent, 10 

Cone, 118 ; of lines, 109 ; of planes, 109 ; 
section of, by plane, is conic, 109; 
as degenerate case of quadric, 308 

Configuration, 38; symbol of, 38; of 
Desargues, 40, 51 ; quadrangle-quad- 
rilateral, 44; of Pappus, 98, 249; of 
Mobius, 320, Ex. 9 

Congruence, linear, 312 ; elliptic, hyper- 
bolic, parabolic, degenerate, 316; de- 
termined by four independent lines, 
317 ; determined by projective planes, 
317; determined by two ^complexes, 
325 ; equation of, 329, 330 

Conic, 109, 118 , theorems on, 109-140 ; 
polar system of, 120-124; equation 
of, 185, 245; projectiyity on, 217; 
intersection of lino with, 240, 242, 
240 ; through four points and tangent 
to line, 250, Ex. 8; through three 
points and tangent to two linos, 250, 
Ex. 9 ; through four points and moot- 
ing given line in two points harmonic 
with two given points, 250, Ex. 10; 
determined by conjugate points, 293, 
Ex. 2 ; 294, Exs. 3, 4 

Conic section, 118 



(Ionics, pencils and ranges of, 128-130, 
287-293; projeetive, 212, ,*K)4 

Conjugate groups, 209 

Conjugate pair of involution, 102 

Conjugate points (lines), with regard to 
conic, 122; on line (point), form invo- 
lution, 124; wifch regard to a pencil of 
monies, U)0, Bx. ;J ; 140, Kx, JU ; 293, 
Kx. 1 

Conjugate projeotivitios, 208; condi- 
tions for, 208, 209 

Conjugate subgroups, 211 

Consistency, of a sot of assumptions, # ; 
of notion of elements at infinity, 9; 
of assumptions of alignment, 17 

Construct, 45 

Constructions, linear (first degree), 230 ; 
of second degree, 245, 249-250, 
Exs.; of third and fourth degrees, 
294-290 

Contact, point of, of lino of lino conic, 
112; of second order between two 
conies, 134; of third order between 
two conies, 130 

Conwell, G. M., 204 

Coordinators, non homogeneous, of points 
online, 152; homogeneous, of points 
on lino, 103; nonhomogeneous, of 
points in plane, 109; iionhomogeno-. 
OIIK, of linos in plane, 170; homogene- 
ous, of point;* and linos in piano, 174; 
in a bundle, 179, Ex. 3; of quadran- 
gle-quadrilateral configuration, 181, 
Ex. 2; nonhomogoneous, in space, 
190; homogeneous, in space, 394; 
Plttokor'H lino, 327 ; Klein's lino, MM 

Coplanar, 24 

Copunctal, 10 

Correlation, between two-dimensional 
forms, 202, 203; induced, 202; be- 
tween two-dimensional forum deter- 
mined by four pairs o homologous 
elements', 204; which interchanges 
vortices and sides of triangle is polar- 
ity, 204 ; between two pianos, analytic 
representation of, 200, 207; repre- 
sented by ternary bilinear form, 207 ; 
represented by matrices, 270; double 
pairs of a, 278-281 

Correlations and duality, 20R 

Correspondence, as a logical term, 5j 
perspective, 12; (1, 1) of two liguros, 
35; general theory of, 04-0(1; idim- 
tioal, 05; inverse of, 05; period of, 
(JO; periodic or cyclic, 00; involutorio 
or reflexive, 00 ; perspective between 
two pianos, 71 ; quadratic, 139, KXH. 
22, 24; 293, Ex. 1 

Correspondences, resultant or product 
of two, 05; associative law for, 00; 
commutative, 60 ; groups of, 07; leav- 
ing a figure invariant form a group, 08 



INDEX 



337 



Corresponding elements, 35; doubly, 
102 

Co variant, 257; example of, 258 

Oemona, L., 187, 138 

OroKH ratio, 151); of harmonic set, 150, 
Hit ; definition of, 100 ; expression for, 
100; iu UoiiH Aeneous coordinates, 
105; theorems on, 107, 108, EXH. ; 
characteristic, of projectivity, 205; 
characteristic, of involution, 200 ; as 
an invariant of two quadratic binary 
forms, 251, Ex. 1 ; of four complexes, 
o>2 

Crows ratios, the six, defined by four ele- 
ments, 101 

Curve, of third order, 217, Exs. 7, 8, 0; 
algebraic, 250 

Cyclic e,orrespondence, 00 

Darboux, <}., 05 

Degenerate conies, 120 

Degenerate regulus, 311 

"Decree of geometric problem, 230 

Derivative, 255 

Desargucs, configuration of, 40, 51 ; the- 
orem on perspective triangles, 41, 
180; theorem on conies, 127, 128 

Descartes, R., 285 

Diagonal point (line), of complete quad- 
rangle (quadrilateral), 44; of com- 
plete u-point (n-line) in plane, 44 

Diagonal triangle of quadrangle (quad- 
rilateral), 41 

Dirkson, L. K., 00 

Difference of two points, 148 

Differential operators, 250 

Dimensions, space of three, 20 ; space of 
n, 30 ; assumptions for space of ?i, 33 ; 
space of live, 331 

"Directrices, of a rogulus, 200 ; of a con- 
gruence, 315; of a special complex, 
IJ24 

Distributive law for multiplication with 
respect to addition, 147 

Division of points, 140 

Domain of rationality, 238 

Double element (point, line, plane) of 
correspondence, 08 

Double pairs of a correlation, 97 

Double* points, of a projeetivity on a 
line satisfy a quadratic equation, 150 ; 
of projectivity on a lino, homogeneous 
coordinates of, 104; of projectivity 
always exist in extended space, 242 ; 
of projectivity on a line, construction 
of, 240 ; of involution determined by 
covariant, 258 ; and lines of collinea- 
tion in plane, 271, 205 

Doublo ratio, 150 

Doubly parabolic point, 274 

Duality, in three-space, 28; in plane, 
29 j at a point, 259 j in four-space, 29, 



Ex. ; a consequence of existence of 

correlations, 208 

Edge of %-point or ?i-plane, 36, 37 
Elation, in plane, 72 ; iu space, 75 
Element, undefined, 1; of a figure, 1; 

fundamental, 1; ideal, 7; simple, of 

space, 34; invariant, or double, or 

lixed, 08 ; lineal, 107 
Eleven-point, plane section of, 53, Ex. 15 
Enriques, F., 50, 286 
Equation, of line (point), 174; of conic, 

185, 245; of plane (point), 193, 198; 

reducible, irreducible, 239 , quadratic, 

has roots in extended space, 242 
Equivalent number systems, 150 
Extended space, 242, 255 
Extension, assumptions of, 18, 24 

Face of ?i-point or n-plaue, 36, 37 

format, P., 285 

Field., 140 ; points on a line form a, 151 ; 
finite, modular, 201 ; extended, in 
which any polynomial is reducible, 200 

Figure, 34 

Fine, H. B., 255, 260, 261, 289 

Finite spaces, 201 

Five-point, plane section of, in space, 
39 ; in space may be transformed into 
any other by projective collineation, 
77 ; diagonal points, lines, and planes 
of, in space, 204, Exs. 16, 17, 18; 
simple, in* space determines linear 
congruence, 319 

Five-points, perspective, in four-space, 
54, Ex. 25 

Fixed element of correspondence, 68 

Flat pencil, 55 

Forms, primitive geometric, of one, two, 
and three dimensions, 55 ; one-dimen- 
sional, of second degree, 109; linear 
binary, 251; quadratic binary, 252; 
of nth degree, 254 ; polar forms, 256 ; 
ternary bilinear, represents correla- 
tion in plane, 207 

Four-space, 25, Ex. 4 

Frame of reference, 174 

Fundamental elements, 1 

Fundamental points of a scale, 141, 231 

Fundamental propositions, 1 

Fundamental theorem of projectivity, 
94-97, 213, 264 

General point, 129 

Geometry, object of, 1; starting point 
of, 1 ; distinction between projective 
and metric, 12 ; finite, 201 ; associated 
with a group, 259 

Gergonne, J. D., 29, 123 

Grade, geometric forms of first, second, 
third, 55 

Group, 66 ; of correspondences, 67 ; gen- 
eral projective, on line, 68, 209; 



338 



INDEX 



examples of, 09, 70 ; commutative, 70 ; 
general project! ve, in plane, 268 

HO, assumption, 45 ; r61e of, 81, 261 

Harmonic conjugate, 80 

Harmonic honiology, 223 

Harmonic involutions, 224 

Harmonic set, 80-82 ; exercises on, 83, 
84 ; cross ratio of, 159 

Harmonic transformations, 230 

Harmonically related, 84 

Hesse, 125 

Hessenberg, G., 141 

Hexagon, simple, inscribed in two inter- 
secting lines, 99 ; simple, inscribed in 
three concurrent lines, 250, Ex. 5; 
simple, inscribed in conic, 110, 111 

Hexagram, of Pascal (hexagramma mys~ 
ticum), 138, Exs. 19-21; 304, Ex. 16 

Hilbert, D., 3, 95, 148 

Holgate, T. P., 119, 125, 139 

Homogeneous coordinates in plane, 
174 

Homogeneous coordinates, in space, 11, 
194 ; on line, 163 ; geometrical signifi- 
cance of, 165 

Homogeneous forms, 254 

Homologous elements, 35 

Homplogy, in plane, 72; in space, 75; 
axis and center of, 104; harmonic, 
223, 275 ; canonical lorm of, in plane, 
274, 275 

Hyperosculate, applied to two conies, 136 

Ideal elements, 7 

Ideal points, 8 

Identical correspondence, 65 

Identical matrix, 157, 269 

Identity (correspondence), 65; element 
of group, 67 

Improper elements, 239, 241, 242, 255 

Improper transformation, 242 

Improperly protective, 97 

Independence, of assumptions, 6 ; neces- 
sary for distinction between assump- 
tion and theorem, 7 

Index, of subgroup, 271 ; of group of col- 
lineations in general protective group 
in plane, 271 

Induced correlation in planar field, 262 

Infinity, points, lines, and planes at, 8 

Inscribed and circumscribed triangles, 
98, 250, Ex. 4 

Inscribed figure, in a conic, 118 

Invariant, of two linear binary forms, 
252 ; of quadratic binary forms, 252- 
254, Ex. 1; of binary form of nth 
degree, 257 

Invariant element, 68 

Invariant figure, under a correspond- 
ence, 67 ; of collineation is self -dual, 
272 



Invariant subgroup, 211 

Invariant triangle of collineation, rela- 
tion between project! vities on, 274, 
276, Ex. 5 

Inverse, of a correspondence, 65; of 
element in group, 67; of projectiyity 
is a projectivity, 68; of projectivity, 
analytic expression for, 157 

Inverse operations (subtraction, divi- 
sion), 148, 149 

Involution, 102 ; theorems on, 102, 103, 
124, 127-131, 133, 134, 136, 206, 209, 
221-229, 242-243 ; analytic expression 
for, 157, 222, 254, Ex. 2 ; character- 
istic cross ratio of, 206 ; on conic, 222- 
230 ; belonging to a projectivity, 220 ; 
double points of, in extended space, 
242 ; condition for, 254, Ex. 2 ; dou- 
ble points of, determined by covari- 
ant, 258 ; complexes in, 333 

Involutions, any projectivity is product 
of two, 223; harmonic, 224; pencil 
of, 225; two, have pair in common, 
243; two, on distinct lines are per- 
spective, 243 

Involutoric correspondence, 66 

Irreducible equation, 239 

Isomorphism, 6; between number sys- 
tems, 150 ; simple, 220 

Jackson, D., 282 
Join, 16 

Kantor, S., 250 
Klein, P., 95, 333, 334 

Ladd, C., 138 

Lage, Geometrie der, 14 

Lennes, N. J., 24 

Lindemann, F., 289 

Line, at infinity, 8 ; as undefined class 

of points, 15 ; and plane on the same 

three-space intersect, 22 ; equation of, 

174; and conic, intersection of, 240, 

246 

Line conic, 109 
Line coordinates, in plane, 171 ; in space, 

327, 333 

Lineal element, 107 
Linear binary forms, 251 ; invariant of, 

251 
Linear dependence, of points, 30; of 

lines, 311 

Linear fractional transformation, 152 
Linear net, 84 
Linear operations, 236 
Linear transformations, in plane, 187; 

in space, 199 
Lines, two, in same plane intersect, 

18 
Liiroth, J., 95 



INDEX 



339 



Maclaurin, C., 119 

MacNeish, H. P., 46 

Mathematical science, 2 

Matrices, product of, 156, 268 ; determi- 
nant of product of two, 269 

Matrix, as symbol for configuration, 38 ; 
definition, 150; used to denote pro- 

. jectivity, 15C; identical, 157, 269; 
characteristic equation of, 165, 272; 
conjugate, transposed, adjoint, 269; 
as operator, 270 

MensBchnuw, 126 

Metric geometry, 12 

Midpoint of pair of points, 230, Ex. 6 

Mbbius tetrahedra, 105, Ex. 6; 326, 
Ex. 9 

Multiplication of points, 145, 231 ; the- 
orems on, 145-148 ; commutative law 
of, is equivalent to Assumption P, 
148; other definitions of, 107, Exs. 
3,4 

n-lino, complete or simple, 37, 38; in- 
scribed in conic, 138, Ex. 12 

nrplane, complete in space, 37 ; on point, 
38 ; simple in space, 37 

n-poinl, complete, in space, 36 ; complete, 
in a plane, 37; simple, in space, 37; 
simple, in a plane, 37; plane section of, 
in space, 53, Exs. 13, 16 ; 54, Ex. 18 ; 
m-space section of, in (n + ] )-space, 
54, Ex. 19 , section by three-space of, 
in four-space, 54, Ex. 21; inscribed 
in conic, 119, Ex. 5; 250, Ex. 7 

n-points, in different planes and per- 
spective from a point, 42, Ex. 2 ; in 
same plane and perspective from a 
line, 42, Ex. 4; two complete, in a 
plane, 53, Ex. 7 ; two perspective, in 
(w l)-spaco, theorem on, 54, Ex. 
26 ; mutually inscribed and circum- 
scribed, 250, Ex. 6 

Net of rationality, on line (linear net), 
84; theorems on, 85; in plane, 86; 
theorems on, 87, 88, Exs. 92, 93; in 
space, 89 ; theorems on, 89-92, Exs. 92, 
93 ; in plane (space) left invariant by 
perspective collineation, 93, Exs. 9, 
10; in space is properly projective, 
97; coordinates in, 162 

Newson, II. B., 274 

Nonhomogeneous coordinates, on a line, 
162 ; in plane, 169 ; in space, 190 

Null system, 324 

Number system, 149 

On, 7, 8, 15 

Operation, one-valued, commutative, as- 
sociative, 141 ; geometric, 236 ; linear, 
236 

Operator, differential, 256 ; represented 
by matrix, 270 ; polar, 284 



Opposite sides of complete quadrangle, 
44 

Opposite vertex and side of simple 
7i-point, 37 

Opposite vertices, of complete quadrilat- 
eral, 44 ; of simple ?i-point, 37 

Oppositely placed quadrangles, 50 

Order, 60 

Ordinate, 170 

Origin of coordinates, 169 

Osculate, applied to two conies, 134 

Padoa, A., 3 

Papperitz, E., 309 

Pappus, configuration of, 98, 99, 100, 

126, 148 

Parabolic congruence, 315 

Parabolic point of collineation in plane, 
274 

Parabolic projectivities, any two, are 
conjugate, 209 

Parabolic projectivity, 101; charac- 
teristic cross ratio of, 206; analytic 
expression for, 207 ; characteristic con- 
stants, 207 ; gives H(MA', AA"), 207 

Parametric representation, of points 
(lines) of pencil, 182; of conic, 234; of 
regulus, congruence, complex, 330, 331 

Pascal, B., 36, 99, 111-116, 123, 120, 

127, 138, 139 

Pencil, of points, planes, lines, 55; of 
conies, 129-136, 287-293; of points 
(lines), coordinates of, 181 ; paramet- 
ric representation of, 182 ; base points 
of, 182 ; of involutions, 225 ; of com- 
plexes, 332 

Period of correspondence, 66 

Perspective collineation, in plane, 71 ; 
in space, 75 ; in plane defined when 
center, axis, and one pair of homol- 
ogous points are given, 72 ; leaving R z 
(R*) invariant, 93, Exs. 9, 10 

Perspective conic and pencil of lines 
(points), 215 

Perspective correspondence, 12, 13 ; be- 
tween two planes, 71, 277, Ex. 20 

Perspective figures, from a point or 
from a plane, 35 ; from a line, 36 ; if 
A, JB, C and A', B', C' on two coplanar 
lines are perspective, the points (AB', 
JR4/), (AC', CA^, and (BC f , CB') are 
collinear, 62, Ex. 3 

Perspective geometric forms, 56 

Perspective n-lines, theorems on, 84, 
Exs. 13, 14 ; five-points in four-space, 
54, Ex, 25 

Perspective (n + l)-points in w-space, 
54, Exs. 20, 26 

Perspective tctrahcdra, 43 

Perspective triangles, theorems on, 41, 
53, Exs. 9, 10, 11; 54, Ex. 23; 84, 
Exs. 7, 10, 11; 246; sextuply, 246 



340 



INDEX 



Perspectivity, center of, plane of, axis 
ot, 36 ; notation for, 57 ; central and 
axial, 57, between conic and pencil 
of lines (points), 215 
Fieri, M., 95 

Hanar field, 55 

Planar net, 86 

Plane, at infinity, 8, defined, 17; deter- 
mined uniquely by three noncollinear 
points, or a point and line, or two in- 
tersecting lines, 20 ; and line on same 
three-space are on common point, 
22 ; of perspectivity, 36, 75; of points, 
55; of lines, 55; equation of, 193, 
198 

Plane figure, 34 

Plane section, 34 

Planes, two, on two points -4, B are on 
all points of line AB, 20, two, on 
same three-space are on a common 
line, and conversely, 22 ; three, on a 
three-space and not on a common 
line are on a common point, 23 

Plucker's line coordinates, 327 

Point, at infinity, 8 ; as undefined ele- 
ment, 15 ; and line determine plane, 
17, 20; equation of, 174, 193, 198; of 
contact of a line with a conic, 112 

Point conic, 109 

Point figure, 34 

Points, three, determine plane, 17, 20 

Polar, with respect to triangle, 46; 
equation of, 181, Ex. 3 ; with respect 
to two lines, 52, Exs. 3, 5 ; 84, Exs. 7, 
9 ; with respect to triangle, theorems 
on, 54, Ex. 22 ; 84, Exs. 10, 11 ; with 
respect to n-line, 84, Exs. 13, 14 ; with 
respect to conic, 120-125, 284, 285 

Polar forms, 256 ; with respect to set of 
tt-points, 256; with respect to regu- 
lus, 302 ; with respect to linear com- 
plex, 324 

Polar reciprocal figures, 123 

Polarity, in planar field, 263, 279, 282, 
283 ; in space, 302 j null, 324 

Pole, with respect to triangle, 46 ; with 
respect to two lines, 52, Ex. 3 ; with 
respect to conic, 120 ; with respect to 
regulus, 302; with respect to null 
system, 324 

Poncelet, J. V., 29, 36, 58, 119, 123 

Problem, degree of, 236, 238 ; algebraic, 
transcendental, 288; of second de- 
gree, 245 ; of projectivity, 250, Ex, 14 

Product, of two correspondences, 65; 
of points, 145, 231 

Project, a figure from a point, 36 ; an 
element into, 58, ABC can be pro- 
jected into A'B'C', 59 

Projection, of a figure from a point, 34 

Protective collineation, 71 

Projective comes, 212, 304 



Projective correspondence or transfor- 
mation, 13, 58 ; general group on line, 
08 ; in plane, 268 ; of two- or three- 
dimensional forms, 71, 152 

Projective geometry distinguished from 
metric, 12 

Projective pencils of points on skew 
lines are axially perspective, 64 

Projective pro jectivi ties, 208 

Projective space, 97 

Projectivity, definition and notation for, 
58; ABC-xA'B'C', 59; ABGD^ 
JSAJDO, 60 ; in one-dimensional forms 
is the result of two perspectivities, 63 ; 
if JET (12, 34), then 1234^1243, 82; 
fundamental theorem of, for linear 
net, 94 ; fundamental theorem of, for 
line, 95; assumption of, 95; funda- 
mental theorem of, for plane, 96 ; for 
space, 97 ; principle of, 97 ; necessary 
and sufficient condition for MNAB -r- 
MNA'B' is Q(MAB, NB'A^, 100; 
necessary and sufficient condition 
for MMAB ^ MMA'B' is Q (MAB, 
MB' A'), 101 ; parabolic, 101 ; ABCD 
-K-ABDC implies H(AB, CD), 103; 
nonhomogeneous analytic expression 
for, 154-157, 206 ; homogeneous ana- 
lytic expression for, 164; analytic 
expression for, between points of dif- 
ferent lines, 167 ; analytic expression 
for, between pencils in plane, 183; 
between two conies, 212-216; on 
conic, 217-221; axis (center) of, on 
conic, 218; involution belonging to, 
226 ; problem of, 250, Ex. 14. 

Projectivities, commutative, example of, 
70 ; on sides of invariant triangle of 
collineation, 274, 276, Ex. 5 

Projector, 35 

Properly projective, 97; spatial net is, 97 

Quadrangle, complete, 44; quadrangle- 
quadrilateral configuration, 46; sim- 
ple, theorem on, 52, Ex. 6 ; complete, 
and quadrilateral, theorem on, 53, 
Ex. 8; any complete, may be trans- 
formed into any other by projective 
collineation, 74; opposite sides of, 
meet line in pairs of an involution, 
103 ; conies through vertices of, meet 
line 111 pairs of an involution, 127 ; 
inscribed in conic, 137, Ex. 11 

Quadrangles, if two, correspond so that 
five pairs of homologous sides meet 
on a line Z, ,tho sixth pair meets on 
, 47, perspective, theorem on, 68, 
Ex. 12 ; if two, have same diagonal 
triangle, their eight vertices are on 
conic, 137, Ex. 4 

Quadrangular set, 49, 79 ; of lines, 79 ; of 
planes, 79 



INDEX 



341 



Quadrangular section by transversal of 
quadrangular set of lines is a quad- 
rangular set of points, 79 ; of elements 
protective with quadrangular set is 
a quadrangular set, 80; Q(MAB, 
NB'A')is the condition for MNAB-^ 
MNATK, 100 ; Q(MAB, MR' A') is the 
condition for MM A B-j- M, M A'B', 101 ; 
Q(ABC, A'B'V) implies Q(A'B'C', 
ABC), 101 ; Q(ABC, A'B'C') is the 
condition that A A', BB', Ware in in- 
volution, 103 ; Q(P Pa-Po, P, PyPx+v) 
is necessary and sufficient for P^ + P 
= P a4 . y , 142; QCP.PoPi, P P,rP,v) 
is necessary and sufficient for P x P y 
= P^145 

Quadrangularly related, 80 

Quadratic binary form, 252; invariant 
of, 252 

Quadratic correspondence, 139, Exs. 
22, 24 

Quadric spread in S a , 331 

Quadric surface, 301 ; degenerate, 308 ; 
determined by nine points, 311 

Quadrilateral, complete, 44 ; if two quad- 
rilaterals correspond so that five of the 
lines joining pairs of homologous ver- 
tices pass through a point P, the line 
joining the sixth pair of vertices will 
also pass through P, 40 

Quantic, 254 

Quaternary forms, 258 

Quotient of points, 149 

Range, of points, 55 ; of conies, 128-136 

Ratio, of points, 149 

Rational operations, 149 

Rational space, 98 

Rationality, net of, on line, 84, 85 ; planar 
net of, 80-88 ; spatial net of, 89-93 ; 
domain of, 238 

Rationally related, 80, 89 

Reducible equation, 239 

Reflection, point-line, projective, 223 

Reflexive correspondence, 60 

Regulus, determined by three lines, 298 ; 
directrices of, 299 ; generators or 
rulers of, 299; conjugate, 299; gen- 
erated by projective ranges or axial 
pencils, 299 ; generated by projective 
conies, 304, 307 ; polar system of, 300 ; 
picture of, 300 ; degenerate cases, 311 ; 
of a congruence, 318 

Related figures, 35 

Resultant, of two correspondences, 65 ; 
equal, 65; of two projectivities is a 
projectivity, 68 

Reye, T., 125, 139 

Rohn, K, 309 

Salmon, G., 138 
Sannia, A., 304 



Scale, defined by three points, 141, 231 ; 
on a conic, 231 

Schroter, H., 138, 281 

Schur, F., 95 

Science, abstract mathematical, 2 ; con- 
crete application or representation of, 2 

Scott, C. A., 203 

Section, of figure by plane, 34 ; of plane 
figure by line, 35 ; conic section, 109 

Segre, C., 230 

Self -conjugate subgroup, 211 

Self -conjugate triangle with respect to 
conic, 123 

Self-polar triangle with respect to conic, 
123 

Set, synonymous with class, 2 ; quadran- 
gular, 49, 79; of elements projective 
with quadrangular set is quadrangu- 
lar, 80*; harmonic, 80; theorems on 
harmonic sets, 81 

Seven-point, plane section of, 53, Ex. 14 

Seydewitz, F., 281 

Sheaf of planes, 55 

Side, of ft-point, 37 ; false, of complete 
quadrangle, 44 

Similarly placed quadrangles, 60 

Simple element of space, 39 

Simple n-point, n-line, n-plane, 37 

Singly parabolic point, 274 

Singular point and line in nonhomoge- 
neous coordinates, 171 

Six-point, plane section of, 54, Ex, 17; 
in four-space section by three-space, 
54, Ex. 24 

Skew lines, 24; projective pencils on, 
are perspective, 105, Ex. 2 ; four, are 
met by two lines, 250, Ex. 13 

Space, analytic projective, 11 ; of three 
dimensions, 20; theorem of duality 
for, of three dimensions, 28 ; n-, 30 ; 
assumption for, of n dimensions, 33 ; 
as equivalent of three-space, 34; 
properly or improperly projective, 
97; rational, 98; finite, 201, 202; 
extended, 242 

Spatial net, 89, theorems on, 89-92; 
is properly projective, 97 

von Staudt, K. G. C., 14, 95, 125, 141, 
151, 158, 100, 286 

Steiner, J., 109, 111, 125, 138, 139, 285, 
280 

Steiner point and line, 138, Ex. 19 

Steinitz, E., 261 

Sturm, Ch., 129 

Sturm, R., 231, 250, 287 

Subclass, 2 

Subgroup, 68 

Subtraction of points, 148 

Sum of two points, 141, 231 

Surface, algebraic, 259; quadric, 301 

Sylvester, J. J., 323 

System affected by a correspondence, 65 



342 



INDEX 



Tangent, to conic, 112 

Tangents to a point conic form a line 
conic, 116 ; analytic proof, 187 

Taylor's theorem, 255 

Ternary forms, 258; bilinear, repre- 
sent correlation in a plane, 267 

Tetrahedra, perspective, 43, 44 ; config- 
uration of perspective, as section of 
six-point in four-space, 54, Ex. 24: 
Mbbius, 105, Ex. ; 326, Ex. 9 

Tetrahedron, 37; four planes joining 
line to vertices of, projective with 
four points of intersection of line 
with faces, 71, Ex, 5 

Three-space, 20; determined uniquely 
by four points, by a plane and a point, 
by two nonintersecting lines, 23 ; the- 
orem of duality for, 28 

Throw, definition of, 60 ; algebra of, 141, 
157; characteristic, of project! vity, 
205 

Throws, two, sum and product of, 158 

Trace, 35 

Transform, of one projectivity by an- 
other, 208 ; of a group, 209 

Transform, to, 58 

Transformation, perspective, 18; pro- 
jective, 13 ; of one-dimensional forms, 
58; of two- and three-dimensional 
forms, 71 

Transitive group, 70, 212, Ex. 6 

Triangle, 37; diagonal, of quadrangle 
(quadrilateral), 44; whose sides pass- 
through three given collinear points 
and whose vertices are on three given 
lines, 102, Ex. 2; of reference of 
system of homogeneous coordinates 



in plane, 174; invariant, of collinea- 
tion, relation between projectivities 
on sides of, 274, 270, Ex. 5 

Triangles, perspective, from point arc 
perspective from line, 41 ; axes of 
perspectivity of three, in plane per- 
spective from same point, are con- 
current, 42, Ex, G; perspective, theo- 
rems on, 53, Exs. 9, 10, 11 ; 105, Ex. 
9 ; 116, 247 ; mutually inscribed and 
circumscribed, 99; perspective, from 
two centers, 100, Exs. 1, 2, 3; from 
four centers, 105, Ex. 8 ; 138, Ex. 18 ; 
from six centers, 246-248 ; inscribed 
and circumscribed, 250, Ex. 4 

Triple, point, of lines of a quadrangle, 
49 ; of points of a quadrangular set, 49 

Triple, triangle, of lines of a quadran- 
gle, 49; of points of a quadrangular 
set, 49 

Triple system, 3 

Undefined elements in geometry, 1 

United position, 15 

Unproved propositions in geometry, 1 

Variable, 58, 150 

Veblen, 0., 202 

Veronese, G., 52, 53 

Vertex, of w-points, 36, 37; of n-planes, 

37; of fiat pencil, 55; of cone, 109; 

false, of complete quadrangle, 44 

Wiener, H., 65, 95, 230 
Zeuthen, H. G., 95 



Date Due 



w 




Demco 293-5 



. S?. r K9!5.M5!!" University Libraries 



3 fiHfiS 01111 "Git 



V39r- 
V.1 (2) 



Carnegie Institute of Technology 

Library 
Pittsburgh, Pa.