Skip to main content

Full text of "Solid Analytical Geometry And Determinants"

See other formats


TEXT FLY 
WITHIN THE 
BOOK ONLY 



00 

64265 >m 

co 

> 



SOLID ANALYTICAL GEOMETRY 
AND DETERMINANTS 



BY THE SAME AUTHOR 



Plane Trigonometry 

This book emphasizes the importance of the 
function concept for elementary trigonometry. 
Cloth; 6 by 9 inches; 110 pages; 58 figures. 

PUBLISHED BY 

JOHN WILEY & SONS, INC. 
NEW YORK 



SOLID 'ANALYTICAL GEOMETRY 
AND DETERMINANTS 



Professor of Mathematics , Swarthmore College 



NEW YORK 

JOHN WILEY & SONS, INC. 

LONDON: CHAPMAN & HALL, LIMITED 

1930 



COPYRIGHT, 1930, 
BY ARNOLD DRESDEN 



Printed in U. S. A. 

Printing Composition and Plates Binding 

F. H. CELSON CO. TECHNICAL COMPOSITION CO. STANHOPE BINDEKY 

BOSTON CAMBRIDGE BOSTON 



PREFACE 

A three-hour course in Solid Analytical Geometry is offered for 
students in the junior year in many of the colleges and universities 
in this country. Though books ,on Plane Analytical Geometry 
frequently devote some chapters ^to the geometry of a space of 
three dimensions, the material covered in these chapters is, with 
few exceptions, not intended to do more than v prt)vide a general 
introduction to the subject, so as to enable students tb understand 
the references to it which have to be made in courses on the cll- 
culus. But it rarely goes far enough to acquaint them with the, 
more interesting and valuable methods of this field. 

For many years, while teaching this subject at the University 
of Wisconsin and at Swarthmore College, it has seemed to the 
author that, in the study of Solid Analytical Geometry, the young 
student of mathematics can find an excellent opportunity for an 
introduction to methods and principles which have an important 
part in various fields of advanced mathematics. Among these are 
the methods based on the theory of determinants and on the con- 
cept of the rank of a matrix. In more advanced mathematical 
subjects these theories are developed and used with a great meas- 
ure of generality; they find relatively simple application in the 
subject to which this book is devoted. Unfortunately, there are 
not readily accessible for use in undergraduate classes treatments 
of these theories which are on the one hand adequate for the uses 
to be made of them here and on the other hand not too advanced 
to be available for an introductory purpose. 

For these reasons, the first chapter of this book presents an 
exposition of some of the properties of determinants and matrices, 
followed in Chapter II by a treatment of systems of linear Aqua- 
tions. The latter subject is not carried so far as to include the 
most general case, but, it is hoped, far enough to serve in the later 
chapters. The repeated applications of the results of the first 
two chapters which are made in the subsequent work (as evidenced 
by the numerous references to Chapters I and II) should indicate 
their importance. With the basis thus provided it becomes pos- 

iii 



IV PREFACE 

sible to deal with the geometrical questions of the later chapters 
in a way which lends itself readily to extension to problems of a 
more general character. Thus the discussion of the theory of 
quadric surfaces in Chapters VII and VIII may be made to serve 
as an introduction to the theory of quadratic forms in n variables. 
Having studied these chapters, the reader should be able to proceed 
to the well-known excellent books, by Bocher and by Dickson, 
nientioned in the introductory paragraph of Chapter I. To these 
books the author owes a large debt. The spirit which pervades 
them has been a guide for him; and it would be a source of grati- 
fic,tion if the present book were to lead its readers to more extended 
sftidy of the subjects treated by these authors. 

Chapters III to X deal with the loci of equations of the first 
and second degree in three variables from the point of view of real, 
metric geometry. Elements at infinity and complex elements 
are considered as non-existent. This point of view has been 
taken because, in the author's judgment, a satisfactory treatment 
of the questions which arise through the inclusion of such elements 
can only be made after the explicit adoption of adequate bases on 
which projective geometry and complex geometry can be erected. 
Since this would involve quite a different orientation than the 
scope of the present book permits, it was deemed better to proceed 
on the implicit assumptions of real metric geometry on which the 
student's earlier work in geometry may be supposed to have been 
founded. 

It is only from this point of view that the detailed classification 
of quadric surfaces made in Chapter VIII can be justified; and 
only when it is kept in mind, can statements like the one in The- 
orem 12 on page 175, to the effect that there are no lines on a 
non-singular quadric with negative discriminant, be explained. 
It must however be recognized that there occur instances in which,, 
if only for the sake of emphasis, infinite elements and complex 
elements must be mentioned. 

In subject matter the last eight chapters follow largely the tra- 
ditional content of introductory courses in Solid Analytical Geom- 
etry. The treatment introduces modes of procedure and devices 
which have been developed in the course of the many years during 
which^the author has taught the subject and which have probably 
also been used by other teachers. 



PREFACE 

It will usually be most satisfactory to work through the greater 
part of Chapters I and II before Chapter III is started. But it 
may be found desirable, as has frequently been the author's prac- 
tice, to begin by spending one hour a week on the geometric part 
of the book, beginning with Chapter III, while the remainder of 
the time is given to the algebraic work of the first two chapters. 

The exercises form an integral part of the course which this 
book presents. The author has not hesitated therefore to refer, 
in a number of cases (see e.g., pages 116, 222, and 228), to results 
established in an exercise. A good many of the problems serve 
no other purpose than that of illustrating the material in the text. 
But there are other problems, and these are doubtless the nA>re 
valuable ones, which require a certain amount of original thinking. 

Thanks are due to other authors besides those which were men- 
tioned in a preceding paragraph; but the uses which I have made 
of their work are too indefinite in character to make explicit 
references possible. It is a pleasure to acknowledge my indebt- 
edness to the mathematical library of Brown University for allow- 
ing photographs to be made of the models in its possession for 
the illustrations of quadric surfaces which appear in this book. 
Furthermore I wish to express my thanks to Mr. George B. 
Hoaclley, a senior student in Swarthmore College, who has drawn 
the figures, and to Miss Alice M. Rogers, research assistant at 
the Sproul Observatory of Swarthmore College, who has given 
valuable aid in the reading of proofs. This preface would not be 
complete without a word of appreciation for the unfailing courtesy 
and patience which the publishers and the printers have contrib- 
uted to the production of this book. 



ARNOLD DRESDEN. 



SWAllTHMORE COLLEGE, 

February, 1930. 



CONTENTS 

CHAPTER I 

DETERMINANTS AND MATRICES 
SECTION 

1. Introduction 1 

2. Definitions and notations 1 

3 The value of a determinant 3 

4. Exlrcises 

5. Elementary theorems ' { 

6. Exercises 10 

T. Minors and cof actors 11 

8. Exercises 15 

0. Matrices. Rank of a matrix 16 

10. Complementary minors. Elementary transformation of matrices . 17 

11. Exercises 19 

12. The Laplace development of a determinant 20 

13. Exercises 24 

14. The product of two determinants 25 

15. Exercises 28 

16. The adjoint of a determinant 28 

17. The derivative of a determinant 31 

18. Exercises 32 

19. Miscellaneous exercises 33 



CHAPTER II 
LINEAR EQUATIONS 

20. Definition and notation 35 

21. The system of n linear non-homogeneous equations in n variables . . 36 

22. The system of n linear homogeneous equations in n variables .... 38 

23. The system of n + 1 linear non-homogeneous equations in n i 

variables 38 

24. Exercises 40 

25. The system of n 1 linear homogeneous equations in n variables . . 41 

26. The adjoint of a vanishing determinant. Symmetric determi- 

nants 42 

27. The system of n linear non-homogeneous equations in n variables, 

continued 44 

28. Exercises 47 

vii 



Vlll CONTENTS 

CHAPTER III 

POINTS AND LINES 

SECTION PAGE 

29. The cartesian coordinates of a point in three-space 49 

30. The coordinate parallelepiped of a point 51 

81. Exercises 52 

32. Two points 53 

33. Direction cosines of a line 55 

J34. Three collinear points 57 

35. Exercises 60 

36. The angle between two lines. The projection method 61 

37. Exercises 65 

38. Miscellaneous exercises 65 

CHAPTER IV 
PLANES AND LINES 

39. Surfaces and curves 67 

40. Cylindrical surfaces. Systems of planes 68 

41. The linear equation ax + by -\- cz -\- d = 71 

42. Exercises 74 

43. The distance from a plane to a point 75 

44. The normal form of the equation of a plane 77 

45. Exercises 80 

46. Two planes 81 

47. The line 83 

48. Exercises 89 

49. The pencil of planes. The bundle of planes 90 

50. Exercises 94 

51. Three planes. A plane and a line 95 

52. The plane and the line, continued 97 

53. Exercises 100 

54. Four planes. Two lines 101 

55. Exercises 105 

56. Miscellaneous exercises 105 

CHAPTER V 

OTHER COORDINATE SYSTEMS 

57. Spherical coordinates 108 

58. Cylindrical coordinates Ill 

59* Exercises 112 

60. Oblique cartesian coordinates 113 

61. Translation of axes 114 

62. Transformation from oblique to rectangular axes 115 

63. Rotation of axes 118 



CONTENTS ix 

SECTION PAGE 

64. Exercises 121 

65. Rotation of axes, continued 122 

66. Linear transformation. Plane sections of a surface 125 

67. Exercises 129 

CHAPTER VI 
GENERAL PROPERTIES OF SURFACES AND CURVES 

68. Surfaces of revolution 131 

69. Exercises 135 

70. The shape of a surface determined from its equation. Contour 

lines 1 35 

71. Some facts from Plane Analytical Geometry 140 

72. Some special surfaces 141 

73. Exercises 148 

74. The intersections of a surface and a line 149 

75. Digression on Taylor's theorem 151 

76. The intersections of a surface and a line, continued 152 

77. Tangent lines and tangent planes. Normals 154 

78. Exercises 156 

79. The shape of a curve in space ^-r*, 157 

CHAPTER VII 
QUADRIC SURFACES, GENERAL PROPERTIES 

80. The quadric surface and the line , 159 

81. Tangent line; tangent plane; normal; polar plane. 161 

82. Polar plane and pole. Tangent cone 165 

83. Exercises 169 

84. Ruled quadric surfaces ( 170 

85. The centers and vertices of quadric surfaces 176 

86. Exercises .-.:,;, ' 182 

87. The asymptotic cone "T'^* wil ^r? [ 182 

88. The diametral planes and the principal planes of a Quadric surface . 185 

89. The discriminating equation . . 190 

90. Principal planes and principal directions . . 194 

91. Exercises 195 

CHAPTER VIII 
CLASSIFICATION OF QUADRIC SURFACES 

92. Invariants 197 

93. Invariants of a quadric surface with respect to rotation and trans- 

lation of axes 199 

94. Invariance of the discriminant of a quadric surface with resect to 9 

rotation 202 



X CONTENTS 

SECTION PAGE 

95* Exercises 205 

96. Two planes 206 

97. Translation of axes to the center of a quadric surface 210 

98. Rotation of axes to the principal directions of a quadric surface.. . 211 

99. Classification of quadric surfaces the non-singular cases 214 

100. Classification of quadric surfaces the non-degenerate singular 

cases 220 

101. Classification of quadric surfaces the degenerate cases 227 

102. The classification of quadric surfaces summary and geometric 

characterization 229 

103. Exercises 234 



CHAPTER IX 
QUADRIC SURFACES, SPECIAL PROPERTIES AND METHODS 

104. The reguli on the hyperboloid of one sheet 235 

105. Reguli on the hyperboloid of one sheet, continued 237 

106. The reguli on the hyperbolic paraboloid 242 

107. The straight lines on the singular, non-degenerate quadrics 243 

108. Exercises 243 

109. Circles on quadric surfaces, the general method 244 

1 10. Circles on quadric surfaces, continued 251 

111. Exercises 252 

112. Tangent planes parallel to a given plane. The umbilics of a quad- 

ric surface 253 

113. The umbilics of a quadric surface, continued 263 

114. Exercises 266 

CHAPTER X 
PROPERTIES OP CENTRAL QUADRIC SURFACES 

115. Conjugate diameters and conjugate diametral planes of central 

quadrics. Enveloping cylinder 268 

116. Exercises 271 

117. Conjugate diameters of the ellipsoid 271 

118. Exercises 276 

119. Linear families of quadrics 277 

120. Focal curves and directrix cylinders of central quadrics 281 

121. Focal conies and directrix cylinders, continued 285 

122. Exercises 287 

123. Confocal quadric surfaces. Elliptic coordinates 288 

Appendix 296 

Index 303 



SOLID ANALYTICAL GEOMETEY 
AND DETERMINANTS 



CHAPTER I 
DETERMINANTS AND MATRICES 

1. Introduction. The Study of Solid Analytical Geometry, to 
which this book is chiefly devoted, leads repeatedly to the problem 
of solving systems of linear equations in several variables, in which 
the number of variables may be less than, equal to, or greater than 
the number of equations. The methods for dealing with this 
problem which are found in books on elementary algebra and in 
college algebra texts are not sufficiently general in character to 
suit the needs of our subject. A more complete treatment of the 
theory of determinants than is found in such books becomes neces- 
sary. For this reason, and also on account of the manifold uses 
of determinants in various fields of mathematics, finally because 
of the great intrinsic interest of the subject, the first chapter of 
this book will be devoted to an introduction to the theory of de- 
terminants and to a few ideas concerning matrices. This will be 
followed in Chapter II by a treatment of systems of linear equa- 
tions. In this treatment the problem is not considered in its 
complete generality, but in a form sufficiently inclusive to suit the 
needs of the later chapters in this book. The reader who desires 
to pursue this subject further can do so in two excellent books, 
dealing with advanced topics in algebra, viz., B6cher, Introduction 
to Higher Algebra, and Dickson, Modern Algebraic Theories. 

2. Definitions and Notations. 

DEFINITION I. A determinant is a square array of numbers to which 
a single number, called the value of the determinant, is attached by 
the method stated in Definition V. 

Vertical bars are placed on either side of the array. The symbol 
so obtained is used also to designate the number that is to be asso- 

1 



DETERMINANTS AND MATRICES 



elated with the array. For example, the symbols 

4-12 



and 







3-5 
1 4 



with each of which a single number is associated in accordance 
with Definition V are determinants. The same symbols are used 
to designate the values of these determinants. 

DEFINITION II. The numbers In the square array which constitutes 
the determinant are called its elements; the horizontal lines in the 
array are called rows, the vertical lines columns; the diagonal which 
runs from upper left to lower right is called the principal diagonal, 
the other diagonal Is called the secondary diagonal. 

DEFINITION III. The order of a determinant is the number of ele- 
ments in any one row or column. 

Remark. A determinant of order n is made up of n 2 elements. 

Notations, In the general form of a determinant every ele- 
ment has affixed to it two indices; the first of these designates 
the row in which the element stands and is called the row index, 
the second designates the column of the element and is called the 
column index. The general forms of the determinants of the third 
and fourth order will therefore be as follows : 



#31 032 



(2) 



#11 #12 #13 #14 

#21 #22 #23 #24 

#31 #32 #33 #34 

#41 #42 &43 #44 

A determinant of the nth order, where n designates a positive 
integer, in the most general form will appear as follows: 

#11 #12 
/Q\ #21 #22 #2 W 



# n l 



# 



These are rather lengthy symbols; they make it desirable to 
have more compact symbols which can be used when it is not 
necessary to designate the elements of the determinant explicitly. 
In such cases we frequently designate a determinant by merely 
writing the elements of the principal diagonal. Thus the 
symbols |#ii#22#33|, |#n#22#33#44| and \ana^ . . . a nn \ are used 



THE VALUE OF A DETERMINANT 3 

to designate the cleterminants (1), (2), and (3) respectively. A 
still shorter way of representing the general determinant consists 
in writing a single element with literal indices and indicating the 
values which these indices a*e to take. In this notation the deter- 
minants (1), (2), and (3) t&mld be represented by the notations 
K>'l, i,i = 1, 2, 3; |oy|, i,j = 1, 2, 3, 4; and |a</|, i, j = 1, 2, 
. . . , n respectively. 
3. The Value of a Determinant. 

DEFINITION IV. Whenever in a set of numbers, consisting: of Inte- 
gers from 1 upward in arbitrary order, a larger integer precedes a 
smaller one, we say that there is an inversion. 

For example, in the row of indices 

463251 

there are 11 inversions: 3 inversions because the number 4 is 
followed by the smaller numbers 3, 2 and 1 ; 4 inversions because 
6 is followed by 3, 2, 5 and 1 ; 2 inversions because 3 precedes 2 
and 1; 1 inversion because 2 precedes 1, and 1 inversion because 
5 precedes 1. 

DEFINITION V. The value of a determinant is the algebraic sum of 
all possible products obtainable by taking one and only one factor 
from each row and from each column, preceded by the plus or minus 
signs, according as the number of inversions of the column indices 
of the factors of a product are even or odd, when the row indices are 
in the natural order 1, 2, 3, etc. 

The indicated sum of these products is called the expansion 
of the determinant. 

Remark 1. We must remember that, although the same sym- 
bol is used for a determinant and for the value of this determinant, 
the concepts " determinant " and "value of a determinant" are 
distinct concepts; the latter is a number; the former is a square 
array of numbers with which a number is associated according to 
Definition V. 

Remark 2. The expansion of the general determinant of the 
nth order (3) will therefore consist of terms of the form a^o^ . . . 
anc n , in which Ci c 2 . . . c* is some permutation of the set of num- 
bers 1,2,. . . n, this term will be preceded by the plus or minus 
sign, according as the number of inversions of the set Ci C2 . . . c n 
is even or odd. Since the number of permutations of the set of 



DETERMINANTS AND MATRICES 



integers 1,2,. . . n is nl, it follows that the expansion of the gen- 
eral determinant (3) consists of n\ terms. 
Examples. 



1. The value of the second order determinant 



011 

021 



012 



is the algebraic 



sum of two terms; each term must contain two factors, one and only one 
from each row and from each column. If we take a\\ from the first row, we 
must take 022 from the second; thus we get the product 011022. If we take 012 
from the first row, we must take a>2i from the second, so that we obtain the 
product 012021. In both these products the row indices are in the natural 
order 1,2. In the first product the set of column indices is 1,2, which has no 
inversions; the column indices in the second product form the set 2,1, which 
has 1 inversion. Consequently the product n 022 is preceded by the plus 
sign, and the product 012021 is preceded by the minus sign. Therefore 



(in 



a u 

022 



012021- 



2. The value of the third order determinant (1) is obtained as the alge- 
braic sum of 6 products; if we write the factors of each product in the order 
of their row indices these products arc 011022033, 0,110290,32, 012021033, 012023031, 
013021032, aisaajsasi. The numbers of inversions in the column indices of these 
products are 0,1,1,2,2 and 3 respectively. We conclude that 

011023032 ~ 012021033 + 0120*3031 



011022033 I = 01102233 



3. To determine the value of the determinant 



013021^32 ~ 013022031. 



1 
3 

G 
10 



1 

4 

10 

20 



we write 



down every possible product of four factors, in each of which there is one and 
only one element from each row and from each column. We order the factors 
in each product according to the rows from which they are taken, and indicate 
below them the columns to which they belong. The number of inversions 
in the column indices then determines the sign to be prefixed to each product, 
in accordance with the rule laid down in Definition V. Thus we obtain the 
following expansion: 



-4- 1 


2- 


6-20-1 


2 


10- 


10 


- 1 


3- 


3 20 + 1 3 10 4 + 1 - 


4- 


3- 10 


i 


2 


3 4 


1 


2 


4 


3 


1 


3 


24 1342 1 


4 


2 -3 


1 


. 4 . 


6-4 - 


1 


1 


6- 


20 


+ 1 


1 


10 10 + 1 3 1 20 - 1 





10-1 


i 


4 


3 2 


2 


1 


3 


4 


2 


1 


43 2314 2 


3 


4 1 


- 1 


. 4 . 


1 10 - 


f 


1 - 


4 - 6 


- 1 


-f 1 


1 


- 3 20 - 1 I 10 - 4 - 1 - 


2- 


1-20 


2 


4 


1 3 




2 


4 3 


1 


* 3 


1 


24 3142 3 


2 


1 4 


-f 1 


2 


10-1 


4- 


1 


4 J 




t 1 


4 


3 1 - 1 1 3 - 10 + 1 


1 


6-4 


3 


2 


4 1 




3 


4 ] 


I 2 3 


4 


21 4123 4 


1 


3 2 


-f 1 


-2- 


1 10 - 


1 


2 


6- 1 





1-3 


1 - 


4 + 1 3 3 1. It follows 


that the 


4 


2 


1 3 


4 


2 


3 1 




4 3 


1 


2 4321 







* The symbol nl, called "n factorial " is an abbreviation for the continued 
product 1 2 3 . . . (n 1) n. 



ELEMENTARY THEOREMS 



value of the given* determinant is equal to 240 - 200 - 180 + 120 + 120 

- 96 - 120 -f 100 + 60 - 30 - 40 + 24 + 60 - 40 - 40 + 20 + 16 - 12 

- 30 + 24 + 20 - 12 - 12 + 9 = 813 - 812 = + 1. 

4. Exercises. 

1. Determine the number of inversions in each of the following sequences of 
numbers: 

(a) 5 2 4 7 3 I G 
(6) 3 6 1 5 4 7 2 
(c) 7645321 

2. How many terms are there in the expansion of a determinant of the 4th 
order? Of the 5th order? Of the 6th order? 

3. Prove that the number of inversions in a row of numbers is not changed 
if all the numbers are increased or decreased by the same amount. 

4. Show that if a row of integers is divided into two sections, such that 
all the numbers in the left section are less than any number in the right section, 
then the number of inversions in the original row is equal to the sum of the 
number of inversions in the left part and that in the right part. 

6. Generalize the theorem of the preceding exercise so as to cover the case 
in which a row of integers is divided into more than two sections. 

6. Determine the values of each of the following determinants by the method 
explained and illustrated in Section 3: 



(a) 



4-1 2 

3 -5 

-2 1 4 



(6) 



-4 
3 



-3 2 
1 -2 
5 -2 



4 -6 3 

-6 1 2 

3 2 5 



7. Determine also the values of the following determinants: 



(a) 



1121 




2-1 3-1 




2 2 2 10 


23-4 
3210 


; (6) 


4 -2 -1 3 
2-1-4 4 


; (c) 


0012 
34-32 


1-11 1 




10 -5 -6 10 




1-245 




23 1-1 




a 1 1 1 






2003 




1 a 1 1 




(d) 


4101 


) \") 


1 1 a 1 






-12-2 1 




Ilia 





6. Elementary Theorems. The determination of the value of 
a determinant by means of Definition V is quite laborious even 
for a determinant of order 4, as will, have been discovered by the 
reader who has done all the parts of Exercise 7. For determinants 
whose order exceeds 4, this method becomes quite useless. Never- 
theless it is important for the reader to do the exercises in the pre- 
ceding set so that he may become thoroughly familiar with the 
content of Definition V. In the next few sections we shall deriVe 



6 DETERMINANTS AND MATRICES 

from this definition a chain of theorems which will supply us with 
the more useful methods for evaluating a determinant which will 
be employed in our further work. 

THEOREM 1. The Interchange of two adjacent numbers in a row of 
integers which is an arrangement of the integers from 1 to n either 
increases or decreases the number of inversions of the row by one. 

Proof. Let p and q be adjacent. We have then to compare the 
number of inversions of the sets 

(1) ... pq . . . and (2) ... qp . . . 

Let us suppose p < q. The inversions which arise in (1) from 
any number preceding p or following q will also occur in (2), for 
all such numbers are followed by the same numbers in (2) as in (1). 
Furthermore p is followed in (2) by the same numbers which 
follow it in (1), except q] but since q > p this change does not 
affect the number of inversions, so that the number of inversions 
due to p is the same in (2) as it was in (1). Finally q is followed 
in (2) by the same numbers as in (1) and moreover by p, which is 
less than g, so that the number of inversions due to q is one more in 
(2) than it was in (1). We conclude that there is a gain of one 
inversion in passing from (1) to (2), and consequently a loss of one 
inversion in passing from (2) to (1). Since in (2) the larger one 
of the two indices that are interchanged precedes the smaller one, 
whereas in (1) the condition is the opposite one, the proof of the 
theorem is complete. 

THEOREM 2. The interchange of any two numbers in a row of in- 
tegers which is an arrangement of the integers from 1 to n changes 
the number of inversions in the row by an odd number. 

Proof. Let there be k numbers between p and q and let us 
compare the two arrangements 

k k 

(3) . . . p ... q . . . and (4) . . . q ... p . . . . 

By interchanging p successively with each of the k numbers 
which lie between p and q in (3), we obtain an arrangement which 
may be represented by 

k 



ELEMENTARY THEOREMS 7 

interchanging now q with p and with the k numbers preceding it, 
we obtain the arrangement (4). Thus we obtain (4) from (3) 
as the last of 2 k + 1 arrangements, each one of which is ob- 
tained from the preceding one by the interchange of two adjacent 
numbers. Hence the number of inversions in, (4) is obtained, in 
virtue of Theorem 1, from that of (3) by 2 k + 1 changes of one, 
some of which are losses and the others gains. Since 2 k + 1 
is odd, no matter what integer k is, the net result will be the loss 
or the gain of an odd number of inversions. For, if there are I 
losses and g gains, and I > g, the net result will be I g = li losses; 
but / + g = 2k + 1 and therefore 2Z = Zi + 2fc + l, from which 
we conclude that Zi is odd. And if the number of gains exceeds 
the number of losses it is shown in exactly similar fashion that 
the net result consists of an odd number of gains. 

Remark. By allowing the change in the number of inversions 
to take negative as well as positive values we can say that every 
interchange of two numbers in a row of integers which is an ar- 
rangement of the integers from 1 to n changes the number of in- 
versions by an odd number. 

THEOREM 3. If an arrangement of the integers 1 . . . n can be 
obtained from the natural order, or can be restored to the natural 
order, by an even (odd) number of interchanges of a pah* of numbers, 
it will have an even (odd) number of inversions. 

Proof. The natural order presents no inversions and every inter- 
change of a pair of numbers changes the number of inversions by 
an odd number. The truth of the theorem follows therefore 
from the fact that the sum of an even number of odd numbers is 
even, while the sum of an odd number of odd numbers is odd. 

THEOREM 4. The value of a determinant is not changed if the 
columns are made into rows and the rows into columns. 

Proof. Let the given determinant be |a#|, i, j = 1, 2, . . . , 77; 
and let the determinant obtained by making the rows into col- 
umns and the columns into rows be designated by |6#|, i,j = 1, 2, 
. . . , n. Then &# = a#. An arbitrary term in the development 
of |o#| has the form a^o^ . . . a^ n in which 

(5) Ci C2 . . . Cn 

is some arrangement of the integers 1, 2, . . . , n; and the si^n 
of this term depends on the number of inversions in the row of 



8 DETERMINANTS AND MATRICES 

numbers (5) (see Remark 2 on page 3). Moreover, this term is 

cqualto (6) Ms- &% 

which is a term in the development of the determinant |&#|, 
except possibly for sign. In order to determine the sign of (6) 
we have to rearrange its factors so as to put the row indices in 
natural order; in doing this we shall put the column indices in 
irregular order and the sign of (6) will depend upon the number 
of inversions in that order. Now this order is obtained from the 
natural order by as many interchanges of pairs of numbers as it 
takes to restore the arrangement (5) to the natural order; conse- 
quently it will present an even or odd number of inversions ac- 
cording as the number of inversions in (5) is even or odd. Con- 
sequently the term (6) will appear in |6#| with the same sign that 
the term a^c^ . . . a nCn had in |a#[. But this last term was an 
arbitrary term in |o#|; therefore every term of the development 
of \0ij\ occurs in the development of |6#| and with the same sign. 
The same argument shows that the terms of |6#| are all reproduced, 
in magnitude and in sign, in the development of |ay|. We have 
therefore proved that |a,y| = |b#|. 

COROLLARY. If a theorem has been proved concerning the rows of a 
general determinant, we may conclude at once that a similar theorem 
holds for the columns; and vice versa. 

THEOREM 5. The interchange of two columns (rows) of a deter- 
minant causes the value of the determinant to change sign. 

Proof. Let the given determinant be |o#| and let the deter- 
minant obtained from it by interchanging the columns whose 
indices are c\ and C2 be designated by |6#|. Then from every term 
of the former determinant, we can obtain one of the latter by 
writing &'s in place of a's and interchanging the column indices 
d and C2. It follows from Theorem 2 and Definition V that these 
two terms will be opposite in sign while equal in numerical value. 
Since moreover every term of |6#| can be obtained in this manner, 
our theorem has been proved; the alternate form, indicated in the 
parentheses, follows by application of the Corollary of Theorem 4. 

THEOREM 6. If all the elements of a row (column) are multiplied 
by the same number, the value of the determinant is multiplied by 
that number. 

This theorem is an immediate consequence of Definition V. 



ELEMENTARY THEOREMS 9 

THEOREM 7. If * two columns (rows) of a determinant are propor- 
tional, the value of the determinant is zero. 

Proof. Let us suppose first that the corresponding elements of 
the columns whose indices are Ci and C2, are equal. Let the value 
of the given determinant be A, and that of the determinant ob- 
tained from it by interchanging the columns of indices Ci and c 2 
be B. We conclude then from Theorem 5 that B = A; and 
from the fact that the two columns which have been interchanged 
are identical, that B = A. Therefore A = A and hence A = 0. 
If now two columns of a determinant are proportional, its value 
is equal, on account of Theorem 6, to a factor of proportionality 
multiplied by the value of a determinant in which two columns 
are identical; its value is therefore also equal to zero. 

THEOREM 8. If the elements of a column (row) of a determinant are 
binomials, its value is equal to the sum of the values of the two de- 
terminants which agree with the given determinant in every element 
except that the particular column (row) concerned consists in one of 
them of the first terms of the binomials and in the other one of the 
second terms. 

This theorem is also an immediate consequence of Definition V; 
the proof is left to the reader (see Section 6). For a determinant 
of the third order the theorem asserts among other facts that 



an i2 + ki 013 
(hi 022 + & 2 023 

31 032 + &3 033 



011 012 013 
021 022 023 
031 032 033 



0u ki ais 

021 & 2 023 
031 &3 033 



THEOREM 9. To the elements of any row (column) of a determinant 
may be added arbitrary multiples of the corresponding elements of any 
other row (column) without affecting the value of the determinant. 

Proof. It is a consequence of Theorem 8 that the value of the 
new determinant is equal to that of the given determinant plus 
the value of a determinant in which two rows (columns) are pro- 
portional. From this remark the present theorem follows by use 
of Theorem 7. 

Remark 1. Theorems 7 and 9 are the first objectives of the 
chain of theorems we are developing. The latter enables us to 
derive from a given determinant another one which is equal to it 
in value but in which all the elements but one of some one row 



10 



DETERMINANTS AND MATRICES 



or column are equal to zero; such a change materially reduces 
the labor involved in the evaluation of a determinant. 

Remark 2. To abbreviate our terminology we shall speak of the 
" addition of one row or column of a determinant to another row or 
column " with the meaning " addition of the elements of one row 
(column) to the corresponding elements of another row (column)/' 

Examples. 

4 -7 
1. To calculate thf* . the determinant 



1 
-3 



3 

-2 
2 



we use 



Theorem 9 as lc j o row 1 we add row 2 multiplied by 4; and to row 

3 we add row 2 mui^plied by 3. Thus we find that A is equal to the value of 
-27 11 , 

This determinant can readily be evalu- 



the determinant 1 52 
21 -4 

ated by means of Definition V; in this way we find that A = (27 
+ 11 - 1 21 = 123. 



1--4) 



2. To evaluate the determinant 



5 

-7 
2 



3 

6 

8 

-1 



2 

-I 
4 
3 



9 

10 

5 

4 



to column 1; 
6 



we, add column 3 



obtain 



4 

-3 
5 



and we add column 2, multiplied by 1, to column 4; thus we 
326 
6-14 



8 
-1 



4 -3 
3 5 



in which there are two equal columns. By 



Theorem 7 the value of this determinant is zero; hence it follows from Theorem 
9 that the value of the given determinant is also zero. 

6. Exercises. 

1. Show by an actual count of the inversions that the number of inversions 
is changed by an odd number when the numbers 2 and 8 are interchanged in 
the row 5 2 4 7 3 8 6 1. 

2. Prove Theorem 1 for the case p > q without assuming it for the case 

P < Q- 

3. Write out a detailed proof of the Corollary to Theorem 4. 

4. Write out a detailed proof of Theorem 6. 
6. Also for Theorem 8. 

6. Illustrate Theorems 4, 5, 7, and 9 by means of determinants of the 3rd 
and 4th orders. 

7. Evaluate each of the following determinants: 



(a) 



-523 

264 

-1 14 11 



1-1 2-2 
-32-23 
-131-3 

21-12 



; (c) 



2-231 

-3 4 17 -2 

5 -6 -9 3 

7 8 25 -4 



MINORS AND COFACTORS 



11 



8. Calculate the ^alue of each of the following determinants: 



(a) 



4323 

5-210 

-1431 

214-3 



(6) 



3 -1 -2 

3 -2 1 

2 -2 1 
-I 2 

3 1 



1 
-4 



7. Minors and Cofactors. The following theorems will enable 
us to reduce still further the arithmetical work involved in the 
evaluation of a determinant. They will moreover furnish a basis 
for the application of determinants to the solution of systems of 
linear equations. 

THEOREM 10. The determinant obtained from a given determinant 
by shifting rows and columns in such a way as to bring a certain ele- 
ment in the upper left-hand corner, without changing the relative 
position of the rows and columns which do not contain this element, 
has a value equal to that of the given determinant or to its negative 
according as the sum of the row and column indices of this element 
is even or odd. 



Proof. Let us consider first the determinant 1 011022033044 1 
let us call its value A. The upper left-hand corner of the deter- 
minant, now occupied by an will be called the " leading position. " 
To form a new determinant in which 34 is in the leading position, 
while the relative order of the 1st, 2nd, and 4th rows, and also of 
the 1st, 2nd, and 3rd columns, remains unchanged, we interchange 
the 3rd row successively with the 2nd and 1st rows; in virtue of 
Theorem 5, this operation causes the value of the determinant to 
change its sign twice, so that we can write 

031 032 033 034 

a\i 012 ai3 0i4 

021 022 023 024 
041 042 043 044 

Now we interchange the 4th column successively with the 3rd, 
2nd, and 1st columns; this leads to the desired result at the cost 
of three changes of sign. Therefore, the new determinant 

034 031 032 033 

014 011 012 013 

024 021 22 023 

044 4 i 042 043 



12 DETERMINANTS AND MATRICES 

in which a^ occupies the leading position while the rows and col- 
umns which do not contain a^ have the same relative order as in 
the original determinant, has a value equal to A. 

It should now be easy to understand the proof for the general 
case. To bring the element a# of the nth order determinant |a#| 
into the leading position without affecting the relative order of 
the rows and columns which do not contain this element, we inter- 
change the ith row successively with the (i l)th, (z 2)th, . . . , 
1st rows; then we interchange the jth column successively with 
the (j l)th, (j 2)th, . . . , 1st columns. This is accomplished 
by means of i 1 + j 1 = i + j 2 interchanges and there- 
fore accompanied by i + j 2 changes of sign in the value 
of the determinant. Consequently the determinant which we 
obtain finally will have a value equal to that of \aij\ or to its nega- 
tive according as i + j is even or odd ; this proves the theorem. 

DEFINITION VI. The minor of an element of a determinant of order 
n is the determinant of order n 1 obtained by deleting the row and 
column in which this element stands. 

DEFINITION VII. The cof actor of an element of a determinant is 
equal to its minor or to the negative of its minor according as the sum 
of the row and column indices of the element is even or odd. 

Notation. The value of the cofactor of the element a% is desig- 
nated by Atj. 

THEOREM 11. All the terms in the expansion of the determinant 
\aij\\ which contain a particular element as a factor are obtained, in 
magnitude and in sign, by multiplying that element by its cofactor. 

Proof. For the element an this theorem is an immediate con- 
sequence of Definition V. Let us again denote the value of 
| a^ | and also the determinant itself by A. To determine the sum 
of all the terms in the development of A which contain the element 
ay as a factor, we consider the determinant A' y obtained from A, 
as in Theorem 10, by putting a# in the leading position without 
affecting the relation of the rows and columns which do not con- 
tain this element. Then A' = (- 1)'+M and hence A = (l)*+*A'. 
Since a# occupies the leading position in A' the sum of the terms 
in A' which contain a# as a factor is obtained by multiplying a,y 
by its cofactor in A'. But the cofactor of a# in A f is equal to its 
minor in A'; and the minor of ay in A f is the same as the minor of 



MINORS AND COFACTORS 13 

this element in A', because the relative order of the rows and col- 
umns which do not contain the element a# has not been changed in 
the transition from A to A f . Therefore the sum of the terms in 
A f which contain the factor a# is obtained by multiplying this 
element by its minor in A. Moreover the terms in A which con- 
tain the factor a# are obtained by multiplying those in A f by 
( I)*' 4 "'". Therefore the sum of the terms in A which contain the 
factor ay is equal to ( 1 )*+*" X a# X the minor of a# in A = a,-. 
X the cof actor of a# in A = 



THEOREM 12. The value of a determinant is equal to the algebraic 
sum of the products obtained by multiplying the elements of any 
column (row) by then* cofactors. 

Proof. In every term of the expansion of a determinant, there 
is one and only one factor from each column (row). If therefore 
we select a column (row) arbitrarily and take the sum of all the 
terms which contain any one of its elements as a factor, we shall 
obtain the value of the determinant. Hence the present theorem 
is an immediate consequence of Theorem 11. 

THEOREM 13. The algebraic sum of the products of the elements of 
any column (row) by the cofactors of the corresponding elements of 
another column (row) is equal to zero. 

Proof. We observe that the cofactors of the elements of any 
column (row) are not affected by changes made in that column 
(row). The cofactors of the elements ai Cj , a^, . . . , a^ in |a#| 
are therefore the same as those of the elements a\ c ^ o^, . . . , o^ 
in the first of the columns so designated in the determinant 

021 ... 02* ... 02* ... 02 



nl ^Wj ^fk; a n 

This determinant is obtained from |a#| by replacing its column 
Ci by its column c^ (we are taking ci < Cz) and leaving everything 
else unchanged. On the one hand it follows from Theorem 7 that 
the value of this determinant is zero; on the other hand we con- 
clude from Theorem 12 and the remark made at the opening of 
this proof, that its value is equal to 

Qlc.Aic -f- QvcA%c. r i 



14 



DETERMINANTS AND MATRICES 



We conclude therefore that 

whenever c\ =|= c%. This proves our theorem. 

Theorems 12 and 13 are the final objectives of the chain of 
theorems which was started in Section 5. By means of Theorems 
9 and 12, the value of any numerical determinant can be deter- 
mined without an amount of arithmetical labor that is out of 
proportion to the Border of the determinant; Theorem 7 can fre- 
quently be used to reduce this labor still further. There are usu- 
ally several effective ways in which Theorem 9 can be used; by 
practice the reader will soon develop skill in applying it. 

Examples. 

2-13 
1. To determine the value A of the determinant 4 3 2 we add 

-3 2 1 

row 1 multiplied by 3 to row 2; and row 1 multiplied by 2 to row 3. We 
find then, by use of Theorem 9 that 

2-1 3 
-2 -7 
1 7 

It follows from Theorem 12 that A is equal to the sum of the elements of 
the 2nd column, each multiplied by its cofactor; but this sum reduces to the 
product of the element 1 by its cofactor. The row index of this element is 
1, its column index is 2; therefore its cofactor is equal to the negative of its 

-2 -7 
1 7 



minor. Therefore A = ( 1) X 



-144-7 = -7. 



Remark. The calculation of the value of a determinant by 
means of Theorem 12, as illustrated in the above example, is 
frequently called " developing the determinant according to a 
column (row)." It reduces the evaluation of any determinant 
to that of one of the next lower order. 



2. To evaluate the determinant 



1 
2 
3 
4 



1 

3 

6 

10 



1 

4 

10 

20 



(see Example 3, page 4), 



we add the first column multiplied by - 1 in turn to the 2nd, 3rd, and 4th 
columns; thus we find that the value A of the given determinant is equal 

1000 

1123 

1259 

1 3 9 19 



to that of the determinant 



and hence, by use of Theorem 



MINORS AND COFACTORS 



15 



12, to that of the third order determinant 



To evaluate this last 



1 2 3 

259 

3 9 19 

determinant, we add to the 2nd and 3rd columns respectively the 1st column 
multiplied by 2 and by 3. It is then found that 
1 



A = 



= 1 



8. Exercises. 



1. Illustrate Theorem 10 by means of determinants of the 3rd order and 
of the 4th order. 

5-70 



2. Calculate the value of the determinant 



by developing it, 



23-1 

642 

without previous reduction, according to the 1st row; also according to the 
3rd column. Verify that the two results are equal. 

3. Verify, in the determinant of Exercise 2, that the sum of the products of 
the elements of the 2nd column by the cofactors of the corresponding elements 
of the 1st column is equal to zero. 



4. Evaluate each of the 


following 


determinants by 


the cofactor method: 


(a) 


5 
4 
-1 


-2 3 
1 -2 
3 2 


; (b 


) 


17 
-5 
6 - 


8 3 
7 2 
11 -4 


'> 


(c) 




2 

(5 
-4 


3 4-1 
-2 5 2 
1-2 3 ' 
5 1 -3 


6. Calculate the values of the 


following determinants: 




18 


3 5 


-3 






-3 2 \ 


7 






4-24 


1 


(a) 


14 

8 


7 -4 

-2 6 


1 
2 


' 


(b) 


-1 J -4 
-3 5 


2 
3 


; 


(c) 


-2 1 -2 -2 
4 -2 4 | 




-10 


11 1 


5 






2 4 -3 


-10 






1 -2 | 


1 


6. Also of the following: 




2 8 


6 


14 


12 






2 -3 


4-25 






7 


1 


6 -4 






3 -4 


3 2 13 




(a) 


3 -6 


3 -5 





; (b) 




3 -4 


4 1 -2 


. 




1 4 


3 


7 


7 






4 - 


-5 


364 






-1 6 


11 


10 


23 






5 - 


-6 


3 9 -2 





7. Show that the cofactor of the element a rs in the determinant |o#|, i, j = 
1, 2, . . . , w, is equal to the determinant obtained from \<HJ\ by replacing the 
element OTS by 1 and all the other elements in the rth row (or in the sth 
column, or in both) by zeros. 

8. Prove that if the rows and columns of a determinant are shifted, as in 
Theorem 10, so as to bring the element Of S into the leading position, then the 
cofactor of any element in the new determinant is equal to the product of the 
cofactor of this same element in the given determinant by ( 1 )*+*. (Hint: 
Make use of the preceding exercise.) 



16 



DETERMINANTS AND MATRICES 



9. Matrices. Rank of a Matrix. Before proceeding to the 
application of the theory of determinants to the solution of sys- 
tems of linear equations, we shall introduce some further concepts, 
which, although perhaps not indispensable, will aid considerably 
not only in the solution of such systems but in all our further work. 

DEFINITION VIII. A matrix is a rectangular array of numbers. 
The numbers composing the array are called the elements of the 
matrix; the horizontal and vertical lines of the array are called re- 
spectively rows and columns of the matrix. 

Notation. In writing a matrix double vertical bars are placed 
on either side of the array. Large parentheses are sometimes 
used instead of the double vertical bars; we shall adhere to the 
former notation. For example 










5 


3 


2 


3 -111 


4 


-1 


2 


-4 


5 6||' 


-3 


2 


-1 






6 





4 



and 







2 

v/7 
-4 
-6 



-3 

6 




2 

-1 
5 



are matrices. 

Abbreviated notations, similar to those used for determinants 
(see Section 2) are also used for matrices; for example, ||o#||, i = 1, 
. . . , 5; j = 1, . . . , 4 represents a matrix of 5 rows and 4 columns 
and ||o$||, i, j = 1, 2, . . . , n represents a square matrix of n rows 
and n columns. We shall also designate a matrix by the single 
letters a or b. 

Remark. We emphasize the fact that a matrix is merely an 
array of numbers and that no numerical value is attached to it. 
In particular it is important to notice the difference between a 
square matrix and a determinant. Although both are square ar- 
rays of numbers, the latter has a number associated with it, 
namely, its "value," but the former has no number associated 
with it. A determinant whose elements are identical with the 
corresponding elements of a square matrix is called the "determi- 
nant of the matrix." We also speak in such a case of the "matrix 
of the determinant." 

DEFINITION IX. The rank of a matrix is a positive integer or zero, 
r, such that it is possible to form a determinant of order r whose value 
is different from zero and whose rows and columns are obtained from 
the rows and columns of the matrix, whereas it is not possible to form 
a determinant of order r + 1 which satisfies the same conditions. 



COMPLEMENTARY MINORS 



17 



Remark. It is an immediate consequence of this definition that 
if the rank of a matrix is r, then the value of every determinant of 
order r + 1, r + 2, etc., whose rows and columns are formed from 
the rows and columns of the matrix, will be zero. And if the rank 
of a matrix is zero, all its elements are zero. 

10. Complementary Minors. Elementary Transformation of 
Matrices. 

DEFINITION X. Any determinant whose rows and columns are 
formed from the rows and columns of a matrix (determinant) Is 
called a minor of the matrix (determinant). 

The terms " two-rowed minor," "three-rowed minor," etc., 
which we shall have frequent occasion to use, should be clear 
without further explanation. 

Remark. The definition of the rank of a matrix can now be put 
in the following form: The rank of a matrix is an integer, r, posi- 
tive or zero, such that the matrix has a non-vanishing r-rowed 
minor but no non-vanishing minor of order higher than r. 

DEFINITION XI. A principal minor of a square matrix (determinant) 
Is a minor formed by using rows and columns of equal Indices only. 

DEFINITION XII. If the rows and columns used In forming the minor 
Af 2 of a square matrix (determinant) are those which were left unused 
In the formation of the minor M 19 then Mi and M 2 are a pair of com- 
pic mem ary minors of the matrix (determinant). Either is the com- 
plement of the other. 

DEFINITION XIII. The algebraic complement of a minor of a square 
matrix (determinant) is equal to its complement multiplied by that 
power of -1 whose exponent is equal to the sum of the indices of the 
rows and columns used In the formation of the minor. 

Remark. The one-rowed principal minors of a square matrix 
(determinant) are the elements of its principal diagonal ; the comple- 
ment of a single element is its minor; the algebraic complement of a 
single element is its cofactor. (Compare Definitions VI and VII.) 

Examples. 

s OH a 



1. The determinants 



0*3 
4 3 



4 6 



and 



2 3 
063 



2 4 

M 



2 6 
O&6 



are two-rowed and 



three-rowed minors respectively of the determinant \04j\ t i,j 1, . . . , 5. 



2. The determinants 



On 

031 



012 
032 



OH 

034 
064 



and 



2 3 
043 



2 6 
4 6 



are cornplementarv 



18 



DETERMINANTS AND MATRICES 



minors of the determinant \aij\, i, j = 1, . . . , 5; and 

principal minor of the same determinant. 
3. The algebraic complement of the two-rowed minor 

terminant |oj/ , i, j = 1, . . . , 5 is equal to ( 1)2+4+3+5 



therefore equal to the complement of this two-rowed minor. 





On QIS 


Ois 




31 33 


035 is a 




o a 63 


O65 


23 25 ^ 


the de- 


4 3 46 




On 012 o 


14 


0i 32 


34 ; it is 



4. The algebraic complement of the three-rowed minor 



the determinant |a#|, i, j = 1, . . . , 6 is equal to (-1) 
and therefore equal to the negative of its complement. 



12 OH 

a& 034 



062 



of 



O21 O23 2 
a 4 l #43 O 4 



The operations upon the rows and columns of a determinant 
which were discussed in Section 5 may also be performed upon the 
rows and columns of a matrix. But since a matrix has no " value," 
the theorems on determinants which were there obtained will have 
no exact analogues; we shall however be interested in the effect 
of these operations upon the rank of the matrix. We introduce 
first the following Definition. 

DEFINITION XIV. An elementary transformation of a matrix Is one 
which consists in performing upon it one of the following operations: 
the interchange of two rows (columns); the multiplication of all the 
elements of a row (column) by a non-vanishing multiplier; the addi- 
tion to the elements of one row (column) of multiples of the corre- 
sponding elements of another row (column). 

Remark. It should be clear that if the matrix b is obtained from 
the matrix a by an elementary transformation, then the matrix a 
is obtainable from b by an elementary transformation. 

THEOREM 14. The rank of a matrix is not affected by an elementary 
transformation. 



Proof. Let us consider the matrix a = ||a#||, i,j=l,.... 9 n 
and let us suppose that its rank is r; it will then contain at least 
one non-vanishing r-rowed minor, while every (r + l)-rowed mi- 
nor vanishes. The interchange of two rows or columns of a and 
Ihe multiplication of the elements of a row or column of a by a 



COMPLEMENTARY MINORS 19 

non-zero constant either have no effect whatever upon its minors, 
or else they will multiply a minor by a non-zero constant; in 
neither case will these operations kill off a non-vanishing minor of 
a nor bring a vanishing minor back to life. These operations will 
therefore leave the rank r of the matrix unchanged. If to the ith 
row of a we add k times the jth row, an (r + l)-rowed minor of & 
will not be changed in value if it does not contain the ith row, nor 
if it contains both the ith and the jth rows. Let us suppose there- 
fore that M is an (r + l)-rowed minor of a which contains the 
ith row but not the jth row; and let us denote by M' the corre- 
sponding minor of the matrix a' obtained from a by adding k times 
the jth row to the ith row. Then it follows from Theorem 8 that 
the value of M ' is equal to the value of M plus k times the value of 
another (r + l)-rowed minor of a; but since every (r + l)-rowed 
minor of a vanishes, it follows from this that the value of M ' is also 
zero. Consequently every (r + l)-rowed minor of a' vanishes, so 
that the rank of a' ^ r ; that is, the rank of a matrix is not increased 
by any elementary transformation. But then it follows from the 
remark preceding this theorem that the rank must remain un- 
changed. For if it were decreased then the elementary transfor- 
mation which carries the new matrix back to the original would 
have to increase the rank; and we have just seen that this can not 
happen. The theorem has therefore been proved. 

COROLLARY. If a matrix a' is derived from another matrix a by a 
succession of elementary transformations the ranks of the two ma- 
trices are equal. 

Remark. This theorem and its corollary can be used in the de- 
termination of the rank of a matrix in the same way as Theorems 
5, 6, and 9 are used in the evaluation of a determinant. 

11. Exercises. 

1. Write out the minors of \\atj\\, i, j = 1, . . . , 6 formed by using the 
following sets of rows and columns: i = 1, 2, 5, j = 2, 3, 6; i = 2, 3, 4, 6, 
j = 2,4,5,6; i = 3,6,,; =3,6. 

2. Determine the algebraic complements of each of the minors of Exercise 1. 

-1 23-5 



3. Show that the rank of the matrix 



3-45 2 
5 -6 13 -1 
2 14 -13 



is 2. 



20 



DETERMINANTS AND MATRICES 



4. Determine the rank of each of the following matrices: 



(a) 



(c) 



-2 3 5 
5 -1 -3 
479 



3 -4 
6 -8 
9 -12 



(W 



W) 



2 




3 




-4 






4 




6 




-5 






6 




9 




-9 








1 







-3 


5 


-2 


3 


2 


-4 




7 


-6 


-13 


25 




1 







-3 


3 



6. Prove that the algebraic complement of a principal minor of a square 
matrix is equal to its complement. 

6. Prove that one of two complementary minors of a square matrix is a 
principal minor if and only if the other one is a principal minor. 

7. Prove that if the minor MI of a square matrix is the algebraic complement 
of the minor M 2} then M z is also the algebraic complement of MI. 

12. The Laplace Development of a Determinant. In the re- 
maining sections of this chapter we shall develop some further 
interesting and important properties of determinants. These 
properties will find application in the later chapters, but they are 
not needed for the solution of systems of linear equations. The 
reader can proceed therefore from this point immediately to 
Chapters II, III, IV, and V, returning to the remainder of Chapter 
I after he has completed these. 

Our first objective is a generalization of the cofactor develop- 
ment of a determinant, discussed in Section 7. 

LEMMA 1. The determinant obtained from a given determinant by 
shifting the rows and columns in such a way as to bring a specified 
fc-rowed minor in the upper left-hand corner without changing the 
relative position of the rows and columns not involved in this minor 
has a value equal to that of the given determinant or of its negative 
according as the sum of the indices of the rows and columns used in 
this minor is even or odd. 

Proof. Let the indices of the rows and columns used in the 
fc-rowed minor under consideration be r i9 r 2 , . . . , r^ and ci, <% 9 
. . . , cu respectively. To accomplish our purpose, we interchange 
the nth row successively with each of the r\ 1 rows which lie 
above it; next we interchange the r 2 th row successively with each 
of the r 2 2 rows which lie above it but below the 1st row, the 
rath row with each of the r 8 3 rows which lie above it but below 
the 2nd row, etc., until we have interchanged the r*th row with 



LAPLACE DEVELOPMENT OF A DETERMINANT 21 

each of the rj& k tows which lie above it but below the (k l)th 
row. Thus the rows whose indices are n, r 2 , . . . , n have been 
placed in the positions of the first k rows, while the relative posi- 
tion of the remaining rows has remained unchanged; and this 
has been done by means of ri 1 + r 2 2 + . . + r* ft 
interchanges of rows, so that the determinant we have obtained 
has a value equal to that of the given determinant multiplied by 
(_!)**.+ **-*<*>/. We p roc eed now to shift the columns 
whose indices are fi, 02, . . . , Ck in such a way as to bring them in 
the position of the first k columns without affecting the relative 
order of the remaining columns; it should be easy to see that this 
is accomplished by means of d 1 + <% "~ 2 + . . . + Ck k in- 
terchanges of columns and therefore at the cost of c\ + c% + . . . + 
c^ k(k + l)/2 changes of sign. The final result in which the 
specified fc-rowed minor is in the upper left-hand corner and in 
which the rows and columns not occurring in this minor have the 
same relative order as in the given determinant has therefore a 
value equal to that of the given determinant multiplied by a power 
of 1 whose exponent is r\ + r 2 + . . . + r^ + Ci + 2 + - + 
Ck k(k + 1). But, no matter what integer k may be, k(k + 1) 
is always even. Consequently the value of the final determinant 
is equal to that of the given determinant if r\ + r 2 + . . . + r* + 
Ci + 02 + . . . + Ck is even, and equal to its negative if this sum 
is odd. 

Remark. Fork = 1, this lemma and its proof reduce to Theorem 
10 and its proof. 

LEMMA 2. All the terms In the development of a determinant which 
contain as a factor any term in the development of a specified fc-rowed 
minor are obtained in the product of this minor by its algebraic com- 
plement; and this product contains nothing but such terms of the 
development of the determinant. 

Proof. We shall prove this proposition first for the principal 
/c-rowed minor in the upper left-hand corner; and we shall denote 
this minor temporarily by Ak- The algebraic complement of this 
minor is equal to its complement (see Exercise 5, Section 11). 
An arbitrary term in the development of this minor is ( l) c ai Cl O2<; 2 
. . . a,kc k , where Ci, c&, . . . , Ck is a permutation of the numbers 
1, 2, . . . , k and c is the number of inversions in this permutation; 
an arbitrary term in the development of its complement is 



22 DETERMINANTS AND MATRICES 

(-1) a k+ i t y k+1 a k+2t y k+2 . . . a nyn , where y k +i t ik+z, , 7* rep- 
resents an arbitrary permutation of the set of numbers 
k + 1, k + 2, . . . , n and 7 is the number of inversions of this 
permutation.* The product of these two terms is (-I)* 4 "* a^o^ 
. . . dkciflk+i, ^+i a *+2, y k+z a vn m Since the numbers of the set 
Cij 02, . . . , Ck are all less than those of the set 7^+1, 7^+2, . . . , 7, 
it follows from Exercise 4, Section 4 that the number of inver- 
sions of the total set ci, 02, . . . , c*, 7^+1, 7^4-2, . . . , 7 W is equal to 
c + 7; hence this product is a term in the development of 
the original determinant. If, on the other hand, ( l^a^o^. 
. . . a<nd n is a term in this development which contains as a factor 
a term of A k) then its first k factors must be elements of A k and 
therefore di, d 2 , . . . , d k and d^+i, d*+ 2 , . . . , d w must be permu- 
tations of the sets 1, 2, . . . , k and k + 1, k + 2, . . . , n respec- 
tively. Hence a^fr^ . . . &u k and a^+i. ^+^+2, d k + 2 . . . Und n will be 
terms in the developments of Ak and of its complement re- 
spectively, and the numerical factors, +1 or 1, will be such 
that their product is equal to ( !)*. Our lemma has been 
proved therefore for the principal minor A&. 

To prove it for an arbitrary fc-rowed minor B k formed from the 
rows and columns whose indices are n, r 2 , . . . , r* and d, (%, 
. . . , Ck respectively, we form first, as in Lemma 1, the de- 
terminant in which Bk occupies the upper left-hand corner. Let 
us call this new determinant, and also its value, A'; then A' = 
( 1)<*-KH- . . . +cj+n+n+ . . . r h A an( j ^ e m i n or Bk of A goes over 
into the minor Ak of A'. Moreover the complement of Bk in A 
is the same as the complement of Ak in A'. In virtue of these 
facts and of the first part of this proof, we conclude that the sum of 
the terms in A which contain a term of Bk as a factor is equal to 
(- !)+*+ qk-Hi+rH- -'kXAk X the complement of A k ' in A' 
= (-!)++ . . . -^H-n+n-f . . . +r k x ^ x t h e complement of B k in 
A = BkX the algebraic complement of B*. This completes the 
proof of the lemma. 

* In the development of the algebraic complement the sign of this term is 
determined by the number of inversions of the set of integers obtained from 
^k r ^k v J 7 n ^ diminishing each of them by k; but it follows from 
Exercise 3, Section 4 that this new set of integers has the same number of 
inversions as the set y k+l , y k+z , . . . , y n . 



LAPLACE DEVELOPMENT OF A DETERMINANT 



23 



Remark. The special case of this lemma which arises when k = 
1 is identical with Theorem 11. 

THEROEM 15. The value of a determinant is equal to the algebraic 
sum of the products obtained by multiplying each of the /c-rowed 
minors that can be formed from any k rows (columns) of the deter- 
minant by their algebraic complements. 

Proof. Let us consider the rows whose indices are r\ 9 r 2 , . . . ,r>. 
Every term in the development of the determinant will contain as 
a factor a product of k elements selected from these k rows, one 
from each; and every such product will be a term in the develop- 
ment of some one fc-rowed minor whose row indices are r\ y r 2 , 
. . . , Tk. Hence we shall obtain the value of the determinant if 
we take the sum of all the terms which contain as a factor any 
term in the development of any one of these fc-rowed minors. 
But, since it was shown in Lemma 2 that for a given fc-rowed 
minor all such terms are found, and without any additional terms, 
in the product of this minor by its algebraic complement, we can 
conclude that the sum of the products of all the &-rowed minors 
formed from the fc rows, which were selected, by their algebraic 
complements is equal to the value of the determinant. 

Remark. The evaluation of a determinant by the method ex- 
plained above is called the Laplace development of the deter- 
minant. For the case k = 1, it reduces to the development accord- 
ing to a row or column discussed in Theorem 12. 

Example. 



The determinant 





-6, 



a 3 
63 
3 




can be evaluated in a very con- 



venient way by means of the Laplace development, 
columns, we find that its value is 



If we use the first two 



ai I 
-ai I 

I -ai 
M -a 2 -6 2 



c 3 
-c 3 



X 



02 



X 







6 2 6 3 
-f 3 



-f 



(a 2 6 3 - a 3 6 2 ) 2 - 



a 3 6 3 
bs - a 3 6 2 ) = 



-a s -6 3 

a 2 i 
a 3 I 



a 3 6 2 ) 2 . 



X 



6 2 6 3 1 



a 3 



24 



DETERMINANTS AND MATRICES 



13. Exercises. 

1. Evaluate each of the following determinants by the Laplace develop- 
ment, using the first two rows: 

ai bi Ct 

c\ 



-2 
3 

-5 
1 



53 

41 

2-4 

-30 



(6) 




3 

-6 
4 



-1 
5 
3 
2 



7 

-2 





-4 
1 

-5 
3 



(c) 



/, 



2. Prove that the sum of the products of the fc-rowed minors formed from 
k columns of a determinant by the algebraic complements of the correspond- 
ing fc-rowed minors of another set of k columns (that is, a set of k columns in 
which there is at least one column that was not among the columns of the 
first set) is equal to zero. 

3. Evaluate each of the following determinants by Laplace's development: 



(a) 



3 
-6 
-3 
4 




1 
-2 


-5 7 
4 -2 
6 8 
5 -3 


'' 


(ft) 


-2 
3 
5 

-8 
-4 


5 4 
1 -2 
-7 10 
4 -1 
fi 3 


6 
8 
3 
12 
-2 



5 

-9 









1 3 


-8 4 

















4 -5 


2 7 


-3 












(c) 


-5 


1 2 


5 














6 


4 -5 


1 














-2 


-3 3 


-4 









4. Prove that the value of a determinant of order n, which contains for 
any integral value of fc, a matrix of k rows (columns) and n k + 1 columns 
(rows) of which every element is zero, is itself zero. 

6. If a determinant of order n contains a matrix of k rows (columns) and 
n k columns (rows) of which every element is zero, the value of the deter- 
minant is equal to the product of the values of a single fc-rowed minor and its 
algebraic complement. 

6. Prove that the algebraic complement of a specified A>rowed minor of a 
determinant is equal to the determinant obtained from the given one by re- 
placing by 1 the elements in the principal diagonal of this fc-rowed minor and 
by all the other elements of throws (columns) from which the minor is 
formed. 

7. Prove that if s is the sum of the row and column indices of a certain 
fc-rowed minor of a determinant and if a new determinant is formed from the 
given one in which this fc-rowed minor has been shifted to the upper left- 
hand corner by the method of Lemma 1, Section 12, then the dofactors of any 
element o# in the two determinants differ only by the factor ( I) 5 . (Hint: 
Compare Section 8, Exercise 8.) 

8. Prove that under the conditions of the preceding problem the algebraic 
complements of two corresponding fc-rowed minors of the two determinants 
differ only by the factor ( 1)*. 



THE PRODUCT OF TWO DETERMINANTS 25 

14. The Product of Two Determinants. The Laplace develop- 
ment enables us to express the product of two determinants of 
order n as a determinant of order 2 n. We denote by 



p = 



= 1,2,. 



I' \bij\ 

the determinant of order 2 n whose first n rows consist of the rows 
of |%-|, each extended by n zeros, while the last n rows consist of 
the rows of |6#| preceded by an arbitrary square matrix of order 
n; it should be easy to see, particularly in view of Exercise 5, 
Section 13, that the value of P is equal to the product of the values 
of \ciij\ and |6#|, that is, P = A B, where A and B designate 
the values of the determinants |o#| and |6#| respectively.* 

We choose now for the matrix /' the n-rowed square matrix in 
which the elements in the principal diagonal are all equal to 1 
and all the other elements are zero. We shall designate by 
Ci, C 2 , . . . , C, C+i, . . . , Gin the matrices of one column and 
2n rows each formed by the successive columns of P. Moreover 
we shall introduce an abbreviation, current in all mathematical 
writing and probably familiar to many readers, for a sum of terms 
which differ in subscript only; namely, we shall write, in general, 

n 

V Uk for the sum u\ + w 2 + . . w- 
*=i 

We proceed now to apply Theorem 9 to change the form of the 

determinant P as follows: we replace C n +i by C n+ i + bnCi + 621^2 + 
. + b n iC nj that is, we add to the (n + l)th column of P each 
of the first n columns after having multiplied them by &n, 6 2 i, . . . , 
b n i respectively. The (n + l)th column will then consist of the 
following elements: an&n + 012621 + . . . + i&ni> &2i&n + 022621 + 

, , . . . , 0; or, 



using the abbreviation which has just been explained, of 

n n 

5) 2*&*i y > 2 nkbki, 0,0,. . . , 0. Next we replace the col- 
*~i A=I 

umn C n +2 by C w + 2 + fc^C^ + 622^2 + . . . + bniCn, the column 
C w +, by C n + 3 + fciaC'i + & 2 3C 2 + + bnaCn, and so forth, until 

* The reader is advised to write out in full the determinant P and to carry 
out the operations described concisely and with abbreviated notation in the 
following paragraphs. 



26 



DETERMINANTS AND MATRICES 



we have replaced the last column of P, namely, C 2n , by C 2n + 
binCi + & 2n <7 2 + . . . + b nn C n , Thus we obtain the following 
result : 



P = AB = 



'21 


22 


tt<t n A 


L 01*6*1 1 

r . .- ^ 

*. #2*6*1 


^ #1*6*2 . . . 2y 
^ 02*6*2 2, 


dlkbkn 
(hkbkn 


.i 


fl2 . 


. . a nn ^ 


L Q>nkbkl A 


J a n ^* 2 ... 2 


&nkbkn 


-1 


0. 


. . 


0- 


... 








-1 . 


. . 





... 








0. 


. .-1 


0- 


... 






in which all the sums designated by the symbol 2) arc to be ex- 
tended over the range k = 1, . . . , n. This determinant is de- 
veloped by Laplace's development using the last n columns; in 
accordance with Exercise 5, Section 13, the result will be equal to 
the product of the n-rowed minor in the upper right-hand corner 
by its algebraic complement. Thus we find that 

P = 



Thus we have shown that the product of the two nth order deter- 
minants |a#| and |6#| is equal to an nth order determinant in which 
the element in the ith row and jth column is equal to the sum 

n 

This result is stated in the following theorem. 

THEOREM 16. The product of two determinants of order n is equal 
to a determinant of order n in which the element of the ith row and 
jth column is equal to the sum of the products of the elements in the 
ith row of the first factor by the corresponding elements of the jth col- 
umn of the second factor. 

Remark. We know from Theorem 4 that the value of a determi- 
nant is not changed if the columns are taken as rows and the rows 
as columns. Hence if we rewrite A (or 5, or both) in such a way 
as to interchange rows and columns and then determine their 



THE PRODUCT OF TWO DETERMINANTS 



27 



product in accordance with Theorem 16, we obtain the following 
extension of this theorem: 

The product of two nth order determinants is an nth order de- 
terminant whose element in the ith row and jih column is 

(1) equal to the sum of the products of the elements in the 

z'th row of the first by the corresponding elements of the 
jth column of the second; or 

(2) equal to the sum of the products of the elements of the 

ith column of the first by the corresponding elements of 
the jih column of the second; or 

(3) equal to the sum of the products of the elements in the 

z'th row of the first by the corresponding elements in the 
jth row of the second; or 

(4) equal to the sum of the products of the elements in the 

ith column of the first by the corresponding elements in 
the jih row of the second. 

This statement is expressed symbolically by the following for- 
mula: 

AB = 

A condensed form of the statement is that determinants may be 
multiplied rows by columns, or columns by columns, or rows by 
rows, or columns by rows. In most cases we shall use the first of 
these methods for multiplying two determinants and, unless the con- 
trary is explicitly stated, it is to be understood that this is the case. 

Examples. 

1. If |ay |, i,j 1, 2, 3 is used to denote the product of the two determi- 



3-56 -17-3 

nants 2 -1 and 6 2 
-489 5-41 

the products of the elements in the second row of the first of these by the 
corresponding elements of the third column of the second; therefore a 2 s 
2 . (-3) 4- o 2 + (-1) 1 = -7. Similarly we find a a2 = (-4) -7 + 8- 
+ 9- (-4) = -16. 
2. The product, rows by columns, of the two preceding determinants is 

27 -33 -13 

-7 18 
49 -16 



then a 2 3 is equal to the sum of 



6 



13 




27 -33 -13 




-1 39 -41 


-7 


= 


-7 18 -7 




-7 18 -7 


37 




110 -12 




110 -12 


-1 39 -41 




-255 280 


= 27,740. 


110 -12 





28 



DETERMINANTS AND MATRICES 



Their product, columns by columns, is 



-23 49 -9 
45 -67 23 
39 -11 



= 4X 



1 

-6 
18 -55 -11 



4 -9 
48 23 



Their product, rows by rows, is 



-56 -18 41 

1 -2 9 

33 66 -43 



-56 -130 545 

100 

33 132 -340 



1 

-6 72 -31 
18 -127 151 



-56 -130 545 

100 

-23 2 205 



= 27,740. 



= 27,740. 



And their product, columns by rows, is equal to 



23 

-19 
-40 



4 3 
16 -17 
12 43 



= 4 X 



= 4X 



1 
-111 4 
-109 3 
2 63 
109 -34 





-29 
34 

= 27,740. 



I 111 29 
1109 -34 



15. Exercises. 



1. Multiply the determinants 

(1) rows by columns; 

(2) columns by rows. 

2. Multiply the determinants 

(1) rows by rows; 

(2) columns by columns. 



1 
3 
9 

27 



1 

4 

16 

64 



and 



and 



3. Determine the square and the cube of the determinant 



1 

2 
2 



-I 


2 




1 

2 
4 



2 
2 

3 


2 
1 



4. Form the products in each of the four possible ways of two general second- 
order determinants and show that the results are equal in value. 

5. Prove that a two-rowed minor of the product of two determinants is 
equal to a sum of products of two-rowed minors of the two determinants. 

6. Prove a theorem for three-rowed minors similar to the theorem of 
Exercise 5. 

7. Prove the general theorem of which the theorems of Exercises 5 and 6 
are special cases. 

8. Prove that the rank of the matrix of a product of two determinants can 
not exceed the rank of the matrix of either factor." 

16. The Adjoint of a Determinant. 

DEFINITION XV. The determinant of which the element in the ith 
row and j th column ( i, j = 1, . . . , n) is equal to the value of the co- 



THE ADJOINT OF A DETERMINANT 29 

factor Aij of the element o# of the determinant |o#|, i, j = 1, . . . , n 
Is called the adjoint of that determinant. 

We shall denote by A and by A' the values of the detenninant 
\a,ij\ and its adjoint respectively. By means of Theorems 16, 12, 
and 13 we should be able to see readily that the product rows by 
rows of | a^ | and its adjoint forms a determinant in which the ele- 
ments of the principal diagonal are all equal to A, and the remain- 
ing elements are all zero; hence that A* A' = A n and therefore 
that A' = A"" 1 , if A 4= 0. This result gives us the following 
theorem. 

THEOREM 17. If the value of the determinant \<HJ\, i 9 j = 1, . . . , n 
is different from zero, then the value of the adjoint of this determinant 
Is equal to the (n l)th power of the value of |a//|. 

The theorem which we have just proved is a special case of the 
following more general theorem. 

THEOREM 18. The value of a fc-rowed minor of the adjoint of a de- 
terminant Is equal to the product of the value of the algebraic com- 
plement of the corresponding fc-rowed minor of the given determinant 
by the (k - l)th power of the value of the given determinant, pro- 
vided this latter value is different from zero. 

Proof. We will prove this theorem first for the A-rowed prin- 
cipal minor in the upper left-hand corner of the adjoint. We 
begin by forming the product of |a#| by the determinant 



An An 


... A hl 





... 





AM A& 


... A to 





... 





A\k Ay, 


... A kk 





... 





Ai, k+i At, t+i 


... At, k+\ 


1 


... 





A,, ^ A t , M 


At, 4 +2 





1 ... 





A ltt A*, 


... At. 





... 


1 



The Laplace development of this determinant which uses the first 
fc rows, together with Theorem 4, shows that its value is equal to 
that of the fc-rowed principal minor which we wish to determine 
(compare Exercise 5, Section 13) ; let us denote this value by V*. 
The product of |a#| by this detenninant is formed in accordance 



30 



DETERMINANTS AND MATP Tri ES 



With Theorem 16; if we make use again of Theorems 12 and 13, 
we obtain the following result: 



A 





. . . 


l, A+l 


01, A+2 - 


In 





A 


. . . 


02, A+l 


O2, A+2 0; 


In 








. . . A 


0A, A+l 


OA, A+2 . . . a 


kn 








. . . 


OA+I, A+i 


OA+1, A+2 ... 


A+l> ft 








... 


0A+2, A+l 


0A+2, A+2 ... 


A+2, 








. . . 


On, A+l 


Oft, A+2 . 


nn 



If the determinant on the right-hand side of this equation is de- 
veloped by Laplace's development, using the first k columns, we 
find, by thinking once more of Exercise 5, Section 13, that it re- 

Ofc+lj A+l flA+li 

duces to A* 

#*, A+l a nn 

pothesis, we conclude that 



hence, since A =f= by hy- 



V k 



A+l 



*+li 



A+i 



X 



Now the determinant on the right of this equation is the algebraic 
complement of the minor of |o#| which corresponds to the minor 
of its adjoint which we are having under consideration; the theorem 
has therefore been proved for this special case. 

Let us now consider an arbitrary fc-rowed minor Mk f of the ad- 
joint; let Mk be the corresponding minor of |a#| and let m^ be its 
algebraic complement. Let us furthermore denote by s* the sum 
of the row and column indices used in Mk (and therefore in M*')- 
We know then from Lemma 1, Section 12 that the value of the 
determinant obtained from \ay\ by a shifting of rows and columns 
which brings the minor M k into the upper left-hand corner without 
altering the relative order of the rows and columns not involved 
in this minor is equal to (l)'kA ; and from Exercise 7, Section 13 
we infer that the cofactors of the elements of this rearranged de- 
terminant differ from the cof actors jrf the same elements in |a#| 
by the factor ( l) f *. Therefore, if MI! denotes the fc-rowed prin- 
cipal minor in the upper left-hand corner of the adjoint of the 



THE DERIVATIVE OF A DETERMINANT 31 



rearranged determinant, every element of M*' is equal to (jj 
times the corresponding element of Mu f and hence MI! = 
( 1)***M*'. To the minor M*' we can apply the conclusion 
reached in the first part of the present proof; the algebraic com- 
plement of the corresponding fc-rowed minor in the rearranged 
determinant |a#| is identical with the complement of Af* and hence 
equal to ( l)**w*. We have therefore the following result: 
(-l^kMk* = (-l)*km k X [(-l) s kA] k ~ l , from which we conclude 
that Mk = mkA k - 1 } this is the relation asserted in our theorem. 
Remark. For k = n, Theorem 18 reduces to Theorem 17; for 
fc = 1, it merely asserts that every element of the adjoint is equal 
to the cofactor of the corresponding element of the given deter- 
minant |%|. If we denote by a# the cofactor of the element Ay of 
the adjoint, we obtain by putting k = n 1, the following 
corollary. 

COROLLARY. The cofactor aij of the element AIJ in the adjoint of the 
determinant \aij\ is equal to aijA*~*. 

17. The Derivative of a Determinant. The elements of the 
determinants whose properties we have been discussing have been 
constants. If these elements are functions of a single variable t } 
let us call them tty(t), then the value of the determinant is also a 
function of this variable. Denoting this function by the symbol 
17(0 > we have 



We inquire now for a convenient form in which to write the de- 
rivative of U(t) with respect to t. To obtain such a form, we 
recall two facts: 

(1) That 7(0 is the sum of terms =bui^(0ti^(0 . . . w^CO* 
in which ci, c&, . . . , c* are successively the different permutations 
of the set of integers, 1,2,. . . , n and the sign of the term depends 
on the character of the permutation; 

(2) That the derivative of a product of two or more functions 
is equal to the sum of all the products obtainable from the given 
product by replacing one factor at the time by its derivative; for 
example, if ' denotes differentiation with respect to t, (uiUju*)' =* 

UiUjUs + UiUJUs + UiUjUs. 

From these facts we conclude that U'() is equal to the sum 
of the sums % Wi'^, . . . u^ 2) =fc u^^ . . . u*c n , . . , 



32 



DETERMINANTS AND MATRICES 



=fc ui c U2c t - u n c n , in each of which ci, 02, . . . , c are succes- 
sively the different permutations of the set of integers 1,2,. . . , n 
and the sign of the term is plus or minus according as the number 
of inversions in the permutation is even or odd. But then it 
follows from (1) that the first of these sums is the expansion of a 
determinant obtainable from |w#(0| by replacing the elements in 
its first row by their derivatives; also, that the second sum is the 
expansion of the determinant obtainable from |w#(0| by replacing 
the elements in the second row by their derivatives; and so forth. 
We can therefore state the following answer to our inquiry. 

THEOREM 19. The derivative with respect to t of the determinant 
U(t] = \uij(t)\, i 9 j = 1, . . . , u is equal to the sum of the n deter- 
minants obtained from U(t) by replacing the elements of one row 
(column) at the time by their derivatives. 



Examples. 

1. The adjoint of the determinant 



2 -1 
1 



is found to be 



1 -2 -1 
-7 1 5 
-142 



1 -3 2 

The value, A, of the first determinant is 9; and the value, A', of the ad- 
joint is 81 = ( 9) 2 . The cofactor of the element in the 3rd row and 2nd 
_4 



column of the adjoint is 



- 
/ 



; its value is 27, which is also the prod- 



uct of the corresponding element in the original determinant, 3, by the 1st 
power of the value of the determinant, 9. 
2. The derivative of the value of the determinant 

t 2 - 3 t 2t 2/ 2 - 7t + 4 

3 t 2 + I 4 2 6 t - 3 

2t + l 2 t* + 6 1 - 3 t 2 + 8 t - 7 

is equal to the value of the sum 

2t -3 2t 2 / 2 - 7 / + 4 

6 / 4 / 2 t - 3 



2 



St-7 



t 2 - 3 1 2 2 


/ 2 -7/ +4 


+ 3 / 2 + 1 8 / 


6 / 3 


21 + 1 4/ + 6 


t*+8t- 7 


2t 4J-7 






4/2 6 







t* -3t 
3/ 2 + l 
2t +1 2 

18. Exercises. 

1. Determine the jtdjoints of each of the following determinants: 



1 A 

1-4 
-4 5 

5 l 



(6) 



0-13 

42-1 



3-141 



(c) 



12-2 1 
21 3-4 
4 5-1-2 
308-9 



THE DERIVATIVE OF A DETERMINANT 
2. Differentiate each of the following determinants: 



33 



(a) 



| sin I cos t 
I cos J sin t 



-I* + 4 
VC + 1) 



ai cos 61 sin / C| 

a-2 sin t b-i cos Z c 2 

c t c 2 



3. Determine the 1st, 2nd, and 3rd derivatives of the determinant 



n - t 
021 



12 

o 22 - t 

032 



033 - 



4. A symmetric determinant being defined as one in which, for every pair 
of indices i and j, y = cy,-, show that the adjoint of a symmetric determinant 
is itself symmetric. 

6. Show that the value of the adjoint of the determinant 

tti2 Ol 

Ol2 #2 

13 2 3 

is zero and the rank of its matrix is 1. 

6. Prove that the value of the determinant 

ait + pi ait + 7> 2 ait -f /> 3 
ait -f- qi a-it -f o 2 a4 -f q* 
a 3 t -f TI azt + ^2 o 3 + r 3 

is a function of the first degree in f. 

7. Verify, by direct computation, the Corollary to Theorem 18 for the case 
of 3rd order determinants. 

8. Work out a formula for the 2nd derivative of the determinant |w#(OI> 
i, j = 1, . . . , n. 



19. Miscellaneous Exercises. 

1. Determine the value of each of the following determinants: 



-4 2 3 
5 -3 2 

7 1 -6 



; (6) 



4-13 2 

12 5 -7 4 

-3 6 5 -9 

14 15 9-10 



3 2 -1 -2 

2010 

-11 3 2 

-2020 



2. Determine the rank of each of the following matrices: 



(a) 



2 -3 5 6 
1 24-5 
4 -13 7 28 


; W 


-321-2 
2-1 3 4 
1 11 8 
, 1 70-5 


; (c) 


4-241 
-2 1 -2 -2 

4-24f 
1 -2 * 1 



3. Compute the adjoin ts of the determinants (a) and (c) in Exercise 1. 



34 



DETERMINANTS AND MATRICES 



4. Determine, by inspection, the sum of the values of the determinants 

-323 
and 1-32 
-6 1 -6 



6. Show that 



4 


2 


3 


-1 


-3 


2 


6 


1 


-6 


1 1 


1 


1 


a b 


c 


d 


a' 2 b 2 


c 2 


d* 


n* b* 


c 3 


d* 



= (a-b) (b-c) (c-d) (a-c) (6-d) (a-d). 



6. Determine the value of the determinant 



1 

On 



7. Show that the square of the value of the determinant in Exercise 5 

4 Si S 2 S 3 
Si S 2 Sa 84 



is equal to the value of the determinant 



S4 



, where sk = a k -\- b k 



+ c* + d*, for fc = 1, . . . , 6. 

8. Using the notation sk, introduced in the preceding exercise, set up an 
wth order determinant which is equal in value to the square of the value of 
the determinant in Exercise 6. 



9. Show that if w 3 = 1, 



1 



1 



0. What is the rank of the matrix 



of this determinant? 

10. Prove that if the elements a# of a determinant \aij\ are independent 
variables then the partial derivative of its value with respect to a particular 
element is equal to the cofactor of this element. 

11. Prove that the second partial derivative of the value of a determinant 
\aij\ with respect to the variables a t j and OTS is equal to the algebraic comple- 

J " 



ment of the two-rowed minor 



dfj 



12. Prove that the fcth partial derivative of the value of the determinant 
\aij\ with respect to the elements a fiCv Or^t, , ^ r k c k ^ wmc h no two have 
the same row-index nor the same column-index, is equal to the algebraic 
complement of the fc-rowed minor whose rows and columns have the indices 
r,, r 2| . . . , n and ci, C2, . . . , ck respectively. 

* A determinant of this form is frequently referred to as a Vandermonde 
determinant. 



CHAPTER It 
LINEAR EQUATIONS 

20. Definition and Notation. 

DEFINITION I. An equation In one or more variables Is called /iomo- 
geneous If, after the right-hand side has been reduced to zero, the 
terms on the left-hand side are all of the same degree in all the vari- 
ables jointly. 

Remark. It follows from this definition that a linear equation 
(that is, an equation of the first degree in all the variables jointly) 
is homogeneous if and only if it contains no term independent of 
the variables. 

Notation. We shall be dealing with systems of equations in n 
variables; the variables will be designated by Xi, Xz, . . , x n . A 
linear equation will therefore have the form : 

+ 02X2 4- . . . + ax n = k. 



It will be homogeneous if and only if k = 0. 

A system of linear equations will be written in the form : 



a p2 x 2 + . . . + a pn x n = k p . 

The first subscript in each coefficient designates the equation in 
which it occurs; the second subscript indicates the variable which 
the coefficient multiplies. It will be a system of homogeneous 
equations if and only if ki = k 2 = . . . = kp = 0. 

It is convenient to designate the entire system of equations 
briefly by writing 

anxi + a i2 x 2 + . . . + a in x n = fc,-, i = 1, 2, . . . , p. 

The coefficients o&, a,- w . . . , a,v, i = 1, 2, . . . , p form a ma- 
trix of p rows and n columns; this matrix will be called the coeffi- 
cient matrix (abbreviated c.m.) of the system of equations. If 
p = n, this matrix will be a square matrix; the corresponding 

35 



36 LINEAR EQUATIONS 

determinant is |o#|, i 9 j=l,. . . , /*>, which will be called the coeffi- 
cient determinant of the system, and its value will be indicated, 
as in Chapter I, by A. If we write not merely the coefficients of 
the variables, but also the known terms which appear on the right- 
hand sides of the equations, we obtain a matrix of p rows and n + I 
columns; this matrix will be called the augmented matrix (abbrevi- 
ated a.m.) of the system of equations. It will be a square matrix 
if and only if p = n + 1, that is, if the number of equations in the 
system is one greater than the number of variables. 

21. The System of n Linear Non-homogeneous Equations in n 
variables. 

THEOREM 1. If the coefficient determinant of a system of n linear 
equations in n variables has a value A, different from zero, then the 
system has a single solution, consisting of one value for each of the 
variables; these are equal to fractions whose denominators are all 
equal to A, and whose numerators are the values of the determinants 
obtained from the coefficient determinant by replacing the coefficient 
of each variable in turn by the known terms as they appear on the 
right-hand side of the equations. 

Proof. Let the equations be written in the form 
anXi + a 12 x 2 + . . . + a in x n = fci, 

(1) 



a n2 X 2 + . . . 

If they are multiplied by the cofactors An, A 2i , . . . , A n i of the 
elements an, 021, . . . , a n i in the first column of the coefficient 
determinant and then added, it follows from Theorems 12 and 13 
of Chapter I that the coefficient of x\ in the sum is equal to A, 
while the coefficients of x 2 , x$, . . . , x n are all zero. Therefore we 
find that 

A xi = kiA n + kzA 2i + . . . + k n A ni . 



If we use as multipliers the cofactors Aw, A 22 , . . . , A n2 of the 
elements in the second column of the coefficient determinant and 
add, we find that 

A x 2 = kiAiz + ktAw + . . . + k n A m ; 

and by the use of the cofactors of the elements in the third, fourth, 
. , . , nth columns as multipliers, we obtain n 2 further equa- 



LINEAR NON-HOMOGENEOUS EQUATIONS 37 

tions of the same general form. They may be written simultane- 
ously in the form : 

(2) A Xi = kiA }i + k 2 A Z i + . . . + k n A ni , i = 1, 2, . . . , n. 

Any set of values of the variables x\, x 2; . . . , x n which satisfy 
equations (1) must satisfy these conditions. Since A 4= 0, there 
is one and only one value for each x,-; and if we recall the observa- 
tion made at the opening of the proof of Theorem 13 of Chapter 
I (see page 13), we will recognize that the right-hand side of (2) 
is the expansion according to its z'th column of the determinant 
obtained from the coefficient determinant by replacing its Oh 
column by the constants &i, 2, . . . , k n on the right-hand sides 
of the given equations. Consequently, if A =J= 0, the system oi 
equations (1) can not have more than one solution, namely, the 
one given by the values 

. . . + k n A ni 



It remains to show that the values given by (3) actually do 
satisfy the equations (1). Substitution of these values in the 
left-hand side of the rth equation of this system gives, by repeated 
use of the abbreviated notation for sums introduced in Section 
14 (see page 25) 



n n 

V a H T kjAjt V kj V 



A A 

n 

But it follows from Theorems 12 and 13 that V a^A^ is equal to 



zero if j 3= T and equal to A if j = r. Hence the only term in the 
sum which is different from zero is the one obtained for j = r, so 
that it reduces to k,A and the values given by (3) do therefore 
actually satisfy the given equations. This completes the proof of 
the theorem. 

Remark. The rule given by this theorem for writing down at 
once the solution of a system of n linear equations in n variables 
whose coefficient determinant has a value different from zero, is 
known as Cramer's rule. 



38 LINEAR EQUATIONS 

22. The System of n Linear Homogeneous Equations in n 
Variables. From the result obtained in the preceding section, we 
derive some important consequences. 

THEOREM 2. A system of n linear homogeneous equations In n 
variables whose coefficient determinant has a value different from 
zero, possesses the solution which consists of zero for each of the 
variables, and no others. 

Proof. It should be clear by inspection that, if ki = A* 2 
= . . . = k n = 0, the equations of the system (1) are satisfied by 
the values Xi = x% = =# = (). That the system has no 
other solution if A =f= follows from the proof of Theorem 1. 

Remark. It should be evident that every system of p linear 
homogeneous equations in n variables Xi, x 2 , . . . , x n possesses 
the solution x\ x 2 = = x n = 0; this solution of such a 
system is called the trivial solution. 

On account of the frequent use to be made of it in the later 
parts of this book (see Sections 41 and 82) we state the following 
corollary which is an immediate consequence of the preceding 
theorem. 

COROLLARY. In order that a system of n linear homogeneous equa- 
tions in n variables shall have a non-trivial solution, it is necessary 
that the value of its coefficient determinant shall be zero* 

23. The System of n + 1 Linear Non-homogeneous Equations 
in n Variables. 

THEOREM 3. A system of n -f 1 linear equations in n variables, 
whose coefficient matrix has rank n, possesses a solution if and only 
if the determinant of its augmented matrix has the value zero; and 
in this case the solution is unique. 

Proof. Let the equations be written in the form: 

= &;, Z = 1, 2, . . . , ft + 1. 



Since the c.m. is of rank n, we can find in it at least one deter- 
minant of order n, which has a non-zero value. And, because 
the order in which the equations that constitute the system are 
written is clearly a matter of indifference, we can suppose without 
loss of generality that this determinant is the coefficient deter- 
minant of the first n equations of the system. These equations 
possess therefore a single solution, which can be written down 



LINEAR NON-HOMOGENEOUS EQUATIONS 39 

according to Cramer's rule; consequently the entire system of 
n + 1 equations can not have more than one solution. It will 
possess one solution if the values of the variables determined by 
the first n equations also satisfy the (n + l)th equation. Now it 
is possible to express these values in terms of the cofactors of the 
elements in the last row of the a.m. If we use the familiar capital 
letter notation to designate the cofactors of the elements of the 
a.m., we find from the first n equations, since K n+i 4= 0: 

n #12 . . . flj, i-i ki ai, 1+1 . . . ai, n -i CL\. 



. . a ni n _i 

l, w-l 



tl fl n 2 . . . (l nj i-l MID i+1 Km n-1 &nn K> 

The latter of these determinants is obtained from the former by 
interchanging its ith column successively with each of the n i 
columns which follow it. But the last written determinant is 
clearly the minor of the element a w +i, t - in the a.m. and therefore 
equal to ( l) n + l +* times the cofactor A n +i, , of this element. We 
conclude therefore that 

K X' = ( i)H t'+n+i+f A 4-\ ' A -LI i = 1 2 n. 

Now we substitute the values of Xi obtained in this way in the last 
equation of the system. We find then that the system possesses 
a solution if and only if a n +\,iA n +iji + a n +i, 2 A n +i,2 + ' 
+ a rt+1 , n A n +i, n + k n +i K n +i = 0. But this is, in virtue of Theorem 
12, Chapter I, the condition that the determinant of the augmented 
matrix be equal to zero. We have therefore proved the theorem. 

COROLLARY. A system of n -f 1 linear homogeneous equations In n 
variables whose coefficient matrix has rank n possesses only the trivial 
solution. 

Examples. 

1. The system of equations 3s-2?/ = 4, 2z + 32/ = 5, x - y = 2 ha,s 
no solution. For, while the c.m. is clearly of rank 2 (the value of the determi- 

Io o i 

is different from zero), the determinant of the augmented 
6 O I 



40 



LINEAR EQUATIONS 



matrix has a value different from zero; this determinant is 



3-24 
2 35 
1 -1 2 



and 



its value is 11. 

2. Let us consider the system of equations 3 x 2y + z = 7, 2x -\- 3y 
4 2 = g t x y + z*=4,x + 2y + 3z = 5. The rank of the c.m. is 3; 
and the determinant formed of the coefficients of the first three equations, 

has the value + 4. The determinant of the a.m. is 



It should not be difficult to show that its value is zero. 



This being done, it follows from Theorem 3 that the system has a single solu- 
tion, which may be found by solving the first three equations of the system 
by Cramer's rule. Thus we find, by use of Theorem 1, that 





3 -2 1 


namely, 


23-4 




1 -1 1 


3-217 




2 3-4-9 




1-114 




1235 





7 -2 1 
-9 3 -4 
4 -1 1 


4 
4 


3 -2 1 
23-4 
1 -1 1 



3 7 1 
2 -9 -4 
1 4 1 



3 -2 1 
2 3 -4 
I -1 1 



-4 
4 



3-27 
23-9 
1 -1 4 



3 -2 1 
2 3 -4 
1 -1 1 



24. Exercises. 

1. Give illustrations of systems of homogeneous and of non-homogeneous 
linear equations in 2, 3, and 4 variables. 

2. Determine the rank of each of the following matrices: 

3 -2 



(a) 



3 -4 

-1 2 

5 -6 



5 2 

5 1 

19 5 



5 -2 
-4 10 
15 -6 



(c) 



1 

-1 432 
7-870 
1 6 16 4 

3. Solve each of the following systems of equations by Cramer's rule: 

4. Proceed in the same way with the following systems of equations: 
(a) 3x + y-4u=9, -5 y + 3z + 2u = 18, x - 6y + 7z = 



(6) 



= 20, 
+ 5 v = 50. 



6. Determine, for each of the following systems of equations, whether they 
possess a solution; solve those systems for which a solution exists: 
(a) 2x-y + l =0, s + 2y + 2 = 0, 15 x + 20 y + 24 = 0. 
(6) x-2/ + 4=0, 3 
(c) 5x-3y-7 = 



LINEAR HOMOGENEOUS EQUATIONS 



41 



6. Also for each of the following systems: 

(a) 6 + 5 ?/ -- 12 2= 11, 5s 2y 

x + 18 y - Sz = -24. 

(6) 2x + 3y + 3s = l, 33-y-4z = 
-4z + ll2/ + 2l2 = -9. 

(c) 3z-i/ + 2z= -3, -2z-f2?/-3 



-3, 
-4, 



26. The System of FI 1 Linear Homogeneous Equations in 
n Variables. 

THEOREM 4. If the rank of the coefficient matrix of a system of 
n 1 linear homogeneous equations in n variables is equal to n 1, 
the system has a single infinitude of solutions; the ratios of the 
variables are equal to the ratios of the cofactors of the elements in 
the nth row of the determinant obtained by writing a row of arbitrary 
elements pi, p 2 , . . . , p n under the coefficient matrix. 

Proof. We write the equations in the form 

aaxi + a i2 x<i + + ai n x n = 0, i = 1, 2, . . . , n 1. 



Since the c.m. is of rank n 1, it contains a non-vanishing de- 
terminant of order n 1, and there is no loss of generality if we 
suppose that this is the determinant formed by the coefficients 
of the variables xi, x 2 , . . . , x n -\. In virtue of Theorem 1, the 
equations possess therefore a single solution for x\, x^ . . . , x n -\ 
for every definite value assigned to x n ] hence there is a single 
infinitude of solutions. 

We can apply Cramer's rule to solve the equations for 
sci, #2, . . . , x n -i in terms of x n \ we find 



an 

021 



022 



t-l #2; +l 



n- 



, 1 n-l>2 



= x n Pi y i = 1, 2, . . . , n, where PI, P 2 , . , P denote the co- 
factors of the arbitrary elements pi, p 2 , . . . , p n in the determinant 
obtained from the c.m. by writing a row of arbitrary elements 



42 LINEAR EQUATIONS 

under it. If we write this result in the form of a continued pro- 
portion, we obtain the result stated in the theorem, namely: 

xi : x 2 : . . . :x n = Pi :P 2 : . . . : P n . 

COROLLARY. If the rank of the coefficient matrix of a system of p 
linear homogeneous equations in n variables is n - 1, the system has 
a single infinitude of solutions; the ratios of the variables are equal 
to the ratios of the cof actors of the elements in some row of an n-rowed 
minor of the coefficient matrix. 

Proof. If the hypothesis is satisfied, it is clear that we must have 
p ^ n 1 and that there must be at least one set of n 1 among 
the p equations which satisfies the conditions of Theorem 4. The 
order in which the equations of the system are written is imma- 
terial and we may therefore suppose that the first n 1 equations 
of the system furnish one such set. We will show now that the 
solutions of this set also satisfy the remaining equations of the 
original system of p equations. For, if we substitute kP iy i = 1 , 
2, . . . , n for Xi in the rth equation (r n, n + 1, . . . , p), the 
left-hand side becomes k (a r fi + <7 f2 P 2 + + a rn P n } ; but the ex- 
pression in parentheses is the expansion of the nth order determinant 
obtained by writing the coefficients of the rth equation under the 
c.m. of the first n 1 equations, so that its value is equal to zero. 
Moreover the set of the first n 1 equations has, in virtue of 
Theorem 4, no other solutions besides those of the single infinitude 
indicated above; consequently solutions which are obtained by 
using another set of n 1 equations selected from the given 
system must be contained among this single infinitude. Hence 
our corollary is proved. 

26. The Adjoint of a Vanishing Determinant. Symmetric De- 
terminants. By means of Theorem 4 we are able to obtain a 
valuable extension of Theorems 17 and 18 of Chapter I. For in 
these theorems we had to make the hypothesis that the value of 
the determinant under consideration was different from zero; 
it is this restriction which we are now able to remove. We begin 
by proving the following theorem. 

THEOREM 5. If the value of a determinant is zero, the rank of the 
matrix of the adjoint of the determinant is equal to or 1. 

Proof. If the cofactor of every element of the determinant 
vanishes, the rank of the matrix of the adjoint is clearly equal to 



ADJOINT OF A VANISHING DETERMINANT 43 

zero; if this is not the case, let us suppose that Aj^ ^ 0. Then 
for every i the equations 



= o 

+ 22^-i2 + ' ' ' + 2n^m 

l*n = 
+ CL n2 Ai2 + * + dnnAfn = 

hold, in virtue of Theorems 12 and 13 of Chapter I. But these 
equations may be looked upon as a system of n 1 linear homo- 
geneous equations in the n variables AH, A&, . . . , A,- rt ; and since 
the coefficient matrix surely contains the determinant Ak r , its rank 
is n 1. We can therefore apply Theorem 4 and we find that 

AH : A i2 : . . . : A in = A kl : A k z : . . . : A knj for i = 1, 2, . . . , n. 

This means that the different rows of the adjoint are proportional, 
so that every two-rowed minor vanishes; the rank of the adjoint 
is therefore less than 2. 

COROLLARY. The conclusions of Theorems 17 and 18 of Chapter I 
still hold when the value of the determinant |o#| is zero. 

We shall now prove a few important consequences of this theorem 
which refer to symmetric square matrices (usually called symmet- 
ric determinants) and which are needed in our later work (see 
Chapter VIII). 

DEFINITION II. A symmetric square matrix Is a matrix \\aij\\, i } = 1* 
2, . . . , n, in which, for every i and every j, aij = a/,-.* 

DEFINITION III. A singular square matrix is one whose determinant 
vanishes. 

THEOREM 6. If all the (n - l)-rowed principal minors of a singular 
symmetric square matrix vanish, its rank is less than n - 1. 

Proof. We have to show that under the hypothesis every ele- 
ment of the adjoint vanishes. Now it should be easy to see that 
the adjoint of the given symmetric square matrix is itself sym- 
metric, so that Aij = Aji for every i and j.* Moreover it follows 

* Compare Exercise 4, Section 18. 



44 LINEAR EQUATIONS 

from Theorem 5 that every two-rowed minor of the adjoint has 



the value zero, that is, 



= for every i and j. Hence, 

= Aij 2 . But AH AJJ= 0, by hypothesis. Consequently 
Aij = 0; and this is what our theorem asserts. 

COROLLARY. If the rank of a singular symmetric square matrix of 
order n is n 1, then it contains a least one no ti- vanishing (n 1)- 
rowed principal minor. 

THEOREM 7. All the (n - 1) -rowed principal minors of a singular 
symmetric square matrix which do not vanish have the same sign. 

Proof. If AH and A# are two non-vanishing principal minors, 
then it follows from the hypothesis, as in the proof of Theorem 6, 
that A#Ajj = Aij 2 > 0; hence AH and A$ have the same sign. 

COROLLARY. If the sum of the (n - 1) -rowed principal minors of a 
singular symmetric square matrix is equal to zero, the rank of the 
matrix is less than n - 1. 

This corollary follows immediately from Theorems 7 and 6. 

27. The System of n Linear Non-homogeneous Equations in 
n Variables, continued. In Section 21 we have seen that a system 
of n linear non-homogeneous equations in n variables, whose c.m. 
has rank n possesses a single solution; this solution may be de- 
termined by means of Cramer's rule. We return now to such a 
system of equations but under the hypothesis that the rank of the 
c.m. is n 1, and we shall prove the following theorem. 

THEOREM 8. If the coefficient matrix of a system of n linear non- 
homogeneous equations in n variables is of rank n 1, the system will 
have a single Infinitude of solutions or no solution, according as the 
rank of the augmented matrix is equal to or greater than n - 1. 

Proof. We write the equations in the form 

+ aaxz + + a in x n = k if i = 1, 2, . . . , n. 



Since the rank of the c.m. is n 1, there is at least one set of n 1 
of the equations and at least one set of n 1 of the variables, 
such that the c.m. of these variables in this particular set of equa- 
tions is of rank n 1 ; these equations can therefore be solved by 
Cramer's rule for n 1 of the variables in terms of the nth vari- 
able, as soon as a value has been assigned to this variable. Con- 



LINEAR NON-HOMOGENEOUS EQUATIONS 



45 



sequently this set of n 1 equations possesses a single infinitude 
of solutions. It remains to determine whether these solutions will 
also satisfy the single remaining equation. This question can be 
answered by means of Theorem 3. 

If the special set of n I variables consists of i, 2 , . . . , a^-i, 
X/+1, . . . #, the determinant of the augmented matrix of the re- 
lated system of n equations in n 1 variables is 



Oil 

021 



022 



0*1 



On . . 
021 . 



fliw K, 
. . Oi* 



0,1 . . 
On 



. am 



But since the rank of the c.m. of the given system is n 1, the 
last term vanishes; therefore the determinant of the augmented 
matrix of the related system reduces to the first term, which is an 
n-rowed minor of the augmented matrix of the given system. If 
the rank of this matrix is n 1, then every one of its n-rowed 
minors vanishes; hence the augmented matrix of the related sys- 
tem has rank n 1 and we conclude, by use of Theorem 3, that 
all the solutions of the special set of n 1 equations also satisfy 
the nth equation. But if the rank of this matrix is n, then, for 
some one of the variables Xj, the determinant of the augmented 
matrix of the related system will be different from zero, no matter 
what value is assigned to xy 9 and in this case we conclude, again 
by use of Theorem 3, that the given system of equations does not 
possess a solution. This completes the proof of the theorem. 

Remark. The theorems proved in Sections 21, 22, 23, 25, and 
27 are special cases of a more general theorem which asserts that 
a system of p linear equations in n variables possesses one or more 
solutions if and only if the ranks of the c.m. and the a.m. are equal. 
A proof of this theorem may be found in the books referred to in 



46 



LINEAR EQUATIONS 



Section 1. The special cases dealt with here suffice for the appli- 
cations to be made in the later chapters of this book; they furnish 
moreover a suitable introduction to the study of the more general 
cases. For this reason and also in order to avoid too elaborate 
algebraic discussions, we have restricted ourselves to these special 
cases. The reader is urged to make himself thoroughly familiar 
with the content and the proofs of these theorems; they will 
repeatedly be referred to in our further work. 
Examples. 

1. The system of equations 



determines the ratios of the variables x, y, and z. For the value of the coeffi- 
2-3 5 



cient determinant 



is readily found to be zero. Since the two- 



2 1 
4 7-7 

rowed minor in the upper right-hand corner is different from zero, the rank 
of the c.m. is 2. Consequently, we conclude from Theorem 4 and its corollary 
that the ratios of the variables are equal to the ratios of the cof actors of the 
elements in the last row. Thus we find that 



x:y:z- 



-3 5 
2 -1 



2 5 

3 -1 



2 -3 

3 2 



- -7 : 17 : 13. 



It is easily verified that any three numbers which have these ratios satisfy the 
given equations. 

2. To determine the ratios of z, y, and z from the system of equations 

x - 2 ;// + 3 z = 0, -3 z -h 6 ?/ 4- 2 = 

we observe first of all that the rank of the c.m. is 2. Jn view of Theorem 4 
we can conclude that the ratios of the variables arc equal to the ratios of the 

1-23 
cofactors of the elements in the third row of the determinant 



hence we find that 



-3 

Pi 



6 1 

Pt Pa 



x ' y . z = 



-2 3 
6 1 



1 
-3 



1 -2 
-3 6 



= -20 : -10: = 2 : 1 : 0. 



^ 3. In virtue of Theorem 8, the system of equations 

2# 37/4-52 = 1, 3 # 4- 2 ?/ z = 4, 4x -\- 7 y 7 z 5 

has no solution. For it should be easy to show that the rank of the c.m. is 2; 

2 -3 1 



and since the value of the determinant 
-26, the rank of the a.m. is 3. 



2 4 



4 7 



formed from the a.m. is 



LINEAR NON-HOMOGENEOUS EQUATIONS 47 

4. In the system of equations 



the ranks of the c.m. and the a.m. are both equal to 2. It follows therefore 
from Theorem 8 that the system possesses a single infinitude of solutions which 
may be found by solving two of the equations for two of the variables in terms 
of the third. If we solve the first two equations for x and y in terms of z, 
we find that 

14 - 7 z . 5 + 17 z 

* = -T3- and = J5 

It is readily verified that these expressions satisfy the three given equations 
for all values of z. 

28. Exercises. 

1. Determine the ratios of the variables from each of the following systems 
of homogeneous equations: 

(a) x 4-3 y -z = 0, -2y +z = 0, 5x + y + 2z = 0. 

GO 4x4- 67/4-82 -84^ =0, 2x + y+3z -48?; = 0, -2 x + y + z 

- 12 v = 0, 4jc4-4y-2-24i;=0. 
(c) 2 x - ?/ 4- 2 z = 0, -x 4- 2 ?/ 4- 2 z = 0, 2 x -f 2 y - 2 = 0. 

2. Proceed similarly with each of the following systems: 
(a) 2x- 7/4-22 =0, x + 2y + 2z =0. 

(&) 5 2: - y + 3 2 = 0, 10 x - 2 y + 4 z = 0. 

(c) 2aj-y-f2jg-t-5tt = l 3a:-f2/-^+2^^ = 0, 4^-2y-h3?/ = 0. 

3. Show that none of the following systems of equations possess a solution: 

(a) 3 x - 4 y = 5, 2 x + 5 y = 3, 6 x - 8 ?/ = 4. 

(b) 2x + 3y - 4z = 3, s-2y + 3 = l, 4o: + 7/-2z=~2, 2o; 

-4y + 5s = 3. 

(c) 5 x - y 4- 2 z = 12, 2 x- + 3 y = 7, 3 x - 2 y - 4 z = -2, 4 s -f T/ 

-2 z = 5. 

4. Determine which of the following systems possess a solution; solve 
those for which a solution exists: 



(a) z4-6 
(6) 4x + 
(c) 2x~ 

6. Determine the conditions under which the system of equations 
aix -\- b\y -f CiZ -f ^ = 0, a 2 x + ^2?/ + C 2 2 + ^2 = 0, 
a& + hy 4- c 8 2 4- ds = 0, o 4 x 4- &4?y 4- 42 4- rf 4 = 0, 

possess one solution. 



48 



LINEAR EQUATIONS 



6. Determine the ranks of the adjoints of the following square matrices: 



-1 

2 
1 

o 

7. Prove: 



2 
3 

-2 
-3 



1 

-2 
-1 

2 



-2 
-3 

2 
4 



(W 



-1 
2 
1 

-2 



2 
3 

-2 
-3 



1 

-2 
-1 

2 



-2 

-3 

2 

3 



(a) If the rank of the matrix 



p u 
u q 



is 2, then the values of no two 



of the expressions pq u 2 , qr w 2 and rp v 2 are opposite in sign. 



p u v 

(b) If u q w 
v w r 
of the matrix in (a) is 1. 



= and pq -f- qr -J- rp u 2 v 2 to 2 = 0, the rank 



8. Prove that if the system of equations 

(liiXi 4- 



= fa, I = 1, 2, . . . , ?i 

in which the value of the coefficient determinant |a#| is different from zero, 
is solved for xi, x 2 , . . . , x in terms of k h k 2 , . . . , k n , then the value of the 
determinant formed by the coefficients of ki, fc 2 , . . . , k n is equal to the recipro- 
cal of the value of the determinant \a#\. 



CHAPTER III 
POINTS AND LINES 

The primary object of Solid Analytical Geometry is the study of 
the geometry of three-dimensional space by algebraic methods. 
This end is accomplished by means of coordinate systems or frames 
of reference. Such systems enable us, as in Plane Analytical 
Geometry, to determine algebraic entities corresponding to various 
geometric elements. We start this study with the simplest geo- 
metric element, the point in three-space. 

29. The Cartesian Coordinates of a Point in Three-space. 
The simplest frame of reference is furnished by three mutually 
perpendicular planes, called the coordinate planes. It is custom- 




Fia. 1 

ary to take one of these planes horizontal, the other two vertical. 
The point common to the three planes is called the origin of co- 
ordinates and is usually designated by the letter O. The lines of 
intersection of the planes with each other are called the coordinate 
axes, the X- and Y-axes being the intersections of the horizontal 
plane with the two vertical planes, and the Z-axis the line in which 
the two vertical planes meet (see Fig. 1). 

It follows from elementary solid geometry that the three co- 
ordinate axes are mutually perpendicular and that each of them 
is perpendicular to the plane formed by the other two. In this 
book we shall take the positive directions on the coordinate axes 

49 



4/0 POINTS AND LINES 

as indicated in Fig. 1.* On d^PpPl^se axes a unit of measure- 
ment is adopted; we shall use equal^||fc on the three axes. 

If an arbitrary point || is now-t||||lin three-space, we drop 
from it perpendiculars ^%e^8rdSfelanes. Let the feet of 
these perpendiculars be designated by P^y, P yz , and P^, the sub- 
scripts indicating the planes in which these points lie (see Fig. 1). 
Now we lay down the following definition : 

DEFINITION I. The x-coordinate of P is the measure of the line 
PyzP, measured in accordance with the unit and the direction specified 
for the X-axis; the ^coordinate is the measure of the line P ZX P 9 
measured in accordance with the unit and direction specified for the 
Y-axis; and the ^-coordinate is the measure of the line P xy P measured 
in accordance with the unit and direction specified for the Z-axis. 

Notation. It will frequently be found convenient to designate 
the coordinates of the point P as XP, yp and ZP, particularly when 
ready identification of the points is desired. When the reference 
is to an arbitrary point of a specified group of points, we shall 
usually omit the subscript; the coordinates of a point PI will be 
denoted by Xi, y\ y z\\ those of a point P 2 by 2 , 2/2, 2 ; etc. 

Remark 1. The coordinates of a point P are signed real numbers, 
the signs depending upon the position of P with respect to the 
coordinate planes. The ^-coordinate of P is positive or negative 
according as P lies to the right or to the left of the 7Z-plane; 
the ^/-coordinate of P is positive or negative according as P lies 
in front of or behind the ZX-plane; the ^-coordinate is positive 
or negative according as P lies above or below the XT-plane. 

Remark 2. It should be clear from Definition I not only that 
every point in three-space has a definite set of three real numbers 
as coordinates, but also that an arbitrary set of real numbers, taken 
in a definite order, determines one and only one point in three- 

* The coordinate system which we have adopted is called a " left-handed 
system " because the thumb, first and second fingers of the left hand can be 
put in such a position as to suggest the positive directions along the X-, 
Y-, and -axes, particularly by a person whose finger joints have not grown 
stiff. If the positive direction along any one of the axes is reversed, we ob- 
tain a right-handed system. It should be clear that any two left-handed 
systems, and also any two right-handed systems, can be made to coincide; 
but that a left-handed system and a right-handed system are symmetric with 
respect to each other and can not be brought to coincidence if we are limited 
to a three-dimensional space. 



THE COORDINATE PARALLELOPIPED OF A POINT 51 

space. The reader should^llfece himself that this point is 
found as the point comnxj^^ three mutually perpendicular planes. 
The point P whose ^-//Iffijbd ^-coordinates are a, 6, and c re- 
spectively will be desig^g^Pby th$ g^tol P(a, 6, c). 

30. The Coordinate Parallelepiped of ,a Point. The three per- 
pendiculars dropped from P on the coordinate planes determine, 
two by two, three mutually perpendicular planes. These three 
planes together with the coordinate planes determine a rectangular 
parallelepiped; we shall call this the coordinate parallelepiped 
of P, a name which we shall frequently indicate by c.p. 

The coordinate axes each meet the faces of the c.p. in and in 
a second point; these points are designated by P x , P y , P z (see 
Fig. 1). The twelve edges of the c.p. are equal, four by four, to 
the lines whose measures are the coordinates of P. The four 
pairs of opposite vertices of the c.p. are and P, P xy and P z , P yz 
and P XJ P& and P y ; the body diagonals are the four lines which 
join pairs of opposite vertices. The lines PP*, PP y , and PP 2 are 
perpendicular to the X-, Y-, and Z-axes respectively; the points 
P x , P y , and P s are therefore the projections of P on the coordinate 
axes. 

If we bear in mind the properties of the rectangular parallelepiped 
which are proved in elementary solid geometry, we obtain at once 
the following theorems. 

THEOREM 1. The *-, y, and ^-coordinates of P are equal respec- 
tively to the projections OP*, OP y , and OP Z of OP upon the X-, Y-, 
and Z-axes. 

THEOREM 2. The square of the distance OP Is equal to the sum of 
the squares of the coordinates of Pi 

OP 2 = * P 2 + y/ + z p \ 

THEOREM 3. The cosines of the angles which the line OP makes 
with the positive directions of the X-, Y-, and Z-axes are equal re- 
spectively to the quotients of xp, y p , and z p by OP. 

If we designate these angles by a op , @ OP , and y op respectively, 
and their cosines by A OP , ju op , and v op respectively, we have : 



z p 



52 POINTS AND LINES 

Remark 1. The square roots in these formulas are to be taken 
with the positive sign; if the sign of the square root is changed, 
we obtain the cosines of the angles which the line OP makes with 
the negative directions along the coordinate axes, or, what amounts 
to the same, the cosines of the angles which the line PO makes with 
the positive directions along the axes. 

Remark 2. It is important that the reader should learn to draw 
the coordinate parallelepipeds of points in various parts of space. 
The following exercises are intended chiefly to develop skill in 
doing this. * 

31. Exercises. 

1. Draw the coordinate parallelepiped for each of the following points; 
determine their distances from the origin and the cosines of the angles which 
the directed lines from the origin to them make with the positive directions 
along the coordinate axes: 

4(-2,5, -4); B(l, -2,3); C(-l, -2,3); D(0, 4, -2); #(4,6,7); 
F(-4, -6, -7); 0(5, -2, -1); ff(-3, 4, 2); /(6, 0, -3); 
K(, 3, -4); L(-5, -2, -3); M(-5, 3, 0). 

2. Determine the loci of the points for which 

(a) x = -2; (6) y = 4; (c) z = -5; (d) x = 4 and y = -3; (e)y = 2 
and z = 6; (/) z = 3 and x = -3; (g) x 1 + y 2 -f z 2 = 9. 

3. Determine the coordinates of the points which are symmetric with 
P(a, b, c) with respect to 

(a) the X-plane; (6) the F-axis; (c) the origin; (d) the Z-axis; 
(e) the FZ-plane; (/) the X-axis; (g) the XF-plane. 

4. Develop one or more algebraic conditions which are satisfied by the 
coordinates of the points which lie 

(a) on a sphere of radius 4 which has its center at the origin; 

(b) on a plane which cuts the F-axis perpendicularly at a point 5 units 

behind the origin; 

(c) on a line parallel to the Z-axis and through the point A (3, 4, 1); 

(d) on a plane parallel to the F-plane and passing through the point 

4(-l, 2, 1); 

(e) on a circle in the Z-X-plane whose center is at the point C(0, 0, 4) 

and whose radius is 3; 
(/) on a line perpendicular to the XZ-pl&ne and through the point 

4(-2, -3, -1); 
(g) on the line determined by the origin and the point A (2, 1, 1). 

6. The point P lies on a line through the origin which makes with the posi- 
tive directions on the coordinate axes angles whose cosines are equal to J, 

i, and ^~, and the distance OP is equal to 3. Determine the coordinates of P. 



TWO POINTS 



53 



6. Prqye that the line from to A(l, 1, 1) makes equal angles with the posi- 
tive directions on the three coordinate axes. 

7. Prove that the lines which join the origin to P(a, 6, c) and to Q(ka, kb, kc) 
make equal or supplementary angles with the positive directions on the 
coordinate axes. 

8. Prove that the sum of the squares of the cosines of the angles which the 
line OP makes with the positive directions on the coordinate axes is equal to 1. 

32. Two Points. The c.p. of the point P is a rectangular 
parallelepiped whose faces are parallel to, or lie in, the coordinate 
planes and of which the origin and P are opposite vertices, while 
the other vertices all lie in the coordinate planes. We construct 
now a rectangular parallelepiped whose faces are parallel to, or 
lie in, the coordinate planes, but of which two opposite vertices 
are to be two arbitrarily assigned points P and Q. The constrvc- 
tion, of which the result is shown in Fig. 2, is carried out most 
readily as follows: 



B 



B 



1 



-- 



Q 1 



A 1 



Q 



Fia. 2 



Through the points P xy and Q xy we draw lines parallel to the 
X- and F-axes, so as to form the rectangle P xy A xy Q xy B xy ] through 
the four vertices of this rectangle we draw lines p z , a z , q t , and 6 a 
parallel to the Z-axis which are the vertical boundaries of the side 
faces of the parallelepiped. To complete the construction, we 
draw through P a line parallel to the X-axis, meeting a s in A; 
through A a line parallel to the 7-axis, meeting q z in Q'; through 
Q' a line parallel to the X-axis, meeting 6, in 5; and through B 



54 POINTS AND LINES 

a line parallel to the F-axis, meeting p z in 1 J . Starting from Q, 
we locate in a similar manner the vertices #', P' 9 and A' of the 
parallelepiped. 

Remark 1. The parallelepiped whose construction is described 
above will be called the coordinate parallelepiped (c.p.) of P and 
Q. It is important that the reader develop skill in carrying out 
this construction; a number of valuable results can be obtained 
readily by means of it. 

Remark 2. The construction of the c.p. of two points P and Q 
can be carried out equally well if we start with the points P yz 
and Qysy or with the points P& and Q&. 

Since the line P'A f is parallel to the X-axis, the segment P'A r 
is equal to the segment P X Q X of the X-axis determined by the 
projections of P and Q on the X-axis. Since 0, P x , and Q x are 
points on the same directed line, namely, the X-axis, we know 
moreover that 

OP X + P X Q X + Q X = 
that is, 

XP + P X Q X -x Q = 0, 

in virtue of Theorem 1, Section 30, page 51. Hence we conclude 

that 

Proj*PQ = P X Q X = x Q - x P . 

Leaving the proof of similar formulas for the projections of the 
segment PQ on the Y- and Z-axes to the reader, we state the fol- 
lowing theorem. 

THEOREM 4. The projections on the coordinate axes of a directed 
segment of a straight line are equal to the differences between the 
corresponding coordinates of the end point and those of the initial 
point of the segment. 

We observe now that the twelve edges of the c.p. of P and Q 
have lengths equal, four by four, to the numerical value of the 
differences between the coordinates of P and Q. If we make use 
once more of the property of the body diagonal of a rectangular 
parallelepiped which was brought forward in Section 30, we 
obtain the following extension of Theorem 2. 

THEOREM 5. The square of the distance PQ is equal to the sum of 
the squares of the differences of the coordinates of P and of Q 9 that Is, 

PQ 2 - (X Q - x p )* + (y Q - y p ) 2 + (M Q - * P )*. 



DIRECTION COSINES OF A LINE 55 

33. Direction Cosines of a Line. We recall that the angle 
between two lines / and m which do not lie in the same plane is 
defined as the angle between any two concurrent lines of which one 
is parallel to / and the other to m. This extension of the concept 
" angle between two lines" enables us to speak of the angles which 
an arbitrary line makes with the coordinate axes and gives sig- 
nificance therefore to the following definition. 

DEFINITION II. The direction angles of a directed line are the angles 
between -180 and +180 which the directed line makes with the 
positive directions of the coordinate axes. The direction cosines of a 
directed line are the cosines of its direction angles. 

Notation. Whenever it is desirable to specify the directed 
line to which reference is made, the direction angles of the line 
PQ will be designated by the symbols a pQ , PQ , and y p ^ its 
direction cosines by X M P and v pQ . Similar notations will be 
used for the direction angles and the direction cosines of an un- 
directed line /. When it is not essential to specify the line, the 
subscripts will be omitted. 

On the basis of this definition we obtain, from a consideration 
of the c.p. of P and Q and by using some of the properties of the 
rectangular parallelepiped mentioned in the second paragraph of 
Section 30 (page 51), the following extension of Theorem 3. 

THEOREM 6. The direction cosines of the directed line PQ are equal 
to the quotients of the differences between the coordinates of Q and 
those of P by the distance PQ ; that is, ! ^ 



= COS a pQ - 



X Q ~ Xp 


V(X Q - * 


p) 2 + (VQ ~ Jp) 2 + (*Q - /) 2 



PQ - COS 7pQ 



Remark. The square root here, as in Theorem 3, is to be taken 
with the positive sign; a change of sign in the square root leads 
to the direction cosines of the line QP. In the case of an undirected 
line I, the sign of the square root remains ambiguous, the two 
signs corresponding to the two directions which may be specified 



56 POINTS AND LINES 

upon the line. The formulas of Theorem 6 lead to the following 
very useful corollaries. 

COROLLARY l. The direction cosines of an undirected line are pro- 
portional to the differences between the coordinates of any two of its 
points. 

COROLLARY 2. The direction cosines of an undirected line through 
the origin are proportional to the coordinates of any one of its points. 

COROLLARY 3. The coordinates of any point on the line which Joins 
the origin to a point P are proportional to the coordinates of P. 

(Compare Exercise 7, Section 31, p. 53.) 

Furthermore we obtain the following important result. 

THEOREM 7. The sum of the squares of the direction cosines of a line 
is equal to 1. 

Proof. If we choose any two points P and Q on the line and 
express the direction cosines of the line in terms of the coordinates 
of P and of Q, as in Theorem 6, we find that 



(XQ- XP)* + (y Q - VPY + (ZQ - ZP)* ' 

Remark. The formula proved in Theorem 7 will be used re- 
peatedly in the sequel. We shall use the phrase "admissible values 
of X, JUL, v" to indicate values of these variables for which X 2 + pt 2 
+ i> 2 = 1. Its principal use will be to enable us to determine the 
direction cosines of a line if we merely know numbers to which 
they are proportional. This will in most cases relieve us of the 
necessity of actually determining the direction cosines and make 
it possible to operate with their ratios. A special case of Theorem 
7 is contained in Exercise 8, Section 31, (page 53). 

THEOREM 8. If the direction cosines of an undirected line are pro- 
portional to three given numbers, then their actual values are equal 
to these numbers, each divided by the square root of the sum of their 
squares. 

For, if X = ka, /* = kb and v = kc, then it follows from Theo- 

rem 7 that fc 2 (a 2 + 6 2 + c 2 ) = 1, so that k = , 1 

Vtt 2 + b 2 + c 2 

The ambiguity of sign in the square root corresponds to the possi- 
bility of two directions on the undirected line. 



THREE COLLINEAR POINTS 57 

34. Three Collinear Points. If A, B and P are points on a line, 
the direction cosines of the segments AP and AB are either equal 
or else equal numerically but opposite in sign, thus : 

\4P = X AB> HAP = f *AB> V AP = V AB> 

where the upper signs are to be used if the segments AP and AB 
have the same direction and the lower signs if they have opposite 
directions. And, if the points A, B and P are not collinear, not 
all of these relations can hold. By means of Theorem 6 we derive 
from these relations the following equations: 

XP - x A = X B ~ x A VP ~ VA 'UB ~ VA 



AB ' A~P AB 

ZP - Z A = Z R - Z A m 

AP AB ' 

or _ 

XP - X A = HP ~ HA = ZP ~ ZA = AP 
XB ~ XA 1/B ~ VA Z B ~ ZA ~AB 

Moreover, if the segments AP and A B have the same direction 
we can take AP = +AP and AB = -\-AB; whereas if they have 
opposite directions, we can take AP = +AP and AB AB. 

AP AP AP 

It follows therefore that we have-r-^ = + =or = according 

At> AB AB 

as the segments AP and AB have the same or opposite directions. 
We obtain therefore the following important result. 

THEOREM 9. The necessary and sufficient condition that the point 
P(** y s) shall lie on the line through A and B is that the coordinates 
x, y, s must satisfy the linear equations 



*B-*A ys-yA S B-*A AB 

Remark 1. It should be clear that by starting with the seg- 
ments BP and BA we find the equations 

x - X B = y - y B = z - Z B ^ BP m 
XA - X B VA ~ VB Z A - Z B BA ' 

and if we use the segments AP and PB, the resulting equations are 

x - XA = y ~ VA = * - ZA = AP ^ 
x B - x VB - V Z B - z PB' 



58 POINTS AND LINES 

Remark 2. The formulas established in Theorem 9, and also 
those given in Remark 1, enable us to determine the coordinates 
of the point P on the line AB as soon as the coordinates of the 

AP 
points A and B are known and also the ratio - of the segments 



AP and PB in which P divides the segment AB. For, if AP : PB 
= n : r 2 , then AP : AB = r\ : n + r 2 , and BP : BA = r 2 : r\ + r 2 . 
Hence we have 

x XA r\ x - X A n x - X B r 2 



XB ~ x r* X B - X A n + r 2 ' X A - X B n + r 2 
From either of these equations we find for the ^-coordinate of P: 



_ 
Xp ~ 



and similar results are obtained for yp and zp. We have therefore 
the following further result. 

COROLLARY. If the points A, B, and P are in a straight line and if 
AP : PB = ri : r 2 , then the coordinates x, y, s of P are given by the 
formulas: 

r& A -f r& B = r*y A + r,y g ^ r 2 s^ + r lgg 

~" n + ^2 ' y n + r a ' ri -f r 2 

Remark 3. The formulas of this corollary are very useful for 
later developments. But they hide to some extent the simple 
geometrical fact from which they have been derived and which 
finds more direct expression in the formulas of Theorem 9 and 
Remark 1. It is advisable therefore, particularly in the begin- 
ning and in numerical problems, to go back to these earlier formulas 
rather than merely to substitute in the formulas of the corollary. 

Remark 4. The first three terms in the equations of Theorem 
9 give two independent linear equations which the coordinates of 
any point on AB must satisfy; the same statement may be made 
for the equations in Remark 1. If we consider also the last term 
in each of these cases and here the last set of equations in 
Remark 1 is particularly useful they give us three equations 
which express the coordinates of an arbitrary point P on the line 
in terms of the coordinates of A and B, and of one parameter, or 

auxiliary variable, namely, r = - , that is, in terms of the ratio r 

7*2 



THREE COLLINEAR POINTS 59 

of the segments into which P divides the segment AB. This pa- 
rameter r varies as the point P moves along the line AB. When 
P coincides with A, r = 0; as P moves from A to B r increases. 
When P lies outside the segment A B, either on the side of A or 
on the side of B, r is negative; and as P moves off indefinitely along 
the line in either direction, r tends towards 1. For 

AP AB + BP 



= _ 
PB ~~ PB ' 

since AB is fixed and PB increases numerically when P moves off 
along the line, the first term on the right tends toward zero and 
hence r tends toward 1. The reader will find it worth while to 
make clear to himself in detail the manner in which the parameter 
r varies as P occupies various positions on the line AB. 

If we combine Theorems 6 and 9, we are led to the following 
valuable theorem. 

THEOREM 10. The necessary and sufficient condition that a point 
P(x 9 y 9 z) shall lie on the directed line through A whose direction cosines 
are A, /u, and v is that the coordinates x 9 y, and s must satisfy the 
equations 



For if P lies on the specified line and if B is another arbitrary, 
but fixed, point on the line, then the equations of Theorem 9 hold. 
But, in accordance with Theorem 6, we have 

y B - y A = /z AB, Z B - Z A = v AB. 



If these values are substituted in the equations of Theorem 9, 
the desired result is obtained. 

COROLLARY 1. The necessary and sufficient condition that a point 
P(x 9 y, s) shall lie on the undirected line through the point A whose 
direction cosines are proportional to 1 9 m and n is that the coordinates 
x 9 y and s shall satisfy the equations 

*-*A = y-y* = * ~*A = AP 

I m n Vl* + m 2 + n 2 

This corollary follows from Theorem 10 in combination with 
Theorem 8. The ambiguity in the sign of the square root corre- 
sponds to the possibility of two directions on the line of which only 
the ratios of the direction cosines are given. 



60 POINTS AND LINES 

The observation made in Remark 4 (page 58) leads from Theo- 
rem 10 and Corollary 1 to two further results of importance. 

COROLLARY 2. The coordinates of a point P(x 9 y, a) on the line 
through A whose direction cosines are equal to X, M v are given by the 
equations 

x = X A -f Xs, y = y A + us, z = Z A -f- vs, 

where s designates the signed measure of the directed segment AP\ 
conversely, any point P whose coordinates are equal to these expres- 
sions lies on the specified line at a directed distance from A equal to s. 
COROLLARY 3. The coordinates of a point P(x 9 y, s) on the undirected 
line through A whose direction cosines are proportional to /, m, n are 
given by the equations 



*=* A + lt 9 y=y A + mt, * - ^ + iu, where t = ^ + = ; 

conversely, any point P whose coordinates are given by these ex- 
pressions lies on the specified line at a distance from A equal to 

t V> -f m 2 -f rt 2 . 

Remark. The equations of Corollary 1 are frequently referred 
to as the "symmetric equations of the line 7 '; those of Corollaries 
2 and 3 as the " parametric equations of the line/' The variable 
s, or t, which changes as P moves along the line, is the parameter. 
The terminology "equations of a line" will be more fully justified 
in Chapter IV (see Section 47, page 83). 

The parametric equations of the line are used a great deal in the 
sequel. The reader is urged to master thoroughly the methods 
by which they have been obtained. It should be observed more- 
over that the equations stated in the Corollary of Theorem 9 are 
also parametric equations of the line, the parameter being the 

ratio r (see Remark 4, page 58). 
i 

35. Exercises. 

1. Construct the coordinate parallelepipeds of the following pairs of points, 
and determine their distances and the direction cosines of the lines joining 
them: 

(a) 4(5, 2, -1) and B(-3, -4, 2); (6) 4(2, 4, 5) and 5(7, 1, 1); 
(c) 4(2, 3, 4) and B(5, -2, 7); (d) 4(3, -2, -1) and B(-3, 4, 5); 
(e)4(-4,3,5)and(-4, -2,0); (/)4(3,4,5)andB(-3, -4, -5); 
(g) 4(0, 3, 6) and J5(4, -1, 6); (h) 4 (-2, 3, 5) and B(-2, 3, -1). 



THE ANGLE BETWEEN TWO LINES 61 

I. Determine which of the following sets of points are collinear: 

(a) A(3, -1, 4), B(-2, 4, -1), (7(1, 1, 2); (6) A(Q, 0, 0), B(2, 5, -3), 
C(4, 10, -6); (c) A(l, -2, 3), B(-l, 2, -3), C(-3, 5, 0); 
(d) A(-2, 2, 3), B(l, -1, 0), C<7, -7, -3); () ^(0, -3, 1), 
B(4, -2, -1), C(2, -4, 3); (/) 4(5, 2, 7), B(l, 5,5), (-3,8,3). 

\Z. Determine the coordinates of the point at which the segment from 
A(3, 2, 5) to B(5, 4, 2) is bisected; also the coordinates of the points at 
which this segment is trisected. 

& The line of the preceding problem is extended beyond B to a point C 
such that (a) BC = AB; (6) BC = 3 AB; (c) BC = \ AB. Determine the 
coordinates of C in each case. 

*0. Find the coordinates of the center of mass of the homogeneous triangle 
whose vertices are A(-2, 5, 4), (3, -1, -2) and C(8, -7, 4). (The center 
of mass of a homogeneous triangle is the point of intersection of the medians.) 

6. Determine the coordinates of the point in which the side AC of the tri- 
angle of Exercise 5 is met by the bisectors of the interior and the exterior 
angles at B. Find also the direction cosines of the bisectors. 

7. Show that the points A(-3, 2, 5), B(l, 0, 1) and C(ll, -5, -9) are 
collinear and determine the ratio of the segments AC : CB. 

8. On a line through the point .4(5, 4, 2) whose direction cosines are pro- 
portional to 2, 1 and 2, a point B is determined such that AB = 4. 
Find the coordinates of B. How many positions arc possible for B? 

9. Determine the center of mass of the homogeneous triangle whose vertices 
are at Pi(xi, yi, z), * = 1, 2, 3. 

10. Show that the three lines which join the midpoints of the three pairs of 
opposite edges of the tetrahedron PiP 2 PsP4 have a common midpoint. This 
common midpoint is called the center of mass of the homogeneous tetrahedron 
PiP 2 PaP4. (Opposite edges of a tetrahedron are edges which have no point 
in common.) 

II. Prove that any vertex of a homogeneous tetrahedron, the center of 
mass of the opposite face and the center of mass of the tetrahedron are col- 
linear; determine the ratio of the segments in which the center of mass of the 
tetrahedron divides the segment determined by the other two points. 

12. Show that if and only if P(x, y, z) lies on the sphere of radius r, whose 
center is at (7(a, 6, c), the coordinates x, y, z satisfy the equation 

(x - a) 2 + (y - 6) 2 + (z - c)* = r 2 . 

13. Determine the equation satisfied by the coordinates of all points which 
are at the same distance from A ( 2, 1, 3) as from B(4, 2, 0). 

14. Determine the equation satisfied by the coordinates of all points on the 
surface of the sphere whose center is at C(3, 2, 3) and whose radius is 5. 

15. Determine the equation satisfied by the coordinates of all points whose 
distance from A(l, 3, 4) is twice as great as their distance from B( 2, 0, 2). 

36. The Angle Between Two Lines. The Projection Method. 

To calculate the angle between two lines whose direction cosines 



62 



POINTS AND LINES 



are known, we shall use a method which will find frequent applica- 
tion in the sequel and which is based on the following two theorems. 

THEOREM 11. The projection of a segment AB of a directed line I 
upon a directed line m is equal to the product of AB by the cosine of the 
angle between the two directed lines. 

Proof. We distinguish two cases, according as the lines I and 
m do or do not lie in one plane. 

(a) For the case when the lines I and m lie in one plane, the 
proof can be found in most books on Plane Analytical Geometry 
and in some books on Trigonometry.* For this reason we shall 

not repeat the proof here. 

(6) If the lines I and m are 
not coplanar, let the angle be- 
tween them be 6. We construct 
a line m r through the point A on 
Z, parallel to m (see Fig. 3); the 
angle between m' and I will then 
also be equal to 6. Through A 
and B we construct planes per- 
pendicular to m; these planes 
will then also be perpendicular 
to m'. If the points in which 

these planes meet m and m' are C, D and A, D f respectively, then 
CD = AD' (Why?). From these facts, in combination with part 
(a) of this proof, we conclude that 

Proj m AB = CD = AD' = AB cos 0. 

This proves our theorem. 

THEOREM 12. The sum of the projections upon a directed line m 
of the segments of a closed broken line in space is equal to zero. 

Proof. If the vertices of the broken line in space are A, B, C, 
. . . , P and if their projections on the directed line m are .4', B', C', 
. . . , P', we have to show that A'B' + B'C' + . . . + P'A' 
= 0. That this is indeed the case follows from a fundamental 
theorem, of which a proof is found in books on Plane Analytical 
Geometry and on Trigonometry,! according to which the sum of 

* See, for example, the author's Plane Trigonometry, p. 36. 
t See, for example, the author's Plane Trigonometry, p. 4. 




FIG. 3 



THE ANGLE BETWEEN TWO LINES 



63 



the directed segments of a line, of which the end point of the last 
segment coincides with the initial point of the first segment, is 
equal to zero. We conclude therefore that 

Pro} m AB + Pro] m BC + . . . + Pro] m PA = A'B' + B'C' 
+ . . . + P'A' = 0. 

We proceed now to the determination of the angle 6 between 
two directed lines I and m. We construct first the c.p. of two 
points A and B selected arbitrarily on one of the lines (see Fig. 4 
in which A and B are taken 
on Z); and we apply Theorem 
12 to the closed broken line 
ABDCA. Thus we obtain the 
equation 

Pro'] m AB + Proj w Z) + 
Pro] m DC + Pro] m CA = 0. 

The segment AB lies on I, the 

segments BD y DC and CA on 

lines which are parallel to the 

Z-j Y-, and X-axes; the angles 

which these lines make with the 

line m are respectively equal to 0, y mj p m , and a m , the last three 

angles being the direction angles of m. The last equation written 

above leads therefore by means of Theorem 11 to the statement 

that 

AB cos 6 + BD v m + DC /% + CA X m = 0. 

Moreover BD, DC and CA are equal to the projections of BA on 
the Z-, F-, and X-axes respectively, so that it follows from Theorem 
11 that 

CA = BA 




Fio. 4 



BD = BA 



DC = BA 



and 



If we substitute these values in the preceding equation and re- 
member that BA = AB ^ 0, we obtain the following result: 



THEOREM 13. The cosine of the angle between two directed lines is 
equal to the sum of the products of their corresponding direction 
cosines. 



64 



POINTS AND LINES 



Remark. If we restrict angles between directed lines to lie be- 
tween 180 and +180, the result contained in Theorem 13 
determines the magnitude of angle 6 but leaves the sign of this 
angle ambiguous. This ambiguity corresponds to the fact that 
either of the two lines may be taken as the initial side of the angle. 

COROLLARY 1. The necessary and sufficient condition that the lines 
I and m arc perpendicular is that \i\ m + ^ m _j_ vp m = o. 

COROLLARY 2. If the direction cosines of two undirected lines are 
proportional to fi, mi, n t and 1 2 , m 2 , n 2 , the necessary and sufficient 
condition for the perpendicularity of the lines is that lih + 
= 0. 



By means of Theorem 13 we can determine the numerical values 
of all the trigonometric ratios of the angles between two directed 
lines. On account of its special interest we develop a formula for 
the sine of these angles. This is done most conveniently by 
means of the following auxiliary theorem, which is of some interest 
on its own account. 

LEMMA. The following identity holds between any six numbers, 
real or complex, a, 6, c and 01, 61, ci: 



(a 2 + 



+ c 2 ) 

b 



+ *>i 2 + ci 2 ) - (aa, 



c o I 2 
* ml" 



f-Wn + < 
& 
61 



The proof of this formula requires merely that the indicated 
operations on both sides of the equation shall be carried out; the 
identity of the two sides will then at once become apparent. 

If this lemma is applied to the direction cosines X /; jjL lf v l and 
\ f Mm, VM of the lines I and m, we find that 

(A/ 2 + M/ 2 + "/ 2 ) (X m 2 + ^ + O 



Mm 



v l 



-h 



X, 
X*, 



Mm 



But in view of Theorems 7 and 13 this result leads to the equation 



cos 2 B = 



Mm v m 



M/ 
Mm 



THEOREM 14. The square of the sine of the angle between two di- 
rected lines is equal to the sum of the squares of the two-rowed minors 
which can be formed from the matrix constituted by the two sets of 



THE ANGLE BETWEEN TWO LINES 65 



direction cosines of the lines; that is, 

2 



sin 2 = 



M/ 



M m "... 

87. Exercises. 

1. Determine the cosines of the angles between the lines whose direction 
cosines Xi, MI, v\ and X 2 , M2, "2 are given by the following data: 

(a) Xi : MI : "i = -2 : 1 : 2, X 2 : M2 : "2 = 2 : 6 : -3 

(6) Xi : MI : "i = 2 : 3 : 6, X 2 : M2 : i* = 3 : 14 : 18 

(c) Xi : MI : ?i = 3 : -2 : -6, X 2 : M2 : z = 1 : - 2 : -2 

(d) X! : MI : "i = 4 : : 3, X 2 : M2 : "2 = 1 : 2 : 3 

2. Determine the cosines of the angles formed by the sides of the triangle 
whose vertices are the points A (3, 1, 4), B( 2, 4, 1) and (7(1, 1, 2). 

3. Test for perpendicularity the pairs of lines whose direction cosines are 
given by the following data: 

(a) Xi : MI : *i = -2 : 1 : 2, X 2 : M2 : ^2 = 1 : -2 : 2 
(6) Xi : MI : "i = 3: - 1 : 2, X 2 : M2 : *> 2 = 1 : 1 : -1 
(c) Xi : MI : v\ = 2 : -3 : 6, X 2 : M2 : "2 = 1 : 2 : 

4. Develop a formula for the tangent, the cotangent, the secant and the 
cosecant of the angle between two lines. 

5. Determine the direction cosines of a line which is perpendicular to the 
two lines whose direction cosines are proportional to 3 : 2:4 and 1 : 3 : 2. 

6. Determine the direction cosines of a line which makes equal angles with 
the three lines whose direction cosines are proportional to 1, 4, 8; to 8, 1, 
4; and to 1, 2, -2. 

38. Miscellaneous Exercises. 

1. Determine the conditions which the coordinates of a point must satisfy 
in order to be equally distant from the points A (2, 1, 0), B( 3, 2, 1) and 
C(l, 3, -2). 

2. Solve the same problem for the points Pi(x{, yi, 2;), i = 1, 2, 3. 

3. Determine a point which is equally distant from the four points A (2, 1, 3), 
B(l, 1, 2), C(2, 0, 5), and D(2, 0, 3). 

4. Establish the condition on the coordinates of the four points Pi(xi, yi, zi), 
i = 1, 2, 3, 4, necessary and sufficient for the existence of a single point that 
is equally distant from these four points. 

5. Set up the equation which is satisfied by the coordinates of any point 
on the sphere which passes through the four points A (3, -2, 4), B(2, -3, 2), 
C(4, -2, 2) and D(5, -1,3). 

6. Determine the direction cosines of a line perpendicular to the two lines 
whose direction cosines are X,-, M, "*> i = 1, 2. Does this problem always have 
a solution? Does it ever have more than one solution? 

7. Remembering that a line is perpendicular to a plane if it is perpendicular 
to two lines in the plane, determine the direction cosines of a line which is 



66 POINTS AND LINES 

perpendicular to the plane of the triangle whose vertices are the points 
4(3, -2, 4), (4, 0, 2) and C(0, -4, -2). 

8. Determine the condition which must be satisfied by the direction cosines 
X,', /if, i/;, i 1, 2, 3 of three lines in order that there may exist one or more 
lines perpendicular to these lines. 

9. Show that if the coordinates of the three points Pi(xi, yi, zi),i = 1, 2, 3 
satisfy an equation of the form ax + by + cz = 0, in which a, 6, c are not all 
zero, then there exists at least one line which is perpendicular to the three 
lines OP,, i = 1, 2, 3. 

10. Determine the direction cosines of a line which is to make equal angles 
with the three lines whose direction cosines are X,-, m, vi, i = 1, 2, 3. Does 
this problem always have a solution? Can it have more than one solution? 

11. Determine the direction cosines of a line which makes equal angles with 
the three lines connecting the origin with the points 4 (2, 3, 6), 5(1, 8, 4), 
and C( 1,2, -2). 

12. The line which joins 4(1, 3, 5) to B(15, -15, 8) is produced beyond B 
to a point P such that BP 5. Determine the coordinates of P. 

13. Prove that every point whose coordinates satisfy the equation 

z 2 + ?/ 2 + z 2 + 2 ax + 2 by + 2 cz + d = 

lies on the surface of a sphere. Determine the center and the radius of this 
sphere. 

14. Determine the center and the radius of each of the spheres represented 
by the following equations: 

(6) x 2 -f t/ 2 + z 2 + 6 x - 2 y + 4 z - 2 = 

(c) x 2 + y 2 -f z 2 -f 2 x - 6 y -f 8 z -f 26 = 

(d) x 2 + ?/ 2 + z 2 - 8 x -f 4 y -\- 6 z - 33 =0 

16. Two lines mi and m* meet at a point under an angle 6, = 180. A line 
l f not necessarily in the same plane with mi and m 2 , makes angles cti and 2 
with mi and m z respectively and an angle with a line in the plane of mi and 
ma, perpendicular to m iy and on the same side of mi as w 2 . Prove that cos ft 

C080 

cos i cos a 2 

of a frame of reference, the line m l as X-axis, and the pljfllHHand m 2 as 
XF-plane. ^^^H 

16. If, in the configuration of the preceding exercise, y I^Wmgle which 
I makes with a line perpendicular to the plane of mi and m%, then cos y sin 
= =fc [1 cos 2 i cos 2 a 2 cos 2 d + 2 cos i cos 2 cos 0]* 



sin d = cos 2 ~ cos i cos 6 = 



Hint: Take as the origin 



= =fc 



1 COS 6 COS ai 
COS 1 COS 2 
COS <Xi COS 2 1 



the + or sign is to be used according as I and the perpendicular to the plane 
of mi and w 2 point to the same side or to opposite sides of this plane. 



CHAPTER IV 



PLANES AND LINES 

39. Surfaces and Curves. If a point P is to be chosen at 
random in space, its determination will, in the coordinate system 
which we have used thus far, depend upon three independent 
choices of arbitrary real numbers, namely, of an x-, a ?/-, and a 
^-coordinate, each of which can be selected without consideration 
of the selections made for the other two. This fact is expressed 
in the statement that a point in space has three degrees of freedom. 

If a point is to be chosen at random on a surface such as we are 
likely to meet in ordinary experience (we may think here of sur- 
faces which limit the objects in our environment, such as tables, 
bottles, lamp shades, trees, etc.) 
the determination will depend 
upon two independent choices 
of arbitrary real numbers; for, 
after two coordinates have been 
chosen at random, the x- and y- 
coordinates for example, the third 
one, the 2-coordinate, must be so 
chosen as to furnish a point on 
the surface (see Fig. 5) ; and this 
will leave, at least in the case of 
surfaces of ordinary experience, 
in general a choice among a finite 
number ofvalues. For this rea- 
son a poiiltt|ferface is said to have two degrees of freedom. 

And if ^^Hlps to be chosen at random on a curve (again our 
reference I51n the first place to curves of cornmOn experience) 
only one coordinate can be selected arbitrarily; a point on a curve 
is therefore said to have one degree of freedom. 

It should be clear from this discussion that to restrict a point in 
three-space to a surface, we have to impose one condition on its 
coordinates; and to limit a point in three-space to a curve, we 
shall have to subject its coordinates to two independent conditions. 

67 




KNJ. 



68 PLANES AND LINES 

These considerations, vague and inconclusive though they are, 
suffice perhaps to indicate that there is some justification for the 
following definitions. 

DEFINITION I. A surface is the totality of all points in three-space 
whose Cartesian coordinates satisfy one equation. 

Remark. The equation which is thus laid down by definition 
as the algebraic counterpart of the surface, is called the equation 
of the surface ; and the surface is referred to as the locus of the 
equation. 

DEFINITION II. A curve is the totality of all those points in three- 
space whose Cartesian coordinates satisfy two independent equations* 

We shall speak of the "equations of a curve" and of the "locus 
of a pair of equations/' 

Remark. These definitions do not specify sharply the con- 
cepts "surface" and "curve" because the word "equation" used 
in them is left without specification. According as the class of 
equations that is taken into consideration is widened or narrowed, 
we shall presumably enlarge or restrict the concepts "surface" 
and "curve." The remarks preceding the definitions are intended 
to make clear that the surfaces and curves of ordinary experience 
are included among the "surfaces" and "curves" introduced by 
the definitions. 

The surfaces and curves which are the loci of algebraic equa- 
tions and pairs of algebraic equations, respectively, are called 
"algebraic surfaces" and "algebraic curves." In this book we 
shall be concerned exclusively with surfaces (and curves) which 
are loci of the simplest types of algebraic equations in three vari- 
ables (pairs of such equations), namely, of algebraic equations of 
the first and second degrees. We shall, however, have to deal 
occasionally with loci of equations of a more general character; 
and we shall begin with some considerations of a general nature. 

40. Cylindrical Surfaces. Systems of Planes. What can be 
said about the space locus of an equation like x* + y z = 4, in 
which only two of the variables are present? The plane locus of 
such an equation consists of the points on the X F-plane, whose 
coordinates satisfy the given equation; it is therefore a plane 
curve. Since the equation does not restrict the ^-coordinate, a 



CYLINDRICAL SURFACES 69 

point P(a y 6, c) will belong to its locus if and only if the point 
P xy (a, bj 0) does, that is, if and only if the projection of P on the 
-XT-plane belongs to the plane locus of the equation. Conse- 
quently the space locus of this equation will be generated by a 
line which moves parallel to the Z-axis and which passes succes- 
sively through the points of the plane locus of the equation. 
We introduce now the following terminology: 

DEFINITION III. A cylindrical surface is a surface generated by a 
line which moves in such a way as to be always parallel to a fixed 
line and in such a way as to pass through the points on a fixed plane 
curve. Any position of the movfng line is called a generating line 
(generatrix); the fixed plane curve is called the directrix. If the 
generatrix is perpendicular to the plane of the directrix, we have a 
right cylindrical surface; if not, an oblique cylindrical surface. 

By the aid of this terminology we can express the result of the 
foregoing discussion in the following form. 

THEOREM 1. The locus of the equation /(*, y) = is a right cylin- 
drical surface whose generating line is parallel to the Z-axis and whose 
directrix is the plane locus of the equation. 

Remark 1. Similar theorems hold of course concerning the loci 
of equations from which the variable x or the variable y is absent. 
It will be good practice for the reader to present in full the argu- 
ment for these cases. The locus of the particular equation 
#2 _|_ y2 _- 4 j s a right circular cylindrical surface, whose gener- 
ating line is parallel to the Z-axis and whose directrix is a circle in 
the XF-plane with center at the origin and radius equal to 2. 

Remark 2. By means of this theorem the whole field of Plane 
Analytical Geometry is seen to be a province in the domain of 
Solid Analytical Geometry. For it shows that the determination 
of the space loci of equations in two Cartesian variables depends 
upon the determination of the plane loci of such equations. 

In particular the space locus of a linear equation in two variables, 
for example, of the equation ay + bz + c = 0, is a cylindrical 
surface of which the directrix is a straight line and the generatrix 
parallel to the Z-axis; the locus of this equation is therefore a 
plane parallel to the X-axis. 

If an equation contains only one Cartesian variable, its plane 
locus is a set of lines, real or complex, parallel to one of the coordi- 



70 



PLANES AND LINES 



nate axes. Its space locus is therefore a set u* planes, real or 
complex, parallel to one of the coordinate planes. For instance, 
the locus of the equation # 2 7# + 12 = consists of two planes 
parallel to the FZ-plane, at distances of 3 and 4 units from the 
FZ-plane. 

We recall that the geometric statement "a point P belongs to a 
certain locus" has as its algebraic equivalent "the coordinates of 
P satisfy the equation of a locus/' With this fact in mind, we 
turn to a particularly interesting case of an equation in two vari- 
ables, namely, the equation obtained by eliminating one of the 
variables from two equations in three variables. Suppose that 
we eliminate z from the equations f(x, y, z) = and g(x, y, z) = 0, 
and that the result of the elimination is the equation F(x, y) = 0. 
What can be said about the loci of these three equations? 

In the first place it follows from Theorem 1 that the locus of 
the equation F(x, y) = is a cylindrical surface parallel to the 
Z-axis. On the other hand, since it is satisfied (provided the 
elimination has been carried out correctly) by all values of x, y, 
and z which satisfy the equations /(x, ?/, z) = and g(x } y, z) = 0, 

it must pass through all tho 
points common to the two 
surfaces which these two 
equations represent, that is, 
through their curve of in- 
tersection (see Definitions 
II and I). The locus of 
the equation F(x, y) = is 
therefore the cylindrical sur- 
face which projects upon the 
-XT-plane the curve repre- 
sented by the equations 

f(x, y, z) = and g(x, y, z) = 0, see Fig. 6. Moreover the plane 
locus of the equation F(x, y) = 0, consisting of the points in the 
-XT-plane whose coordinates satisfy the equation, is clearly the 
projection of this curve on the .XT-plane. A similar statement 
can be made concerning the equation obtained by eliminating x or 
y from the two given equations. We have obtained therefore the 
following theorem. 




THE LINEAR EQUATION 71 

THEOREM 2. The equation in two variables, obtained by eliminating 
one variable from two equations in three variables, has as its space 
locus the cylindrical surface which projects the curve represented by 
the two equations upon the plane of the two remaining variables; 
and as its plane locus the projection of the curve on the same coordi- 
nate plane. 

41. The Linear Equation ax + by + cz + d = 0. 

DEFINITION IV. A plane is a set of points of such character that if 
any two points A and B belong to the set, then every point on the line 
Joining AB also belongs to it. 

On the basis of this definition it is a simple matter to prove the 
following theorem. 

THEOREM 3. The locus of any equation of the first degree in *, y, 
and a is a plane. 

Proof. The most general equation of the first degree in x, y, and 
z is ax + by + cz + d = 0. If an arbitrary point P is taken on 
the line AB, its coordinates will be, according to the corollary to 
Theorem 9, Chapter II (see Section 34, page 58) 



_ _ r 2 y A 

XP ~ ' yp ~ ' ** 



rt+r, ' ~ r, + r 2 ' " r, + r, ' 

when TI : r 2 = AP : PB. We have to show therefore that the 
identities ax A + by A + CZ A + d = and ax B + by B + CZ B + d s= 
have as a consequence the identity 



, ^ A s , A i B , d s 

"~ 



r 2 n r 2 



If we write this equation in the equivalent form obtained by clear- 
ing it of fractions (n + r 2 3p 0) and collecting the terms in r\ and 
those in r 2 , namely, 



by A + CZ A + d) + ri(ax B + by B + CZ B + d) ss 0, 

it should be indeed evident that it results from the two given iden- 
tities. The theorem has therefore been proved. 
Next we shall establish its converse. 

THEOREM 4. The equation of any plane is a linear equation in x, y, 
and 



72 PLANES AND LINES 

Proof. We shall divide the proof into four parts. 

(a) Suppose that the plane is parallel to two of the coordinate 
axes, that is, to one of the coordinate planes; let us say for the sake 
of definiteness that the plane is parallel to the ZF-plane at a dis- 
tance k from it. Then for every point on the plane, and for no 
other points, the x-coordinate is equal to fc. Therefore the equation 
of this plane is x k = 0, which is obviously a linear equation. 

(6) If the plane is parallel to one coordinate axis, for example, 
the Z-axis, let the line in which the plane meets the Jf F-plane, re- 
ferred to the X- and F-axes, have the equation ax + by + c = 0. 
It follows then that the equation of the plane is also ax + by + c 
= 0, see Theorem 1 and Remark 2 following it (Section 40, page 
69) ; this is again a linear equation. 

(c) Suppose that the plane cuts all three axes but does not pass 
through the origin. Let it cut the X-axis in the point P(p, 0, 0), 
the F-axis in the point Q(0, q, 0), and the Z-axis in the point 
72(0, 0, r); then p y q, and r are all three different from zero. And 

3* 77 Z 

the given plane must be the locus of the equation (- ~ + - = 1. 

For this equation is linear; its locus is therefore a plane, by virtue 
of Theorem 3. Moreover, the points P, Q, and R clearly belong to 
this locus since their coordinates manifestly satisfy the equation. 
Hence the given plane and the plane which is the locus of this 
equation have three points in common; therefore they coincide. 

(d) Suppose finally that the plane passes through the origin. 
We select in the plane two points PI(XI, y\, z\) and P 2 (#2, 2/2, 2) not 
collinear with 0; and we seek to determine three numbers a, 6, and 
c, not all zero, such that ax\ + byi + cz\ = and ax 2 + by 2 + cz z 
= 0. Is this possible? Yes, for since the points 0, Pi, and P 2 are 
not collinear, the coordinates x\ 9 y\, z\ are not proportional to the 
coordinates x 2 , y^ 2 (see Corollary 3 to Theorem 6, Chapter III, 
Section 33, page 56 and Exercise 7, Section 31, page 53) and 

therefore not all the two-rowed minors of the matrix l ^ M 

%2 2/2 3> II 

can vanish, that is, the rank of this matrix is 2. But this fact 
enables us to conclude, by means of Theorem 4, Chapter II 
(Section 25, page 41) that we can indeed determine the numbers 
a, 6, and c so as to satisfy the conditions mentioned above. The 
locus of the equation ax + by + cz = has therefore the three 



THE LINEAR EQUATION 73 

points 0, PI, and P% in common with the given plane; since this 
locus is moreover a plane (Theorem 3) it coincides with the given 
plane. 

This completes the proof of the theorem. 

Remark. 1 The directed distances p, q, and r from the origin 
to the points in which a plane cuts the coordinate axes are called 

the intercepts of the plane. The equation - + - + - = 1 which 

p q r 

can be written down as soon as the intercepts of a plane are known, 
is usually referred to as the intercept form of the equation of the 
plane. It is a special case of the equation of the plane in terms of 
the coordinates of three of its points. 

COROLLARY. The intercepts of the plane which Is the locus of the 
equation ax -f by ~h ex -f d = 0, In which neither a, nor 6, nor c are 

equal to zero, are equal to , , , and . 

a o c 

Remark 2. It follows from this corollary that if d is a constant, 
different from zero and a, 6, and c are variables which tend to 
zero, then the intercepts of the plane represented by the equation 
ax + by + cz + d = increase beyond all bounds and hence the 
plane moves farther and farther from the origin the usual 
phrase is " the plane moves to infinity." This is the sense in which 
we are to understand the statement that the equation x + y 
+ z + d = represents the " plane at infinity." A satisfac- 
tory treatment of the question here hinted at belongs in the field 
of Projective Geometry; we shall not undertake such treatment 
in this book. Whenever it becomes desirable to recognize ex- 
plicitly that the plane under discussion is not the "plane at 
infinity" we shall speak of a plane at finite distance. 

THEOREM 5. The plane which passes through the three non-colllnear 

y 
points Pi, P 29 and P 8 has the equation 



y\ 
y* 



Proof. The required equation is linear in x, y, and z, by virtue 
Df Theorem 4. If therefore P(x t y, z) is any fourth point of the 
plane determined by the non-collinear points PI, F 2 , and P 3 , there 
tnust exist four numbers a, 6, c, and d which are not all zero, such 



74 PLANES AND LINES 

that 

ax + by + cz + d = 0, 

axi + byi + czi + d = 0, 

ax 2 + by 2 + cz 2 + d = 0, 
and axz + by* + C2 3 + d = 0. 

This is a system of four linear homogeneous equations in the four 
variables a, fr, c, and d. In order that this system of equations 
may possess a non-trivial solution, it is necessary, in view of 
Theorem 2, Chapter II and its corollary (Section 22, page 38), that 
the value of the coefficient determinant must be zero. Hence the 
coordinates of an arbitrary fourth point of the plane PiP 2 Ps must 
x y z 1 

2/1 2i 1 

y-2, 2 1 

?/3 23 1 



satisfy the equation 



- 0. 



On the other hand this equation is linear in x, y, and z, in view of 
Theorem 12, Chapter I (Section 7, page 13). It is moreover satis- 
fied by the coordinates of the points PI, P 2 , and P 3 , as should be 
evident by use of Theorem 7, Chapter I (Section 5, page 9) ; there- 
fore its locus is the plane determined by PI, P 2 , and P 3 . 

The equation established in this theorem is usually referred to 
as the three-point form of the equation of the plane. 

42. Exercises. 

1. How many equations are required to specify a curve in a space of four 
dimensions? To specify a surface in a four-space? To specify a three-space 
in a four-space? 

2. Formulate a general statement of which the answers to the preceding 
exercise and the statements preceding the definitions in Section 39 are special 
cases. 

3. Determine the loci in three-space of the following equations: 

(a) I _ |! = 1 (6) x * -f 2/2 + Z 2 16 

( c ) 4 z = y i (d) x* - 5 x + 6 = 

(e) 4 * 2 + 6 2/ 2 - 12 = (/) z 3 - 6 z 2 + 11 z - 6 - 

4. Show that if PI, PI, and Pa are collinear points, the equation in Theorem 
5 is satisfied identically, that is, for all values of x, y, and z. 

6. Show that, if three points lie on a plane which passes through the origin, 
the determinant whose rows are the coordinates of these points, all taken 
in the same order, has the value zero. 



THE DISTANCE FROM A PLANE TO A POINT 75 

6. Determine the point on the line through the points A( 4, 2, 5) and 
B(l, 3, 2) which also lies on the plane determined by the points Pi(0, 2, 1), 
P 2 (-6, -2, 0) and P(-4, 0, 1). 

7. Determine four points which lie on the plane 2x 4y + z + 7=Q. 

8. Determine four points which lie on the locus of the equation 3 a: 2 4 y* 
+ 5 * 2 = 22. 

9. Determine three points which lie on the curve whose equations are 

z + 2?/-32 = 5 and 2 x - 3 y + z = 3. 

10. Determine three points on the curve which is the locus of the pair of 

equations 

2 x - y -h 2 z = 9 and x 1 -f y* + z 2 = 26. 

11. Determine the equations of the projections on each of the coordinate 
planes of the curve of the preceding exercise. 

12. Determine the equations of the planes which pass through the following 
sets of three points each; find the intercepts of each of these planes: 



(a) P t (l, 2, 3), P 2 (2, 3, 4), 


*M3, 5, 7) 


(6) PiC-2,3,4), /M-1,2,5), 


P 3 (7, 0, 2) 


(c) P,(3, -2,5), P 2 (-2, 1, 3), 


P 3 (8, -3, 7) 


(d) P,(-4, 5, -2), P 2 (-4,3, 1), 


P s (-4, -7,3) 


(e) Pi (2, 4, -5), P 2 (-3, 1,2), 


P a (-5, 11, -4) 


(/) Pi(3, -4,2), P a (-2, -5, 1), 


P 8 (-l, -2,4) 



13. Set up the condition which the coordinates of three points must sa v 
in order that the plane determined by them shall be parallel to (a) the F-pl 
(b) the 2LY-plane; (c) the ^TF-plane; the Jf-axis; the F-axis; the Z-axihe 

14. Determine the conditions which the coordinates of three points m^ r 
satisfy in order that they lie on a line. 

43. The Distance from a Plane to a Point. To determine the 
distance from a plane to a point we make use of the projection 
method explained in Section 36; we divide the discussion into 
two parts. 

(a) The plane does not pass through the origin. 

Suppose that the direction cosines of the directed perpendicular 
from the origin to the plane arc X, M, and v and that the positive 
direction on this line is taken to be the direction from the origin to 
the plane. * Let the unsigned length of the distance from the origin 

* This agreement as to the positive direction on the perpendicular from 
the origin to the plane is entirely arbitrary; if the opposite agreement were 
made, the interpretation of some of the results obtained in the following 
pages would be different, but equally useful. The convention adopted here 
is in accord with general practice. It would be a good exercise for the reader, 
after having thoroughly mastered the next few sections, to develop this part 
of the work on the basis of the opposite convention. 



76 



PLANES AND LINES 



to the plane be designated by p and the foot of the perpendicular 
by H (see Fig. 7). Suppose furthermore that the given point is 
P(a, ft, 7) and its projection on the given plane is Q. 




FIG. 7 

We consider now the projection on the directed line / of the closed 
broken line which goes from to P along the edges of the c.p. of 
and from P back to by way of Q and H. By virtue of Theorem 
Chapter III (Section 36, page 62), we find that 

Proj/OA + Proj/AJS + Proj/JSP + Proj/PQ + Proj,Q# 
+ Proj/#0 = 0. 

If we evaluate these projections by means of Theorem 11, Chap- 
ter III (Section 36, page 62), noticing that QP \\ I and QH J_ I, 
we conclude that 

a\ + ftu + 7" + PQ + + HO = 0. 

Here we have to bear in mind that the sign of PQ is to be taken 
in accordance with the direction specified on Z, also that HO = p. 
Accordingly we obtain the result given in the following theorem. 

THEOREM 6. The distance from a plane which does not pass through 
the origin and for which the perpendicular directed from the origin 
to the plane has direction cosines X, /* and t> 9 and unsigned length />, 
to the point P(, /?, 7) is equal to 

<?P = a\ + fo + yv - p. 

(b) In case the plane passes through the origin, the specifica- 
tion of the positive direction on the line I becomes meaningless; 
but, if we designate by X, /*> v the direction cosines of either direc- 



THE NORMAL FORM OF THE EQUATION OF A PLANE 77 

tion on a perpendicular to the plane, the proof goes through as in 
part (a). We conclude therefore, since now p = 0, that the dis- 
tance QP from a plane through the origin to the point P(a, 0, 7) 
is equal to X + AU + yv, where X, /* and v are the direction cosines 
of either direction on a perpendicular to the plane. 

COROLLARY 1. The unsigned distance from a plane through the ori- 
gin to the point P(, 0, 7) is equal to the numerical value of x -f 
M _!_ 7l/> where x, M> and v are the direction cosines of either direction 
on a perpendicular to the plane. 

Remark. It follows from the above discussion that if the 
distance QP from a plane not through the origin to P turns out to 
be positive, then the direction of QP agrees with the positive di- 
rection along /, that is, P lies on the side of the plane opposite to 
that on which the origin lies; whereas, if the distance QP turns 
out to be negative, P lies on the same side of the plane as the 
origin. Furthermore, if for a plane through the origin, the ex- 
pression a\ + #M + yv turns out to be positive (negative), the 
point P lies on the side of the plane (on the side opposite to that) 
indicated by the direction which the direction cosines X, ju, v 
specify. 

If, whether the plane passes through the origin or not, the 
distance from the plane, calculated by means of Theorem 6 or 
Corollary 1, turns out to be zero, the point P lies on the plane. 
And conversely, it should be clear that if P lies on the plane, its 
distance from the plane is zero. This simple fact enables us to 
state an important further corollary of the theorem : 

COROLLARY 2. The coordinates *, y, as of a point P on a plane satisfy 

the equation 

Xx + ny + vz - p = 0. 

If the plane does not pass through the origin, then X, /i, and v are the 
direction cosines of the perpendicular directed from the origin to 
the plane and p is the unsigned distance from the origin to the 
plane; if the plane passes through the origin, then p = and 
X, ju, v are the direction cosines of either direction on a perpendicu- 
lar to the plane. v 

44. The Normal Form of the Equation of a Plane. A compari- 
son of the equation established in Corollary 2 of the preceding 
section with the general linear equation in x y y, and z yields a 



78 PLANES AND LINES 

valuable result. For we have seen in Theorem 4 (Section 41, 
page 71) that every plane can be represented by an equation of 
the form ax + by + cz + d = 0, in which a, 6, c, and d had no 
particular significance; and in Corollary 2 of Section 43 we estab- 
lished the fact that every plane can be represented by an equation 
of the form \x + ny + vz p = 0, in which X, /x> v, and p have 
the geometrical meanings stated in this corollary. But if these 
two equations represent the same plane, they must be equivalent; 
hence their coefficients must be proportional; thus there exists a 
non-vanishing number k such that 

a = k\j b = kfj,j c = kv and d = kp.* 

From the first three of these equations we conclude (see Theorem 
8, Chapter III, Section 33, page 56) that k = dbVa 2 + 6 2 + c 2 ; 
from the last it follows, since p is an unsigned number, that the 
sign of fc must be opposite to that of d. Thus k is completely de- 
termined if the plane does not pass through the origin, whereas 
its sign is left ambiguous if the plane passes through the origin. 
We have therefore obtained the following geometrical interpreta- 
tion of the coefficients in the general linear equation in x, y, and z. 

THEOREM 7. The coefficients a, 6, and c of the variables x, y, s in 
the equation of a plane, ax + by + cz -f r/ = 0, are proportional to 
the direction cosines of a line perpendicular to the plane; if d 4= 0, 
the quotients of a, 6, and c by that square root of the sum of their 
gquares which is opposite in sign to c/, are equal to the direction co- 
sines of the perpendicular directed from the origin to the plane, and 
the quotient of d by the same square root gives the unsigned dis- 
tance from the origin to the plane. 

Remark 1. Corollary 2 of Section 43 gives us another form in 
which the equation of a plane may be written. It is called the 
normal form of the equation of a plane. This form of the equation 
of the plane is characterized by the two facts that the sum of the 
squares of the coefficients of the variables is equal to 1 and that 
the constant term is negative or zero. 

Remark 2. Division of the form ax + by + cz + d = of the 
equation of a plane by +Va 2 + 6 2 + c 2 or by - Va 2 + 6 2 + c 2 

* Two equations are equivalent if any values of the variables which occur 
in it that satisfy either one of them also satisfy the other. It is a nice exer- 
cise in algebra to show that, if two linear equations in x, y, and z are equiva- 
lent, their coefficients arc proportional. 



THE NORMAL FORM OF THE EQUATION OF A PLANE 79 

according as d is negative or positive is called " reduction of the 
equation of the plane to the normal form." 

If we combine Theorems 7 and 6, we obtain the following corol- 
laries. 

COROLLARY 1. The distance from the plane ax + by + cs + d = 0, 



d * 0, to the point P(, 0, 7) Is equal to the + or _ 

=b Va 2 -f 6 2 + c 2 

sign being used according as d Is negative or positive; this distance 
will be positive or negative according as P and the origin lie on oppo- 
site sides or on the same side of the plane. 

COROLLARY 2. The unsigned distance from a plane through the ori- 
gin ax + by + cs = to the point P(, 0, 7) is given by the numerical 

aa -f- bft -f- cy 



value of 



-f 6 2 -f c 2 



If a definite choice of the sign of the square root in this last 
formula is determined upon, then those points P for which the 
distance turns out to be positive (negative) lie on the side of the 
plane (on the side opposite to that) indicated by the direction 
whose direction cosines are equal to the quotients of a, 6, and c 
divided by that square root. Although in this case the parts of 
space on opposite sides of the plane are not so readily character- 
ized as when the plane does not pass through the origin, we still 
have this essential fact that, once the sign of the square root has 
been fixed in either way, two points P(a, , 7) and P'(a', /ft', 7') 
will lie on the same or on opposite sides of the plane according as 

aa + bp + cy , aa' + fe/3' + cy' . . . . . 

, and ,, are equal or opposite in sign, 

Va 2 + fr 2 + c 2 Va 2 + b' 2 + c 2 

that is, according as aa + b0 + cy and aa' + 6/3' + cy' have the 
same or opposite signs. 

From Theorem 7, in combination with Theorem 13, Chapter III 
(Section 36, page 63) we obtain moreover the following result. 

COROLLARY 3. The angles between a line whose direction cosines 
are x, /*, v and the plane ax + by -f cs + d = are determined by the 

equation 

\a-\-nb-i- vc 



sin e = 



6 2 + c 2 



The angle between a line and a plane is the angle between that 
line and its projection on the plane; the sine of this angle is there- 
fore equal to the cosine of the angle made by the given line and a 



80 PLANES AND LINES 

line perpendicular to the plane. Corollary 3 follows from these 
observations. The ambiguity of sign corresponds to the fact that 
neither the direction on the given line nor that on the projection 
have been specified. 
Examples. 

1. To find the distance from the plane II: 2 x + 3 y 4 z + 5 = to the 
points A(-l, 2, 4), B(3, -2, 0), 0(0, 0, 0) and C (3, 3, 5), we determine the 
direction cosines of the perpendicular directed from the origin to the plane; 

2 3 _4 

it is found that X = - 7=, M = ^7=, v = - 7=- The unsigned length 
-\/29 -V29 -\/29 

rj 
of the perpendicular from the origin to the plane is * __ It follows that the 

r* nA -2 + 6-16 + 5 7V29 .,. 6-6-0 + 5 

distance II A = - 7 - = on , that II B 



7 - on , - -= - 

-V29 29 , -V29 

-5 V29 ., . nn -5 V'29 . ., . 6+9-20 + 5 _ w 

, that nO = and that 11(7 = - - = 0. We con- 

clude that A and lie on opposite sides, B and O on the same side of the plane, 
while C is on the plane. These are the geometrically essential facts concern- 
ing the positions of these points and the plane; that the side of the plane 
on which the origin lies happens to be the negative side is not of geometric 
importance, but is a result of the convention made in Section 43 (see the foot- 
note on page 75). 

2. To find the distances from the plane II : 3 .c 12 y + 4 z = to the 
points A(-3, 1, 4), 5(3, -12, 4) C(-3, 12, -4) and D(5, 1, 1), we determine 
the direction cosines of the line perpendicular to the plane. We find that 
X : M : ^ = 3 : -12 : 4, so that X = f 8 a , ju = =F}J, v = T \, in which either 
all the upper signs or all the lower signs are to be used. If we choose the 

upper signs, we find that UA = ~ 9 ~ 12 + 16 = - T 8 3 , HB = +13, IlC = -13 

and nD = T \, from which we conclude that A and C lie on the side of the 
plane opposite to that indicated by the direction whose direction cosines are 
iV> it iV b ut B and D lie on the side indicated by that direction. If we 
choose the lower signs, the signs of the four distances are reversed; this means 
that A and C lie on the side of the plane indicated by the direction whose 
direction cosines'are r \, {|, A, and B and D lie on the opposite side. These 
conclusions are obviously identical in geometrical content with those stated 
in the preceding sentence. 

45. Exercises. 

1. Determine the distances of the points A( 3, 2, 1), B(5 t 3, 1), 
(7(2, 4, 2) and D(-l, 2, -4) from the plane 3z+2?/-6z-2=0, and 
determine their positions relative to the plane. 

2. Also the distances of the points A(l, 4, -3), B(3, -2, 2), C(-5, 1, 3) 
and D(l, 0, 2) from the plane 2z-3t/ + z = 0. 



TWO PLANES 81 

3. Find the direction cosines of the lines which are perpendicular to the 
following planes: 

(a) 14 x - 3 y + 18 z + 1 = (c) 6 x - 2 y - 3 2 -f 2 = 

(&)2a? + 3y-2+4=0 (rf) .c + 4 ?/ - 8 2 - 3 = 0. 

4. Determine the distances: 

(a) from the X-axis to the plane 3 <y 4 z + 7 
(6) from the F-axis to the plane 5z 2z" 3=0 
(c) from the Z-axis to the plane 5 x 12 ?/ 8 = 0. 

6. Determine the coordinates of the point in which the plane 2 x y 2 z 
-H 4 = is met by a line through A (3, 1, 2) perpendicular to the plane. 
Find the distance from the plane to A in two ways. 

6. Find the coordinates of the point in which the plane ax -\- by -\- cz ~\- d 
= is met by the perpendicular from P(a, p, y) to the plane. 

7. Through the point A (2, -2, 6) in the plane 3z-f2?/-z-f4=0, 
a line is drawn whose direction cosines are proportional to 3, 6 and 2. Find 
the angles which this line makes with the plane. 

8. Set up the equation of a plane through the point A (2, 3, 1) and per- 
pendicular to a line whose direction cosines are proportional to 3, 4, 2. 

46. Two Planes. Two distinct planes either intersect in a line 
or else they are parallel. Since two planes are perpendicular to 
the same line if and only if they are parallel, it follows immedi- 
ately from Theorem 7 that two planes 

(1) aix + biy + CiZ + di = and (2) a 2 x + Iwj + o>z + r/ 2 = 

are distinct and parallel if and only if ai : a 2 = &i : &2 = Ci : c* ^ 
d\ : dz y that is, if the rank of the coefficient matrix of the two 
equations is 1, and the rank of the augmented matrix is 2 (see 
Definition IX, Chapter I, Section 9, page 16 and Section 20, last 
paragraph). If two equations represent the same plane, their co- 
efficients are proportional (see footnote on page 78), the two- 
rowed minors of the augmented matrix all vanish and the rank of 

the a.m. is 1. We can therefore state the following conclusion. 

. i 
THEOREM 8. The planes represented by two linear equations are 

(1) coincident If and only if the rank of the augmented matrix Is 1; 

(2) parallel If and only If the rank of the augmented matrix is 2 and the 
rank of the coefficient matrix Is 1; (3) Intersecting If and only if the 
rank of the coefficient matrix is 2. 

To determine the angles between two intersecting planes, we 
make use once more of Theorem 7. These angles are the same 
as the angles between two lines perpendicular to these planes 



82 



PLANES AND LINES 



(see Fig. 8). Therefore the angles between the planes (1) and (2) 
are equal to the angles between the lines whose direction cosines 
are proportional to a it 61, c A and to a 2 , b 2 , c 2 . Consequently, if 6 is 




Fia. 8 

used to designate any one of these angles, we conclude, using also 
Theorem 13, Chapter III (Section 36, page 03), that 

bib 2 + CiC 2 



(3) cos 6 = 



c 2 2 ) 



THEOREM 9. The cosine of any of the angles between the planes 
represented by two linear equations is equal to the sum of the prod- 
ucts of the coefficients of the like variables in the two equations, di- 
vided by the product of the square roots of the sums of the squares of 
these coefficients. 

Remark 1. The ambiguity of sign in the formula corresponds to 
the fact that the different angles formed by two planes are related 
in such a way that their cosines differ at most in sig^. 

Remark 2. If the square roots in the denominator of formula 
(3) are given the signs opposite to those of di and d 2 respectively, 
we obtain the cosine of the angle between the perpendiculars to 
the plane, directed in each case from the origin to the plane (see 
Theorem 7, page 78), that is, the cosine of the supplement of that 
angle between the planes in which the origin lies. We are sup- 
posing in this statement that neither plane passes through the 
origin. 

COROLLARY. Two planes, a\x -f b\y + ci* -f rfi =0 and a*x -f 6 2 y + 
c< 2 s -f- dz = 0, are perpendicular if and only if a { a> -f- b>b 2 + cic 2 = 0. 

An equation of the form (a\x + b\y + c\z + di) (a 2 # + b^y 
+ c%z + d 2 ) = is satisfied by values of the variables which cause 



THE LINE 83 

at least one of the factors of its left-hand side to vanish, and by 
such values only. The locus of this equation consists therefore 
of the two planes represented by the equations a\x + biy + c\z 
+ di = and a^x + btfj + c 2 z + d 2 = 0. It should be clear that 
this observation may be generalized as in the following theorem. 

THEOREM 10. If F(x 9 y, z) =/i(, y, a) fi(x 9 y, )... fk(x 9 y, ), 
the locus of the equation F(x 9 y, s) = consists of the loci of the 
equations /i (*, y, s) = 0,/ 2 (*, y, s) = 0, . . . ,/*(*, y, s) = 0. 

DEFINITION V. If the function F(*, y, s) is factorable into real 
factors (that is factors which involve only real operations on the 
variables), the locus of the equation F(x 9 y, s) = is called a degener- 
ate locus. 

47. The Line. The coordinates of every point on the line of 
intersection of the two planes represented by equations (1) and 
(2) of the preceding section satisfy, in case (3) of Theorem 8, these 
two equations. Conversely, every point whose coordinates sat- 
isfy these equations lies on the line of intersection of the planes. 
We say therefore, in accordance with Definition II (Section 39, 
page 68), that "two linear equations ciix + b^y + c\z + c?i = 
and a 2 x + b' 2 y + c$z + cl 2 = 0, whose coefficient matrix has rank 
2, are the equations of a line/' 

Remark. A line, thus defined as the intersection of two 
planes, has as its equations those of two planes passing through 
it. But there is a single infinitude of planes which pass through a 
given straight line (see Section 49) and the equations of any two 
of these planes can be taken as the equations of the line. Thus it 
is seen that one and the same straight line can be represented by 
any one of an infinite number of pairs of linear equations. The 
reader may at first be troubled by this lack of definiteness; he 
will do well to think this question through until it has become 
clear to him. 

The results obtained in Chapter III, where the line was dis- 
cussed as a locus of points, can now be interpreted in the light of 
the point of view presented in the first paragraph of this section. 
The equations found in Theorems 9 and 10, and in Corollary 1 of 
Theorem 10, Chapter III (see Section 34, pages 57 and 59) are 
all linear equations; and it is readily seen that in each case a pair 
of equations can be selected whose coefficient matrix has rank 2. 
For example, the equations of Theorem 9 may be written in the 



84 PLANES AND LINES 

following form: 

(yB - VA)X - ( X B *A)y + yA*B - ysXA = 0, 

(ZB - z A )y - tea - VA)* + z A ys - ZWA = 0, and 



The second order determinants of the coefficient matrix of the first 
two of these equations have the values 



those formed from the coefficient matrix of the second and third 
equations have the values 

-(XB ~ X A ) (^B ~ Z A ), (y B - VA) (ZB - ZA), -(ZB - z A ) 2 ; 

and those obtained from the third and first equations have the 
values 

-(XB - XA)*J (ys - VA) (XB - XA), -(XB ~ X A ) (Z B - Z A ) . 

Now it should be clear that if A and B are distinct points at least 
one of these second order determinants must have a value which 
is different from zero; therefore a pair of equations can be selected 
whose coefficient matrix has rank 2. 

Similar arguments can be made for the equations of Theorem 
10 and Corollary 1 of Theorem 10. A mere restatement of the 
earlier results in terms of the terminology which was introduced 
and justified at the beginning of the present section, leads to the 
following theorems. 

THEOREM 11. The equations of the straight line which passes 
through the points A and B may be written in the form: 



*-* y ~y * - 



*B 

also in the forms: 



or 



- x y-y * - * 



THEOREM 12. The equations of the directed line which passes 
through A and whose direction cosines are x, /* and v may be written 
In the form: 

* -* A _y ^y A ^*~*A 

M 



THE LINE 



85 



THEOREM 13. The equations of the undirected line which passes 
through the point A and whose direction cosines are proportional to 
f, m and n may be written in the form: 

x - X A y -y A * - * A 



I m n 

Remark 1. It should be clear that in each of these theorems the 
line is the intersection of three planes, any two of which suffice to 
determine it. The three planes are, in each case, parallel to the 
X-, F-, and Z-axes; they are indeed the planes which project 
the line on the three coordinate planes (see Fig. 9), the diagonal 
planes of the c,p. of any two points on the line. 




Fio. 9 

Remark 2. The equations of the line, established 
11, 12, and 13, lose meaning whenever one of the denominators 
vanishes. In spite of this disadvantage these forms for the equa- 
tions of the line are, in general, more convenient than the extended 
form obtained by equating two of the fractions at a time and then 



86 PLANES AND LINES 

clearing of fractions. This extended form becomes imperative if 
one of the denominators vanishes. Frequently one finds the 
condensed form used even in such cases; this is however not to be 
recommended even though this apparently meaningless form is 
intended to be symbolic for the extended equations. The diffi- 
culty referred to here can be obviated by use of the parametric 
equations, already obtained in Chapter III (see the Corollary of 
Theorem 9, and Corollaries 2 and 3 of Theorem 10, Section 34, 
pages 58, and 60), whose existence is formulated again as fol- 
lows. 

THEOREM 14. The equations of the line through the points A and B 
may be written In the following form: 



S 



_ 

* ~ 



- 1+r ' ""14- r ' ~ 1 +r * 

THEOREM 15. The equations of the directed line through the point 
A whose direction cosines are equal to X, M and v may be written in the 
form: 

x = X A + Xs, y = y A + /us, * = S A -f vs. 

THEOREM 16. The equations of the undirected line through the 
point A whose direction cosines are proportional to 1 9 m and n may be 
put in the form: 

* = X A -f It, y y A + mt 9 * = S A + nt. 

Remark. In each of the last three theorems the line is given by 
means of three equations; but these equations involve four vari- 
ables, namely the coordinates x y ?/, and z of the variable point 
along the line, and the parameter r, s, or t. The locus of these sets 
of equations has therefore one degree of freedom (see Section 
39, page 67). The geometric significance of the parameters r, s, 
and t was discussed in Corollaries 2 and 3 of Theorem 10, Chapter 
III, and in Remark 4 following the Corollary of Theorem 9 (see 
Section 34, page 58) ; it is desirable that the reader recall this in- 
terpretation of the parameters at this point. 

If an undirected line is given by means of two linear equations 
in the general form, like equations (1) and (2) of Section 46, whose 
coefficient matrix has rank 2, these equations can be reduced to 
any one of the forms given in Theorems 11 to 16, as soon as the 
coordinates of two points on the line and the ratios of its direction 
cosines have been determined. 



THE LINE 87 

Since the coefficient matrix is of rank 2, the equations can be 
solved for two of the variables in terms of the third. By assigning 
values to this third variable arbitrarily, an infinite number of solu- 
tions of the equations can be obtained; but each of these solutions 
furnishes the coordinates of a point on the given line. 

The direction cosines of the line are found by means of the fol- 
lowing theorem: 

THEOREM 17. The direction cosines of the line of intersection of 
two intersecting planes are proportional to the two-rowed minors of 
the coefficient matrix of their equations, taken alternately with the 
plus and the minus signs. 

Proof. The proof of this very useful theorem can be made in 
various ways. We shall make use here of Corollary 1 of Theorem 
6, Chapter III (Section 33, page 56). Suppose that P\ and P 2 
are two arbitrary points on the line of intersection of the planes. 
Then 

+ biiji + CiZi + h = 0, a&i + 6 2 ?/i + c 2 2i + r/ 2 = 0, 

Ci2 + r/i = and a 2 z 2 + 6 2 t/ 2 + c 2 2 2 + ^2 = 0. 



If we subtract these equations in pairs, we find that the differences 
of the coordinates of PI and P 2 satisfy the following two linear 
homogeneous equations : 

- 7/ 2 ) + ci(zi - z 2 ) = and 02(0?! - x 2 ) 
Zz) = 0. 



Since the two given planes intersect, it follows from (3) in Theorem 
8 (Section 46, page 81) that the rank of the coefficient matrix 
of these equations is 2. Theorem 4, Chapter II (Section 25, 
page 41) gives us the means therefore to determine from these 
equations the ratios of the coordinate differences of PI and P 2 . 
But we know from Corollary 1 of Theorem 6, Chapter III (Section 
33, page 56) that these coordinate differences are proportional 
to the direction cosines X, /i, v of the line. We find therefore that 



bz ( 
This completes the proof of the theorem. 



88 PLANES AND LINES 

Examples. 

1. The planes represented by the equations 2z-2/-f3z-4=0 and 
2 x y + 5z + 3 = intersect in a line; for the rank of the matrix 

2-13 
2 -1 5 

is clearly 2. To determine points on the line of intersection, we solve the 
equations for x and z in terms of ?/ (we could equally well solve them for ?/ and z 
in terms of x, but not for x and y in terms of z\ why not?) by Cramer's rule. 
We find 



By selecting values for y and calculating the corresponding values of x and 
z from these equations, we can find. as many points on the line as we wish; 
thus we locate the points A (8, , -}, #(*/, -2, -I) and C Y (% 0, - J) on the 
line of intersection of the given planes. Having determined these points, we 
can find the direction cosines of the line most simply by use of Corollary 1 
of Theorem 6, Chapter III directly; it is found that 

X : M : * = 8 - V : I + 2 : - J + J = I : I : = 1 : 2 : 0. 

1 2 

Consequently X = j-r, /* = ^, v - 0, so that the line makes an angle 

=fc v 5 =fc v 5 

of 90 with the Z-axis and is therefore parallel to the X F-plane; this could 
have been foretold from the fact that all of its points have the same z-co- 
ordinate, j. 

The ratios of the direction cosines can also be found by applying the formula 
proved in Theorem 17; this gives us \ 



\ '. fJL '. V = 



-1 3 
-1 5 



12 3 
2 5 



2 -1 
2 -1 



= -2 : -4 : = 1 : 2 : 0. 



In accordance with Remark 2, following Theorem 13 (page 85), the non- 
parametric forms of the equations of the line as given in Theorems 11, 12, and 
13 are not desirable in this case. The parametric forms of the equations are: 

a: = 8 + f, y = 2 + 2 , 2 = - j, where t AP 



-2+|r PB 

---' *--*' where r = ; 



29 , a 2s , 

x = -r TT , ?/ = ^^ , z = I, where s = 4P. 
* V5 V5 

We observe that the three sets of values of x, y, and z given by the different 
parametric equations satisfy the equations of the two given planes identically 
in t, r, or s respectively. 



THE LINE 89 

2. The planes represented by the equations 2x ?/ + 3 z 4 = U and 
4 z 2 T/ + 6 z + 3 =()are parallel; for the rank of the c.m. is 1 and the 
rank of the a.m. is 2. The parallelism or coincidence of two planes can readily 
be recognized upon inspection of the equations. 

3. The direction cosines of the line of intersection of the planes represented 
by the equations 10z + 3?/-4z+8 = and 4u;+3?/-3z--4=0 

are proportional to Ur ; <**Ct 
7>-V. *** 



3 -4 
3 -3 



" 4 -3 



10 3 
4 3 



that is, X : /z : v = 3 : 14 : 18. 



Since 3 2 + 14 2 + 18 2 = 529 = 23 2 , it follows that X = &, M = ii " = =*=it- 



Solution df the equations for x and i/ in terms of z leads to x = -A ' 2, 

y + 4 ; we arc now able to determine readily as many points on the line 
J 

as we wish. 

The parametric equations of the line may be written in the following forms: 



a: = -2 + 3 , ?/ = 4 -f 14 /, z = 18 <; 
_ -2 -f 7r _ 4 + 46 r 54 r 

* - 1 +r ' ^ ~ 1 -hr ' " ~ IT"r * 

We can verify that these values of x, y, and z satisfy the given equations of 
the planes identically. 
The angles between the two planes are given by the equation: 

10-4 -h 3-3 -f (-4) (-3) 61 



COS0 = 



V10 2 -f 3 2 + (-4) 2 X V4 2 -1- 3 2 -f (- 3) 2 5 Vl70 



The direction cosines of the directed lines from the origin perpendicular to 

.u T i* 2 3 4 ^ 4 3 3 , 

the plane are equal to T ., 7-, 7^ and p=, T=, 7=- There- 

V5 5V5 5\/5 V34 V34 >/34 

fore the cosine of that angle between the planes in which the origin lies is 

i . 6l 
equal to T=T 

5V170 

48. Exercises. 

1. Determine by inspection which of the following pairs of equations repre- 
sent intersecting, which parallel planes, and which coincident planes: 

(a)3z-?/-h42 + l=0 and 2x-f?/-2z-f3=0 

(b)2s + y-3;8 + 4 = and 2z + ?/-32-4 = 

(c)z + 2y + 4z-3=0 and x-2y+4z + l=Q 

(d) x -y + z = and 2ar-2?/ + 22 + 7 = 

2. Determine the angles between the planes of the intersecting pairs of 
planes in the preceding exercise. 



90 PLANES AND LINES 

3. Determine the distances between the following pairs of parallel planes: 

(o)a:-8y + 4-3=0 and x - Sy + 4 z + 15 = 0; 
(6) 2z-3?/-6z + 5=0 and 2 x - 3 y - 6 z -f 19 = 0; 
(c) a; + y + z + 6 = and z + 2/-fz-8 = 0. 

4. Write the equations of the lines which pass through the following pairs 
of points: 

(a) 4(-3, 5, 2) and 5(5, 4, -2), (c) A(5, 2, -3) and B(-l, -1, -1), 
(6) A(4, -3, 1) and B(-8, 3, 5), (d) A(-2, 4, 1) and 5(3, -5, 2). 

6. Write the equations of the line through A (3, 4, 1) and perpendicular 
to the plane 2 x y -\- 2 z 5 =0. Determine the coordinates of the 
point in which the plane is met by this line. 

6. Set up the equation of the plane through the point A( 2, 3, 4) and 

(a) parallel to the plane 3 x + y 5 z -f 7 =0; 

(b) perpendicular to the line ^ = y = ^ . 

7. Determine the parametric equations of 

(a) the line of intersection of the planes 3 x y -J- 3 z 2 and 

z+2?/-3z-f4=0; 
(6) the line through the point .4(1, 3, 5) and parallel to the line of 

intersection of the planes 3z + 2/ + 2z 3 = and 6 x -f- 3 y 

+ 2 z + 5 = 0. 

8. Determine a plane through the points A (2, 1, 4) and #( 1,3,2), 
which intersects the plane 3 z ?/ 2 z = 4 in a line that makes equal angles 
with the coordinate axes. 

9. Find the equation of a plane through the points A(l, 3, 2) and 
B(2, 4, 5) which is perpendicular to the plane 3 x + 6y 4 z 5 = 0. 

10. Find the parametric equations of a line through the point A (2, 5, 3) 
and parallel to the line of intersection of the two planes represented by the 
equations 2 - 6 xy + 9 y* - 4 z* + 12 z - 9 = 0. 

11. Set up the equation of a plane through the point A( 1, 4, 3) and 
perpendicular to the line of intersection of the two planes represented by the 
equation 9 x 2 4 y 2 -f z 2 6 xz 4 y - 1 =0. 

12. Find the equation of a plane through the point A (3, 2, 1) and per- 
pendicular to the two planes x y + z -f- 4 =0 and 2 x y 2 z -f 3 =(). 

13. Determine the coordinates of the point in which the plane 3 x - 4 y 
-f z - 3 = is met by the line :c = 2-2J, ?/= - 1 3 J, z = 5 -H. 

14. Determine the coordinates of the point in which the plane 4 x -f y 3 z 
-|- 5 = is met by the line which joins the points A ( - 1 , 3, - 2) and B (4, - 3, 1 ) . 

49. The Pencil of Planes. The Bundle of Planes. If the 
left-hand sides of the equations 

(1) aix + biy + ciz + di = 



THE PENCIL OF PLANES 91 

and 

(2) Otx + b 2 y + c 2 z + d 2 = 

arc denoted by EI and E 2 respectively* and if k t and k 2 are arbi- 
trary constants, then the equation 

(3) kiEi + k 2 E 2 = 

is also a linear equation. Its locus is therefore a plane. Moreover 
equation (3) will be satisfied by the coordinates of those points 
which lie on both planes EI and E 2 , that is, by the points on the 
line of intersection of these planes; and this last statement holds 
true whether ki and k 2 are constants or not, because the coordi- 
nates of the points on the line of intersection of the planes cause 
both EI and 7 2 to vanish. Consequently, the equation represents 
a plane through the line of intersection of the planes EI and E 2 for 
any constant values assigned to &i and k 2 . 

On the other hand the equation of every plane through this line 
can, by suitable choice of the values to be given to the constants 
ki and & 2 , be put in the form (3). We see, in particular, that for 
fci = 1 and k 2 we obtain the plane E\\ and for k\ = and 
k 2 = 1, we obtain the plane # 2 . And if any other plane through 
the line of intersection, I, of the two planes is given, and if P(a, 0, 7) 
is an arbitrary point in such a plane but not on J, then equation 
(3) will represent the given plane, provided ki and k 2 are so chosen 
that 

kifaa + bi0 + Ciy + di) + k 2 (o^a + b 2 $ + C 2 y + d 2 ) = 0. 

Since P does not lie on I and hence not on both planes, the two 
expressions in the parentheses do not both vanish; consequently 
the ratio ki : k 2 can always be determined in such a manner that 
the last written equation is satisfied. If numbers which have 
this ratio are substituted for ki and k 2 in equation (3), this equation 
will indeed have the given plane as its locus. If we introduce now 

* When this abbreviated notation is employed for the left-hand side of the 
linear equation in x t y, and z, it is usually convenient to use the same letter to 
designate the plane which is the locus of the equation. Thus we shall speak 
of "the plane EI " instead of using the longer and more explicit phrase "the 
plane whose equation is EI s a\x + b\y + c\z + d\ = 0"; this usage does 
not frequently lead to confusion. 



92 PLANES AND LINES 

the expression " pencil of planes" to designate the set of all the 
planes which pass through a line, we obtain the following theorem. 

THEOREM 18. The pencil of planes through the line of Intersection 
of the planes E { and E 2 Is represented by the equation kiEi -f k 2 E-> = 0, 
In which k L and k> 2 are arbitrary constants not both zero. 

Remark 1. Since the geometrical significance of equation (3) 
is not altered when it is multiplied through by a non-zero constant, 
the pencil of planes through the intersection of the planes E\ and 
E% is also represented by the equation EI + kE 2 = 0, where 

k 

k = T- , except that in the latter form the plane E 2 which is ob- 
KI 

tained from equation (3) when ki = 0, is excluded. For this 
reason the form (3) of the equation of the pencil deserves preference. 

Remark 2. The lack of definiteness in the equations of a line 
pointed out in the remark at the beginning of Section 47 (page 83) 
can now, at least partially, be provided for, inasmuch as we can 
say that the line which is given by the pair of linear equations 
EI = and E z = is also determined by any two equations of the 
form (3), that is, by any two planes of the pencil of planes through 
this line. 

Remark 3. The ratio ki : k% is a parameter in the equation (3) 
of the pencil of planes. It is constant for any one plane of the 
pencil, it varies as we pass from one plane in the pencil to another. 
The pencil of planes is a "one parameter family of planes." It 
will be instructive for the reader to compare the character of the 
parameter ki : & 2 in equation (3) with that of the parameters r, s, 
and t in the parametric equations of the line (see Section 47, 
page 86). 

Remark 4. The method used for determining the equation of 
the pencil of planes finds frequent application throughout Ana- 
lytical Geometry (see e. g. Exercise 3, Section 73, page 148 and 
Section 82, page 168). In connection with one of the remarks 
made in the opening paragraph of the present section, we observe 
that equation (3) does not represent a plane, if ki and k z are not 
both constants. The surface which it does represent will, how- 
ever, still pass through the line of intersection of the planes EI 
and Z? 2 . Furthermore if Si = and fi> 2 = are the equations of 
two arbitrary surfaces, the equation fciSi + feS 2 = represents a 



THE PENCIL OF PLANES 93 

surface which passes through all the points common to the two 
given surfaces, no matter what k\ and fc 2 may be. 
Examples. 

1. To determine the equation of a plane which passes through the line of 
intersection of the planes 3 x 2y + z 4 =0 and x + 5y 2 z + 3 =0 
and which is moreover perpendicular to the plane 2x + y 3z + l = 0, 
we consider the pencil of planes through the given line. The equation of this 
pencil is 

fa (3 x - 2 y + z - 4) + kt(x + 5 y - 2 z + 3) = 0. 

If a plane of this pencil is to be perpendicular to the plane 2x + y 3 z -f 1 
= 0, fa and fa must satisfy the condition which follows from the Corollary 
of Theorem 9 (Section 46, p. 82), namely, (3 fa + fa)2 + (-2 fa + 5 fa) 
+ (fa - 2 fa) (-3) = 0. This leads to the condition fa + 13 fa = 0, that 
is, fa : A~ 2 = 13 : 1. The equation of the required plane is therefore 
13 (3 x - 2 // 4- z - 4) - (x + 5 y - 2 z + 3) = 0, or 38 x - 31 y + 15 z 
- 55 = 0. 

2. To determine a plane through the line whose parametric equations are 

x = -4 + 3 , y = 5 - , z = 3 + 2 J 

and through the point P( 4, 3, 3). 

First solution. The parameter t may be eliminated between the first two of 
the parametric equations of the line and also between the last two. This 
furnishes the two linear equations x + 3 y 11 =0 and 2 y + z 13 = 0; 
and the given line is the line of intersection of the planes which these equations 
represent. The equation of the pencil of planes through the given line can 
therefore be written in the form fa(x + 3 y 11) -f- fa (2 y + z 13) =0. 
Since the point P( 4, 3, 3) must lie on the required plane, the constants fa and 
fa must be so selected that fa(-4 + 9 - 11) + fe(6 + 3 - 13) = 0, or so 
that 6 fa 4 fa = 0; hence fa : fa = 2 : 3. We conclude that the 
equation of the required plane is 2(z-f3?/--ll) 3(2 y + z 13) =0 
or 2 x - 3 z + 17 = 0. 

Second solution. The required plane is determined by P and any two points 
on the line. Such points can be found at once when the line is given by para- 
metric equations, by assigning two values arbitrarily to the parameter. 
The values t = and t = 1 yield the points A (-4, 5, 3) and B(-l, 4, 5). 
The equation of the plane can now be written in the three-point form (see 
Theorem 5, Section 41, page 73). Thus we find the equation 

x y z 1 
-4 3 3 1 
-4531 u ' 
-1451 

which, upon development, reduces to the form 2z 3z-}-17=0, found 
by the first method. 



94 PLANES AND LINES 

If the plane ax + by + cz + d = is to pass through a fixed 
point P(cx f j8, 7), its coefficients must satisfy the condition aa + bfi 
+ cy + d = 0, so that d = aa b0 cy; and it should be 
clear that if d has this value, the point P will lie on the plane. 
Upon introduction of the term "bundle of planes" to designate 
the set of planes which pass through a fixed point, we can state the 
following theorem. 

THEOREM 19. The bundle of planes through the point P(a, , 7) Is 
represented by the equation a(* - ).+ b(y - 0) + c( - 7) = 0, in 
which a, 6, and e are arbitrary constants, not all zero. 

Remark. The ratios of the constants a, b, and c are the param- 
eters in the equation of the bundle of planes; the bundle of planes 
is a "two parameter family of planes.'' 

50. Exercises. 

1. Determine the equation of a plane through the line of intersection of the 
planes 3x-y + 2z-\-2=Qa.nd 2x + 4y-3s + l = 0, and 

(a) through the point A ( 1,3, 2); 

(6) perpendicular to the plane 4 x 5y + z 2 = 0; 

(c) through the origin; 

(d) parallel to the 7-axis; 

(e) parallel to the Z-axis. 

2. Determine the equation of a plane through the line x = 2 3 t, y I 
+ 6 t, z = -3 - 2 t and through the line x = 2 + t, y = I - 2 t, z = -3 
+ 2t. 

3. Write the equation of a plane through the point A( 3, 4, 1) and per- 

,. , . ,, ,. x -\- 2 z 4 

pendicular to the line 5- - y 2 - = 

j z 

4. Prove analytically that every pencil of planes contains at least one 
plane parallel to the X-axis, at least one parallel to the F-axis and at least one 
parallel to the 2T-axis. Under what conditions will a pencil contain more 
than one plane in such position? 

6. Prove that every bundle of planes contains exactly one plane parallel 
to the FZ-plane, one parallel to the ZJf -plane and one parallel to the XY- 
plane. 

6. Determine the equation of the plane through the line of intersection of the 
planes 3x 6y 2z + 5 =0 and 2 x y 2z + 3 = which is per- 
pendicular to the first of these planes. 

7. Find the equations of the planes which bisect the angles between the 
planes 2z-6?/-32-M=0 and 4*H-y-8z + 5=0. Hint: This 
problem can be solved by observing that the bisecting planes belong to the 
pencil of planes through the line of intersection of the given planes. Another 



THREE PLANES 95 

method of procedure is based on regarding the bisecting planes as the locus 
of points whose distances from the two given planes are equal or equal nu- 
merically but opposite in sign. 

8. Find the equations of the planes which bisect the angles between the 
planes atf + b\y -f- c\z + dt = and a 2 x + hy + c& + d 2 = 0. 

51. Three Planes. A Plane and a Line. In this section we 
shall be concerned with the question of determining from the 
equations of three planes, E\, E%, and E 3 , how they are placed with 
respect to each other, that is, with the problem of extending the 
result stated for two planes in Theorem 8 (Section 46, page 81). 
Let the equations of the planes be 

#1 = aix + biy + ciz + di = 0, E 2 55 a 2 x + b^y + c 2 z + d 2 = 0, 
#3 = azx + b*y + c 3 z + r/ 3 = 0. 

It should be clear that if no two of the planes are parallel or 
coincident, the ranks of the c.m. and of the a.m. of the system of 
equations must be at least 2 (compare Theorem 8, Section 46, 
page 81). We obtain further results by means of Theorems 1 
and 8 of Chapter II (see Sections 21, page 36, and 27, page 44). 
If the coefficient matrix is of rank 3, the system of equations has 
a unique solution; in this case the three planes have a single point 
in common. If the coefficient matrix has rank 2, the system of 
equations possesses a single infinitude of solutions or no solution, 
according as the rank of the a.m. is 2 or 3; in this case therefore the 
planes will have a line in common or no point in common, according 
as the rank of the augmented matrix is 2 or 3. If the rank of the 
c.m. is 1, the rank of the a.m. can not exceed 2 (Why?). In case 
it is 2, at least one pair of planes must be parallel; if it is 1, the 
three planes must be coincident. In this case therefore the planes 
are coincident or else they have no point in common, according as 
the rank of the a.m. is 1 or 2. We have therefore obtained the 
following conclusion. 

THEOREM 20. Three planes will (1) have a single point In common 
If and only if the rank of the coefficient matrix of its equations is 3; 
(2) have a single line in common if and only if the ranks of the coeffi- 
cient matrix and the augmented matrix are both 2; (3) be coincident 
if and only if the ranks of the coefficient matrix and the augmented 
matrix are both equal to 1; (4) have no point in common if and only 
if the ranks of the coefficient matrix and the augmented matrix are 
unequal. 



96 



PLANES AND LINES 



Remark 1. If the planes have a single point in common, they 
form a trihedral angle, see Fig. 10a; if they have a single line in 
common, they are three planes of a pencil (see Fig. 106), unless 
two of the planes coincide ; if they have no points in common, they 
form a triangular prism (see Fig. lOc), unless there is a pair of 
parallel or coincident planes among them. Since parallelism 
and coincidence of planes are readily determined by inspection of 
their equations (compare Exercise 1, Section 48, page 89), the 
following corollary of Theorem 20 is of considerable use in nu- 
merical cases. 






FIG. 10a 



FIG. 106 



FIG. lOc 



COROLLARY 1. Three planes of which no two are either parallel or 
coincident will form (1) a trihedral angle if and only if the rank of the 
coefficient matrix of their equations is 3; (2) a pencil of planes if and 
only if the ranks of the coefficient matrix and the augmented matrix 
are both equal to 2; (3) a triangular prism If and only if the rank of the 
coefficient matrix is 2, while the rank of the augmented matrix is 3. 

For future reference it is convenient to state separately the fol- 
lowing immediate deduction from Theorem 20. 

COROLLARY 2. Three planes have one or more points in common if 
and only if the ranks of the augmented matrix and of the coefficient 
matrix of their equations are equal. 

If the rank of the c.m. is not less than 2, two of the three equa- 
tions can be taken as the equations of a line. The corresponding 
restatement of Theorem 20 leads to the following Corollary. 



TIP* PLANE AND THE LINE 97 

COROLLARY 3. A plane and a line will (1) meet in a point if and 
only if the rank of the coefficient matrix of the three equations used 
to represent them is 3; (2) be parallel if and only if the ranks of the 
coefficient matrix and the augmented matrix of these equations are 
2 and 3 respectively. The line will lie in the plane if and only if the 
rank of each of these matrices is 2. 

Remark. It is of interest to observe that these conditions must 
continue to hold true when the equations of the line are replaced 
by the equations of any two planes of the pencil of planes through 
the line. Hence the rank of the matrices of the 3 linear func- 
tions EI, EZ, E 3 is not changed if the functions EI and E 2 are 
replaced by k\Ei + k z E 2 and l\Ei + l^E^ respectively. Thus we 
obtain, for the special case n = 3, a geometrical interpretation of 
a part at least of Theorem 14, Chapter I (Section 10, page 18). 

62. The Plane and the Line, continued. The geometrical 
content of Corollary 3 in the preceding section will become more 
apparent if the conditions of that corollary are interpreted in 
terms of the direction cosines of the line. Let the equations of 
the line be given in the form stated in Theorem 12 (Section 47, 
page 84) : 

x - a. = y - 13 = z - y 

X p. V 

It will always be possible to select among the three equations here 
represented two whose c.rn. has rank 2; these equations can then 
be taken as the equations of the line. If we suppose that p = 0, 
these equations may be taken to be the following two : 

nx \y (XIJL + 0X = and vy M^ ~ &v + 7M = 0. 

The condition that the given line shall meet the plane ax + by 
+ cz + d = in a point, can therefore, in virtue of Corollary 3, 
Section 51, (1), be written in the form: 



0) 







-X 



v - 



a b c 



4=0. 



If we develop this determinant and divide out the factor /u which 
was supposed to be different from zero, we find the condition 

(2) aX + bfji + CP 4= 0. 



98 PLANES AND LINES 

On the other hand the line will lie in the plane or be parallel to 
it if and only if the determinant in (1) vanishes, that is, since 
M 4 1 0, if and only if 

(3) a\ + bn + cv = 

while at least one two-rowed minor of the determinant in (1) 
has a value different from zero. 

To distinguish the case of parallelism from the case in which the 
line lies in the plane, we have to consider the augmented matrix 
of the system of equations 

(4) nx \y oifjL + 0\ = 0, vy IJLZ $v + yn = 0, 
ax + by + cz + d = 0. 

There is no loss of generality, by virtue of the hypothesis that 
the rank of the matrix of the determinant in (1) is 2, if we suppose 
that not all the cofactors of the elements in the second column 
of this determinant vanish; let us denote the values of these 
cofactors by Ci, C 2 , and (7 3 and let us suppose that Ci 4= 0. We 
know then from Theorem 14, Chapter I (Section 10, page 18), 
that the rank of the augmented matrix of the system of equations 
(4) is not changed if the first row is replaced by Ci times the first 
row, plus C 2 times the second row plus Cs times the third row. 
But if this operation is carried out, the first and third elements 
of this row will reduce to zero by Theorem 13, Chapter I (Section 
7, page 13), the second element vanishes on account of (3), 
and the fourth element becomes 



Consequently the augmented matrix of the system of equations 
(4) will have rank 2 or 3 according as this last expression is or is 
not equal to zero. Now, Ci = a/*, C 2 = c/z and C 3 = /x 2 - If 
these values are substituted in the expression for the fourth 
element, above, we find that the rank of the augmented matrix 
is 2 or 3 according as the equation 



- jSX) + cp(-fr + TM) + M 2 d = 
is or is not satisfied. This condition reduces to 

(act + cy + d)^ - (aX + ci/)ftu = 0. 



THE PLANE AND THE LINE 99 

But from (3) it follows that a\ + cv = fr/*; if this is used in the 
preceding equation and if the non-vanishing factor p, 2 is divided 
out, we are led to the condition 

(5) aa + bp + cy + d = 0. 

We have therefore obtained the following equivalent form of 
Corollary 3 of Theorem 20. 

THEOREM 21. The line I through the point P(, (3, ?) and with di- 
rection cosines X, /*, v will (1) meet the plane ax -f by -f cs -f- d = in 
a single point if and only if aX -f &/* + cv = 0; (2) be parallel to the 
plane If and only if aX + b 4- cv = and aa + bft + cy + d 4= 0; or 
lie in the plane if and only if a\ -f &M + cv = and aa -f bp + cy -f 
d = 0. 

Remark 1. The geometrical interpretation of these conditions 
should be obvious. For it follows from Corollary 3 of Theorem 
7 (Section 44, page 79) that the condition (3) requires that the 
angle between the line and the plane shall be or 180, that is, 
that the line shall lie in the plane or be parallel to it. And the 
condition (5) clearly states that the point P(a, 0, 7) must lie in 
the plane ax + by + cz + d = 0. 

Remark 2. We have been interested in deriving Theorem 21 
from Theorem 20 in order to illustrate the power of this general 
theorem. It must be observed, however, that the conditions of 
Theorem 21 are obtained more directly if the equations of the line 
are taken in the parametric form of Theorem 15 (Section 47, 
page 86), x = a + X,s-, y = + JJLS, z = y + vs. For if these 
expressions are substituted for x, y, and z in the equation of the 
plane, ax + by + cz + d = 0, we obtain the linear equation : 

(a\ + &M + cv)s + (aa + h0 + cy + d) = 

from which the value of s is to be ascertained, which determines 
the point of intersection of the line with the plane. From this 
equation it is evident that, if aX + 6/i + cv 4 1 0, the equation 
has a single root and the line meets the plane in a single point; 
if a\ + bfj, + cv and aa. + fc/3 + cy + d = 0, the equation has 
no solution, and the line is parallel to the plane; if aX + fyu + cv 
= and act + 6/3 + cy + d 0, the equation is satisfied by 
every value of s and the line lies in the plane. 



100 



PLANES AND LINES 



Examples. 

1. Let it be required to determine the equation of a plane through the 
point A( 4, 1, 2) and parallel to the lines 

/ .^=J_?LJ~*_I_ 3 i i -*. + 3_y-_4_* + 2 



-2 



and 



-1 



-2 



If P(;r, ?/, z) is an arbitrary point on the required plane, it must be possible 
to determine four numbers a, b, r, and d, not all zero, such that 

ax + by + cz -f- d = 0; such that 

4 a + 6 -}- 2 c -|- rf =0, since -A is to lie in the plane; and such that 
4 a - 2 b -f 3 c = 0, and 

a + 3 6 2c =0, since h and J 2 are to be parallel to the plane 

and their direction cosines are proportional to 4, 2, 3 and to 1, 3, 2 
respectively. These four linear homogeneous equations in a, 6, c, and d possess 
a non-trivial solution only if the value of the coefficient determinant is zero 
(see Corollary of Theorem 2, Chapter IT, Section 22, page 38). Hence the 

x y z 1 



condition on the coordinates of P is that 



- 0, or that 



z 

-4121 
4-2 30 
-1 3-20 

x y 2 z + 9 = 0. The locus of this linear equation is a plane; it is easy to 
verify that this plane meets the required conditions. 

2. If it is required to show that the line of intersection of the planes x 2 y 
-f z 4 = o and 3x + 5y 2 z + 4 = is parallel to the plane 7 x 3 y 
-f 2 z 5 = 0, we can proceed in various ways. Using Corollary 3 of 

Theorem 20 (Section 51, page 97), we can show that 



and that the rank of the matrix 





1 -2 1 




that 


3 5 -2 


= 




7 -3 2 




is 3; thus the problem 



1-2 1-4 
35-24 

7 o o r 

i o <-/ j 

is made to depend on the evaluation of determinants. We can also reduce the 
equations of the line to the point-direction form, established in Theorem 12 
(Section 47, page 84) and then apply Theorem 21. 

53. Exercises. 

1. Determine the relative positions of the planes in each of the following 
sets: 

(a) 3z-2i/-f42 = 0, 2s + 3y-z + 3=0, x~ 4y + 2z + 2 = 0; 
(6) z-2^-42+3 =0, 3s + ?/-z + 2 = 0, 3x + 8y + Wz - 5 = 0; 

(c) 2z + i/-3z + 4=0, 4z + 27/-6z + 5=0, 3z- Qy+z + 1 = 0; 

(d) z-i/ + 2-3=0, 2z+?/ + 3z = 0, 3z+52/ + 3z-l=0; 

(e) 4z + ?/-32:-f-2 = 0, 2z-3?y + z-4=0, 72/-5z + 4=0; 
(/) * + 77 + 2 + 1=0, 2x + 2y+3z-4 

(0) 5z-22/-7z + 3 =0, 10z-4i/- 14z-2 = 0, 
-8 = 0. 



FOUR PLANES 101 

2. For those sets of planes in the preceding exercise which form a trihedral 
angle, determine the point common to the three planes. 

3. For those sets of planes in Exercise 1 which belong to a pencil of planes, 
determine the direction cosines of the line common to the three planes. 

4. Set up the equation of a plane through the point A (2, 1, 3) and paral- 
lel to the lines x = 2 - 3t, y = l+2t,z = 3 t and z = 4 + 3f, y = 
-3 + 5 t, z = 1 - 2 t. 

5. Set up the equation of a plane through the point P(a, ft 7) and parallel 
to the lines x = ai -f X/s, y = fo -f M*S, z = yi + i/,-s, i = 1, 2. 

6. Prove that the ranks of the coefficient matrix and the augmented matrix 
of the equations of three planes are 2, if these planes belong to the same pencil 
of planes; in proving, use the results of Section 49 and Theorem 14, Chapter I. 

64. Four Planes. Two Lines. The number of possible relative 
positions of a set of planes increases quite rapidly when the set 
contains more than three planes. The methods to be used in such 
cases do not differ in any essential respect, however, from those 
employed in the preceding sections. We shall not study this 
general problem therefore; we shall restrict ourselves to the fol- 
lowing special cases. 

THEOREM 22. Four planes meet In a single point if and only if the 
ranks of the coefficient matrix and the augmented matrix of their 
equations are both 3. 

Proof. If the planes meet in a single point there must be at 
least one set of three among them which form a trihedral angle; 
in that case at least one of the three-rowed minors of the c.m. has 
a value different from zero and therefore the rank of the c.m. is 
3. Moreover, it follows from Theorem 3, Chapter II (Section 23, 
page 38) that the determinant of the augmented matrix vanishes; 
hence the rank of that matrix is less than 4. But since the rank of 
the a.m. can certainly not be less than that of the c.m., the rank 
of the a.m. is 3. 

Conversely, if the ranks of the c.m. and the a.m. are both 3, we 
conclude by means of Theorem 3, Chapter II, that the equations 
have a unique solution, that is, that the planes have a single point 
in common. 

COROLLARY. Two lines meet in a point if and only if the coefficient 
matrix and the augmented matrix of the four linear equations used 
to represent them, both have rank 3. 

THEOREM 23. Four planes have a single line in common if and only 
if the ranks of the coefficient matrix and the augmented matrix of 
their equations are both equal to 2. 



102 PLANES AND LINES 

Proof. If the four planes have a single line in common then every 
set of three of them have at least a line in common, and there must 
be at least one set of three which have but a single line in common. 
It follows therefore from Theorem 20 (Section 51, page 95) that 
the ranks of the c.m. and the a.m. of any three of the four equations 
is at most 2, and that there is one set of three equations among 
them at least, whose c.m. and a.m. both have rank 2. We con- 
clude from this that the ranks of the c.m. and the a.m. of the four 
equations must both be 2. 

Conversely, if the rank of both these matrices is 2, then for any 
three of the four equations the ranks of both the c.m. and the a.m. 
are at most 2, whereas there is at least one set of three equations 
for which the ranks of the c.m. and the a.m. are exactly 2. It 
follows therefore from Theorem 20 that there is at least one set 
of three among the four planes which meet in a line, while the 
fourth plane either passes through this same line, or else is parallel 
to it. The latter alternative is ruled out by Corollary 3 of Theo- 
rem 20, Section 51, page 97. Hence tho fourth plane passes 
through the line common to the other three, as is required by the 
theorem. 

THEOREM 24. Four planes coincide if and only if the ranks of the 
coefficient matrix and the augmented matrix of their equations are 1. 

The proof is left to the reader. 

From Theorems 22, 23, and 24 we obtain immediately the fol- 
lowing corollaries: 

COROLLARY 1. Four planes have one or more points in common if 
and only if the ranks of the coefficient matrix and the augmented 
matrix of their equations are equal. 

COROLLARY 2. Four planes have no points in common if the aug- 
mented matrix of its equations is non-singular. (Compare Definition 
III, Chapter II, Section 26, p. 43.) 

We return now to the Corollary of Theorem 22. The criterion 
for deciding whether two lines meet in a point, which it supplies, 
is not very convenient if the parametric forms of the equations of 
the lines are used. We shall therefore develop this criterion in a 
different form. 

Let us consider the lines 

h : x = ai + Xis, y = ft + ins, z = 71 + vis, 



TWO LINES 



103 



and 



I X = 



= 72 



is 



Whether or not these lines have one or more points in common 
depends upon whether or not it is possible to determine one or 
more values of s and ,s' such that 

that is, upon whether or not the system of equations 

(1) XiS \2$' + Oil <*2 ~ 0, Ml* jJL'rt' + ft ~ ft = 0, 
/ I f\ 

v\s v%8 ~r Ti T2 = " 

possesses solutions. 

The c.m. of this system of equations i 

Xi > 

Mi A 

Vi V 

its rank is 2, unless the lines are parallel or coincident. We con- 
clude therefore, on the basis of Theorem 3, Chapter II (Section 
23, page 38), that if the lines are neither parallel nor coincident 
they will have one point in common or none according as the 
rank of the augmented matrix of the system (1), that is, of the 
matrix 

Xi \2 Oil Oi2 

(3) Ml M2 ft ft 
v\ V<L 71 72 

is 2 or 3. If two non-parallel lines have no point in common, 
they are said to lie skew with respect to each other. 

If the rank of the matrix (2) is 1, the lines are parallel or coin- 
cident; and to distinguish between them we have to consider the 
rank of the matrix (3). If this is 1, all its rows are proportional, 
so that if any one of the equations is satisfied, the other two will 
also be satisfied. This means that to any value of s there corre- 
sponds a value of s' such that together they will satisfy the equa- 
tions (1). Hence every point on l\ coincides with some point on 
Z 2 ; in other words, the lines l\ and Z 2 coincide. If the rank of the 
matrix (3) is 2, there is at least one pair among the equations 
which do not possess a solution, in virtue of Theorem 3, Chapter 



104 



PLANES AND LINES 



II; in this case therefore the lines can have no point in common. 
Finally, it is clear that the rank of (3) can not be 3 if the rank of 
(2) is 1 ; for in that case all the cofactors of the elements in the 
last column of (3) vanish. 

We summarize the conclusions in a theorem. 

THEOREM 25. The two lines * = + \ts 9 y = fo -f #, * = yi -f vis 
are skew if and only if the rank of the matrix (3) is 3; they meet in a 
point if and only if the ranks of the matrices (2) and (3) are both 3; 
they are parallel if and only if the rank of the matrix (2) is 1, while the 
rank of the matrix (3) is 2; they are coincident if and only if the ranks 
of the matrices (3) and (3) are both 1. 

Remark. It should be clear that similar conclusions are ob- 
tained if the equations of the lines are taken in the parametric form 
of Theorem 16 (Section 47, page 86). 

Examples. 

1. To find the relative position of the lines l\, given by the equations 
2 2 y -}- 3 z + 4 = 0, x -}- 2 y z 3 = 0, and / 2 , given by the equations 
3x + 2?/ 2z-|-5=0, 2 x 3 y -}- z 4 = 0, we begin by finding the 
direction cosines of each. We find, by use of Theorem 17 (Section 47, page 87) 
that 



-1 3 

2 -1 

2 -2 

-3 1 



13 -2 

2 1 



1:1, and 



3 2 
2 -3 



4:7: 13. 



It is evident from these results that the lines are neither parallel nor co- 
incident. To decide whether or not they are skew, we determine a point on 
each of the lines. The point PI( 2, 3, 1) lies on \i\ the point P 2 (l, 1, 5) 
lies on Zg. And the determinant 



-1 4 -3 
1 7 2 
1 13 -4 



-1 4 -3 
11 -1 

17 -7 



=4=0. 



Hence the lines are skew. 

2. The lines h : x = -4 -f t, y = 3 - 2 1, z = 2 + 3 t and 1 2 : x -2 
4. 3 i' t y = 1 6 t', z = 8 + 9 /' are parallel or coincident, since their 
direction cosines are proportional to each other. The augmented matrix 
of the equations *-3*'-2=0, -2i-t-6i / -f4=Oand3-9< / -6 = 

1 -3 -2 
2 6 4 

3 -9 -6 
two lines coincide; 
over into those of fe- 



s 



; and we see by inspection that its rank is 1. Hence the 
the substitution I =* 3tf + 2 carries the equations of l\. 



EXERCISES 105 



55. Exercises. 



1. Show that the four planes 3x -\- y z 5 = 0, a; 2y + 3z 2 = 0, 
2 # + 4 ?/ 5 z 2 = 0, and 4 s + 3 y 7 2 + 3 =0 meet in a point. 
Determine the coordinates of this point. 

2. Show that the four planes 3z-?/ + 2z-3 = 0, 2x + 2y-3z 
+ 4=0, 3 + 5 j/ -82 + 11=0 and Sy - 13 2 + 18 = meet in a line. 
Determine the direction cosines of this line. 

3. Show that the lines .c = 4 2 1, y = 3 + 2 , 2 = 5- 3 t and 
x = t, y = I 41, z 1 + 3 meet in a point. Determine the co- 
ordinates of this point. 

4. Determine whether the lines # = 5 3, 2/ = 4-M, 2= 3 + 4 tf and 
x = Q 6 t, y 2 + 2 <, 2 = 5 + 8 are parallel or coincident. 

5. Prove that if four planes form a tetrahedron, the ranks of the coefficient 
matrix and the augmented matrix of their equations are 3 and 4 respectively. 

6. Prove that if four planes form a four-sided prism, the ranks of the 
coefficient matrix and the augmented matrix of their equations are 2 and 3 
respectively. 

7. Prove analytically that if two lines are parallel there exists one and only 
one plane in which they both lie. 

8. Remembering that skew lines are non-parallel lines which have no point 
in common, prove that if two lines are skew, there exists no plane in which they 
both lie. 

56. Miscellaneous Exercises. 

1. Determine the distance of the point A (2, 3, 8) from the line I : x = 3 
2 t, ?/ = 1 + 2 /, z = 6 t. Hint: The distance from the line to A 
is equal to the product of the distance from A to the point #(3, 1, 6) on the 
line by the sine of the angle between AB and /; use Theorem 14, Chapter III, 
(Section 36, page 64). 

2. Determine the distance of the point P(a\ t ft, 71) from the line x = a 
+ As, y = + s, z = 7 -f vs. 

3. Find the distances of the point A (4, 5, 3) from the planes 2 x 6 y 
32 + 4 = and 3x -{- Qy 2z 5=0 and from their line of intersection. 

4. Determine the relative positions of the planes in the following sets: 

(a) 2z-?/4-3z-4=0, 3x + 2y - z + 2 = 0, s-42/ + 7z-10 = 0; 
(6) 2z-2/ + 3z-4 = 0, 3x + 2y-z + 2=0, -42/4-72-6=0. 

6. Find the points which 

(a) the plane 3x 2y + z +2 = and the line x = l-\-t,y= 2 

+ 2t, z = -3+4*; 

(b) the plane 3 x - 2 y + z + 2 = and the line x = -1 + 2 *, y = 2 

+ 2J, z = 3-2 t-, 

(c) the plane 3 x - 2y + z + 2 = and the line x=-l+ t, y = 2 

+ *, z = -3 - / 
have in common. 



106 PLANES AND LINES 

6. The intercepts of a plane are a, 6, and c. On the axes of another rec- 
tangular reference frame with the same origin as the original axes, the inter- 

cepts of the same plane are a\, 61, and d. Show that -$ + r^ + -} = ~~2 

CL 0" C Q i 



7. Show that the locus of all points which are equally distant from two 
given planes consists of two planes through the line of intersection of the 
given planes. For the points on one of these planes the distances from the 
two given planes are equal in sign as well as in magnitude; for those on the 
other plane the distances from the two given planes are equal in magnitude, 
but opposite in sign. 

8. Prove that the two planes determined in Exercise 7 are perpendicular to 
each other. 

9. Prove that each of the planes determined in Exercise 7 makes equal 
angles with the two given planes. The planes found in Exercise 7 are called 
the bisecting planes of the dihedral angle formed by the two given planes. 

10. Prove that if three planes meet in a point, the six bisecting planes of 
the three dihedral angles formed by them meet three by three in a line. Hint: 
Take the equations of the given planes in the normal form. 

11. Write the equation of the plane through the origin determined by the 
two lines through the origin whose direction cosines are Xi, MI, "i and X 2 , ^2, v*. 

12. Show that three concurrent lines are coplanar (lie in one plane) if 
and only if the determinant formed by their direction cosines vanishes. (Com- 
pare Exercise 5, Section 42, page 74.) 

Note. The determinant, mentioned in this exercise, whose rows consist 
of the direction cosines of three concurrent lines, will be called the ori- 
entation determinant of these lines. 

13. A point moves in such a manner that its distances from two fixed lines 
are always equal to each other. Determine the equation of the locus which 
this point describes. 

14. Determine the equations of the line which passes through the point 
Pi(<*i, ft, 7i), is perpendicular to the line joining P 2 ( 2 , ft, 72) and I\(a z , ft, 73), 
and lies in the plane determined by the points PI, P 2 , and P 3 . 

15. Determine the distance of the point PI(I, ft, 71) from the line joining 
the points P 2 (a 2 , ft, 72) and Ps(aa, ft, 73). 

16. Show that four times the square of the area of the triangle whose ver- 
tices are Pi(ai, ft-, 7;), i = 1, 2, 3 is equal to the sum of the squares of the 



determinants 



7i 



1 



ft 72 1 
1 



73 



73 3 



<*2 



17. Find the distance of the point P(a, ft 7) from the plane determined 
by the points Pi (a/, ft', 71), i = 1, 2, 3. 



EXERCISES 



107 



18. Prove that the volume of the tetrahedron whose vertices are the points 
Pi(<*i, fa, yi) i 1, 2, 3, 4 is equal to one sixth of the value of the determinant 

1 /3i 71 1 

2 & 72 1 
/?3 73 1 

* Pi 74 1 

19. Determine the equations of the three planes, each of which passes 
through one of three concurrent edges of a tetrahedron and is perpendicular 
to the face opposite to the vertex in which these edges meet. 

20. Show that the three planes determined in the preceding exercise meet 
in a line which passes through one vertex and is perpendicular to the opposite 
face. 

21. Prove that the four perpendiculars from the vertices of a tetrahedron to 
the opposite faces meet in a point. 

22. Determine the equation of the plane through one edge of a tetrahedron 
and through the midpoint of the opposite edge. 

23. Prove that the six planes of the kind described in the preceding exercise 
meet in a point. 

24. Determine the equation of the plane through the midpoint of one edge 
of a tetrahedron and perpendicular to the opposite edge. 

26. Prove that the six planes of the kind described in the preceding exercise 
have one point in common. 



CHAPTER V 
OTHER COORDINATE SYSTEMS 

In the development of the subject up to this point we have used 
a rectangular Cartesian coordinate system; and we have had no 
occasion to change from one such system to another. Many of 
the problems to be taken up in later chapters require such tran- 
sitions; and for other purposes it is frequently desirable to use 
reference frames different from that furnished by the rectangular 
Cartesian coordinates. We shall therefore consider in the present 
chapter some other reference frames in three-space, and also the 
transition from one reference frame to another. 

67. Spherical Coordinates. The reference frame consists of: 
(1) a fixed plane II; (2) a fixed half -line, I, in this plane, called the 

initial line ; (3) a fixed point 
on the line /, called the origin; 
and (4) a unit for linear meas- 
urement and a unit for angular 
measurement. To determine 
the coordinates of a point P in 
space with reference to this 
FIG. 11 frame, we connect with P and 

we drop a perpendicular from P 

to the plane II; let P' be the foot of this perpendicular (see Fig. 11). 
The spherical coordinates of P are then defined as follows. 

DEFINITION I. The spherical coordinates of a point P in space are: 
(1) the unsigned distance r from O to P, measured in terms of the unit 
specified for linear measurement this is called the radius vector of 
P; (2) the angle between -90 and 90 which the plane II makes with 
OP, measured in terms of the unit specified for angular measurement 
this is called the latitude; and (3) the angle e between and 360 
which the line I makes with the projection of OP on the plane n, 
measured in terms of the same unit this is called the longitude. 

Remark 1. It follows from this definition that to every point 
in space, except 0, there corresponds a definite set of three real 
numbers, which are its spherical coordinates. But it is not true 
in this case, as it was when Cartesian coordinates were used, that' 

108 




SPHERICAL COORDINATES 



109 



for every set of three real numbers there exists a point of which 
these numbers are the spherical coordinates. The radius vector 

is an unsigned real number, the latitude must lie between - 

i 

and x , and the longitude between and 2 ?r, if the radian is the 
J 

unit of angular measurement (between 90 and 90, between 
and 360 if the degree is the unit). 

Remark 2. There are various ways in which the definition of 
spherical coordinates may be modified. The radius vector may 
be defined as a signed number, with a possibility of its being either 
positive or negative ; and the ranges of value for the latitude and 
the longitude may be changed. Although there are some advan- 
tages to be derived from such different agreements which the co- 
ordinates, as defined above, do not possess, the present definition 
has the desirable property, mentioned in Remark 1, of assigning to 
every point in space, except 0, a single set of spherical coordinates. 




FIG. 12 



To establish a connection between spherical coordinates and 
rectangular Cartesian coordinates, we make the point the origin 
of a rectangular reference frame, the line I the positive X-axis 
and the plane n the -XT-plane. Moreover we adopt the unit of 
linear measurement as the unit on the three axes of this super- 
imposed rectangular frame. It should now be easy to see (from 
Fig. 12) that the Cartesian coordinates of a point P are connected 
with its spherical coordinates by the following formulas: 

x = OP X = OP' cos = r cos </> cos 0, 
y = p x p f = OP' sin = r cos sin 0, 
z = p'p = r sin <. 



110 OTHER COORDINATE SYSTEMS 

If these equations are squared and then added together, we find, 
in accordance with the conventions laid down in Definition I, that 

r = \Vx 2 + y' 2 + z 2 \ and hence that <t> = Arc sin. , 

' ' ^ 



+ I/ 2 + 2 2 

Squaring and adding the first two equations leads to r cos < = 

, . _ , y , 

|vV 2 + ?/ 2 |, and hence to the result that sin 6 = i / 2 . f[ alKl 

# 

cos = j . ; by means of these conditions the angle 

|vV + j/2| 

is completely determined between and 2 w. We state our results 
as follows. 

THEOREM 1. The transformation from spherical coordinates to rec- 
tangular Cartesian coordinates, and vice versa, when in the two 
reference frames the origins coincide, the initial line coincides with 
the positive half of the X-axis, the initial plane with the XI -plane, 
and the units of linear measurement are the same, is accomplished 
by means of the equations: 

x = r cos < cos 0, y r cos </> sin 0, z r sin </>; 
and 



4- Z 2 | f = Arc sin -. 



sin e = ; -., cos e = 

I ^x' 2 -f y 2 \ 

By means of the first set of formulas an equation in Cartesian 
coordinates may be transformed into an equation in spherical 
coordinates; and the second set of formulas enables us to trans- 
form an equation in spherical coordinates into an equation in 
Cartesian coordinates. In view of Definitions I and II of Chap- 
ter IV (Section 39, page 68), we conclude from this that the 
locus of a single equation in spherical coordinates is a surface, and 
the locus of a pair of equations a curve. The geometrically 

* We are using here the notation Arc sin u to indicate the " principal value " 

of the angle whose sine is u, that is, the angle between and ~ which has 

Z Z 

its sine equal to u\ this function has a single real value for every real value 

of u between 1 and 1. It should be clear that -. , is never 

\Vx* + y* + z>\ 
more than 1 and the angle <t> as defined in the text always exists and lies between 

7T j 7T 




CYLINDRICAL COORDINATES 111 

simplest surfaces, that is, the planes, are not represented by the 
algebraically simplest equation when spherical coordinates are 
used. By using Theorem 1, in connection with Theorem 4, 
Chapter IV (Section 41, page 71), we find that the general 
equation of a plane in spherical coordinates is 

r(a cos </> cos + 6 cos <j> sin 6 + c sin 0) + d = 0. 

On the other hand, the equation of a sphere whose center is at 
the origin and whose radius is a has, in spherical coordinates, the 
very simple equation r a. 

58. Cylindrical Coordinates. 
The reference frame now con- 
sists of the initial plane II, the 
initial line I, the origin 0, units 
of linear and of angular meas- 
urement, and besides of a di- 
rected perpendicular to the 
plane II at 0, called the z?-axis. 
The cylindrical coordinates of 

an arbitrary point P in space are then defined as follows (see 
Fig. 13). 

DEFINITION II. The cylindrical coordinates of a point P in space are: 
(1) the perpendicular distance, f , from the initial plane to P, measured 
in accordance with the unit of measurement and the direction speci- 
fied for the -axis; (2) the undirected distance, p, from O to the pro- 
jection P' of P on the initial plane, measured in terms of the specified 
unit of linear measurement; and (3) the angle ^, between and 360, 
which the initial line I makes with OP', measured in terms of the 
specified unit of angular measurement. 

Remark. The cylindrical coordinates of a point evidently com- 
bine polar coordinates in the initial plane with a Cartesian f- 
coordinate. For every point in space, except those which lie 
on the 2-axis, there exists a unique set of 3 real numbers, which are 
its cylindrical coordinates. But again it is not true that with 
every set of three real numbers there is associated a point of which 
these numbers are the cylindrical coordinates. The f -coordinate 
is a signed real number, the coordinate p an unsigned real number, 
and the coordinate ^ is restricted to the range 2 TT, if the radian 
is the unit of angular measurement. 



112 OTHER COORDINATE SYSTEMS 

A rectangular Cartesian reference frame can be superimposed 
on the reference frame used for cylindrical coordinates in the 
manner used in the preceding section for spherical coordinates. 
Thus we find the following relations between the rectangular 
Cartesian coordinates z, y, z of a point and its cylindrical co- 
ordinates p, \p, f (see Fig. 13) : 

x = OP X = p cos ^, y = P X P' = p sin \f/, z = f ; 
and 



,-=*, . x- =-,, . 

\Vx 2 + y 2 \ \Vx 2 + ?/| 

Moreover, the reference frame for cylindrical coordinates con- 
tains a reference frame for spherical coordinates. It is therefore 
a simple matter to connect the spherical coordinates r, 0, of a 
point with its cylindrical coordinates p, ^, f. We find: 

p = OP 1 = r cos 0, ^ = 0, f = r sin <; 
and 

r = |Vp' 2 + H, * = Arc tan , = ^. 

P 

A repetition or the argument made in the last paragraph of 
Section 57 should make it clear that the locus of a single equation 
in cylindrical coordinates is a surface, and the locus of a pair of 
equations a curve. The equation p = a represents a right cir- 
cular cylindrical surface whose radius is a and whose axis is along 
the *?-axis; the pair of equations p = a, f = b determines a circle 
of radius a, in a plane parallel to the initial plane at a distance 6 
from it and having its center on the 2-axis. 

59. Exercises. 

1. Determine the loci of each of the following equations: 

(a)r = 2; (6) P - 3; (c)0 = ~; (</)f=-l; (e) 4 = -|; (/) * = ~ 

2. Write the equations in spherical coordinates of the surfaces whose equa- 
tions in rectangular Cartesian coordinates are: 

(a) x* -f 2/ 2 - 5; (6) y + * = 3; (c) 3 x - 2 y = 0; 
(d) 3 z 2 4- 2 r/ 2 + 4 z 2 = 1; (e) 4 x 2 - y 2 = 1. 

* The notation Arc tan ?z is used to designate the angle between and 

2 2 

whose tangent is u, that is, the " principal value " of the multiple-valued func- 
tion Arc tan u; see the footnote on page 110. 



OBLIQUE CARTESIAN COORDINATES 113 

3. Determine the locus of each of the following pairs of equations: 
(a)r = 3,0 = ~s (6)J"=4,* = *; (c) * = - *, * = ?^; 

(d) P =5, r = -2; W r = 4, <#> = ~- 

4. Transform the following equations into equations in rectangular Car- 
tesian coordinates: 

(a) r = tan d; (b) f = 2 <; (c) r(cos + sin tan 0) = 4 sec 0; 
(df) P (3 cos ^ - sin t) + 2 r - 4 = 0; (e) P 2 + f 2 = 9; 
(/) sin 2 + 2 sin 2 0-3 cos 2 (9=4. 

60. Oblique Cartesian Coordinates. A reference frame for 
a system of oblique Cartesian coordinates is furnished by any 
three planes which meet in a 
point, 0. These planes meet 
two by two, in lines through 0; 
we call these lines the X-, Y-, 
and Z-axes and denote them 
by OX, OF, and OZ respectively 
(Fig. 14). On each axis we 
specify a positive direction and 
a unit of measurement. Through Fro. 14 

an arbitrary point P in space, we 

draw lines parallel to the coordinate axes, meeting the given 
planes in the points P yz , P^, and P xy . We can now give the 
following definition. 

DEFINITION III. The (oblique) Cartesian coordinates of the point 
P are the lengths of the lines PyzP, PzxP, and P xy P measured in accord- 
ance with the units and directions specified for the X-, Y-, and Z-aies 
respectively. 

The coordinate planes, together with the planes determined by 
the lines P yz P, P&P, and P xy P, taken two at the time, form an 
oblique parallelepiped. This parallelepiped can be used con- 
veniently to develop generalizations of some of the results obtained 
in Chapter III, just as these results themselves were found by the 
aid of the rectangular parallelepipeds, which we designated as the 
c.p. of a point and the c.p. of a pair of points (see Sections 30 and 
32). 

Notation. The coordinate frame which we have just described 
will be designated by the symbol 0-XYZ. If the angles between 




114 



OTHER COORDINATE SYSTEMS 



the coordinate axes have to be specified, we shall use the symbol 
O-XYZ-aQy, where a = / YOZ, ft = zZOX, and y = Z.XOY. 

It will be supposed throughout that the units of measurement 
on the three axes of any Cartesian reference frame, and also those 
used on the axes of two such frames whose mutual relations are 
under consideration, are equal to each other. 

61. Translation of Axes. We consider now the relations exist- 
ing between the two sets of coordinates of a point P with reference 
to two Cartesian reference frames, whose axes are parallel; these 
may be rectangular or oblique frames. Let the two reference 




FIG. 15 

frames be 0-XYZ and O'-X'Y'Z', and let the coordinates of 0' 
with respect to 0-XYZ be a, b, and c. For an arbitrary point P(x, 
y, z) we construct now the c.p. of P and 0' with respect to 0-XYZ. 
Since the axes of the two frames are parallel, this parallelepiped 
will also be the c.p. of P with respect to O'-X'Y'Z' and therefore 
its edges will be equal in unsigned length to the numerical values 
of the coordinates of P with respect to O'-X'Y'Z', that is, of 
x', y 1 , and z'. The X-, F-, and Z-axes will meet the faces of this 
parallelepiped in the points P x , <V; P y , O/; and P 2 , 0,' respec- 



OBLIQUE TO RECTANGULAR AXES 115 

lively; and the segments O X 'P X , O y 'P y , and 0/P 2 are equal to 
x', y', and z r respectively (see Fig. 15). We have now, independ- 
ently of the positions of P and of the reference frame O'-X'Y'Z', 
the following relations: 

0/0+ OP X + P X X ' = 0, 0/0 + OP y + PyO y ' = 0, and 
0/0 + OP Z + P 3 2 ' = 

and therefore 

a + x x' = 0, b + y y' Q, and c + z z' = 0. 

The result of the discussion can be summarized in the following 
theorem : 

THEOREM 2. The coordinates *, y, z of an arbitrary point P with 
reference to a Cartesian frame of reference O-XFZ, and the coordi- 
nates *', y' 9 z' of the same point with reference to a parallel Cartesian 
frame of reference O'-X'Y'Z', whose origin has the coordinates a, 6, c 
with respect to O-XYZ satisfy the relations 

x' ~ x a, y' y 6, s' = z c. 

Remark 1. The coordinates #', y' y and z f of the point 0' are all 
0; hence we find from the theorem just stated, that the coordinates 
of 0' with reference to O-XYZ are a, 6, c, as stated in the hypothe- 
sis of the theorem; this simple fact serves as a check on the 
formulas. Similarly, we find that the point whose coordinates 
in the system O-XYZ are (0, 0, 0) has the coordinates ( a, 6, 
-c) in the system O'-X'Y'Z'. 

Remark 2. It should be noted that the formulas established in 
Theorem 2 are the same, independently of whether the two refer- 
ence frames are oblique or rectangular. 

62. Transformation from Oblique to Rectangular Axes. Before 
taking up the transformation of coordinates which results when we 
pass from one arbitrary frame of reference to another, we shall 
consider what happens when we change from an oblique frame of 
reference to a special associated rectangular frame. A system of 
rectangular axes can be superimposed upon a given oblique refer- 
ence frame 0-XYZ-af$y by using in both systems the same XY- 
plane, the same X-axis, the same origin, and the same units of 
measurement. This is illustrated in Fig. 16, in which the axes of 
the rectangular frame are designated by OX', OF', and OZ'. To 
obtain the relations between the coordinates of an arbitrary point 



116 



OTHER COORDINATE SYSTEMS 



P with respect to these two frames of reference, we make use of 
the projection method (see Section 36). The coordinate parallele- 
pipeds of P with respect to these two sets of axes furnish a closed 
broken line, leading from to P along the edges OP*, P x P xy , P xy P, 




X,X l 



and back from P to along the edges PP xy ', P xy 'P x , P X 0. We 
infer now from Theorem 12, Chapter III (see Section 36, page 62) 
that 



P,'0 = 0. 

To evaluate these projections, we make use of Theorem 11, Chap- 
ter III (see Section 36, page 62), remembering that the angles 
YOZ, ZOXj and XOY formed by the original axes are equal to 
a, 0, and 7 respectively, and that the X', Y', and Z' axes are mu- 
tually orthogonal. In this way we find that 

x + y cos 7 + z cos )8 x f 0. 

By projecting the same path OP x P xy PP xy P x O upon the F'- and 
Z'- axes, we find 

y cos z YOY' + z cos z ZOY' - ?/ = and 
2 cos z ZOZ' - z' = 0. 

Clearly z F07' = ~ - Z-YOF, so that cos z F07 X = sin y. 

t 



To determine cos ^ZOY f and cos z^OZ', we make use* of the 
result of Exercises 15 and 16, Section 38 (page 66), from which 
we find that 

cos a. cos /5 cos 7 



sin 7 



and 



cos Z ZOZ' = -. Vl cos 2 a cos 2 ft cos 2 7 + 2 cos a cos cos 7, 



sin 7 



OBLIQUE TO RECTANGULAR AXES 117 

in which the + or sign is to be used according as the Z-axis 
and the Z'-axis point toward the same side or toward opposite 
sides of the .XT-plane. Substitution of these values in the pre- 
ceding equations leads to the following theorem. 

THEOREM 3. If x, y, and z are the coordinates of an arbitrary point 
P in the Cartesian frame O-XVZ-apy, and if *', y', and s f are the coo'r- 
dinates of the same point with reference to the orthogonal Cartesian 
frame O'-X'Y'Z' in which the units of measurement, the origin, the 
A -axis, and the A Y-plane are the same as the corresponding elements 
of the frame O-XYZ-apy, then 

/ , / , , * (COS a COS COS 7) 

* = x 4- y cos 7 + * cos 0, y' = y sin 7 H - - -fa - - - 

f =fc*(l - COS 2 a - COS 2 - COS 2 7 + % COS a COS COS 7)* 

* = - sin^ 

In which the plus or minus sign is to be used according as the two 
reference frames are of the same or of opposite type (see footnote on 
p. 50). 

By means of the formulas of this theorem we can express the 
distance of a point from the origin of a system of oblique axes in 
terms of the oblique coordinates of this point. For 

OP 2 = x'* + y'* + z' 2 

= (x + y cos 7 + z cos /3) 2 + 

S COS0COS7)! 2 

- - 



[ 
y si 



. 

sin 7 + 

sin 7 

2 2 (1 cos 2 a cos 2 ft cos 2 7 + 2 cos a cos cos 7) 

sin 2 7 
= x 2 + y* (cos 2 7 + sin 2 7) 

2 2 [cos 2 /3 sin 2 7 + (cos a cos ft cos 7) 2 + 1 cos 2 a 
__ cos 2 j8 cos 2 7 + 2 cos a cos ft cos 7] _ 

sin 2 7 

+ 2 xy cos 7 + 2 xz cos 
+ 2 2/2 [cos 7 cos j8 + cos a cos jS cos 7] 
= rr 2 + i/ 2 + z 2 + 2 j/2 cos a + 2 zz cos jS + 2 XT/ cos 7. 

COROLLARY 1. The square of the distance from the origin of a ref- 
erence frame O-XYZ-<xpy to a point P(*, y, *) is equal to * 2 4- y 2 + * 2 + 
2 y* cos a -f 2 ** cos -f 2 *y cos 7. 

The expression for cos ZOZ' which was used in the proof of 
Theorem 3 enables us moreover to obtain a convenient formula 



118 



OTHER COORDINATE SYSTEMS 



for the volume of the c.p. of the point P(a, 6, c) in an oblique 
reference frame. For the area of the base of this parallelopiped 
is equal to a&sin 7 and its altitude is c cos ZOZ'. This leads 
readily to the following result. 

COROLLARY 2. The volume of the coordinate parallelopiped of the 
point P(a 9 b, c) In the reference frame O-XYZ-afty Is equal to abc (1 

1 cos 7 cos ft 



COS 2 a COS 2 ft COS 2 7 -h 2 COS a COS ft COS 7]* = abc 



COS 7 

COS ft 



1 

COS a 



COS a 

1 



63. Rotation of Axes. The projection method also enables us 
to determine in a direct manner the relations which connect the 
coordinates of one point in two Cartesian reference frames which 
have the same origin. We shall first develop these equations for 
the general case in which both systems are oblique and then ob- 
tain as a special case the formulas for two rectangular systems. 

Let the systems be 0-XYZ-a(5y and 0-X^YiZr-arfiy^ the 
units being the same in the two. Let the cosines of the angles 
formed by the axes of these systems be indicated in the following 
table: 





X 


Y 


Z 


X, 


h 


nil 


HI 


Y l 


k 


W 2 


n* 


Zi 


h 


W.3 


n.s 



so that we have cos XOXi = h, cos XOYi = k, cos ZOYi = HZ, 

etc. 

We consider now the closed broken 
line which leads from to P along 
the edges OP*, P x P xy and P xy P of the 
c.p. of P in the system 0-XYZ, and 
which returns from P to along the 
edges PP* iyi , PxwPxi, Pxfl of the c.p. 
of P in the system 0-X \YiZi (see 
Fig. 17). We project this closed 
broken line in turn on the axes OX, 
p IG 17 OYj and OZ, and then on the axes 

OX ly OYi, OZi. If we make use in 

each of these projections of Theorems 11 and 12 of Chapter III 
(see Section 36, page 62) and if we employ also the notation in- 




ROTATION OF AXES 



119 



troduced above for the angles between the two sets of axes, we 
find: 

Ix + y cos 7 + z cos ft - xili - 7/1/2 - Zi/ 3 = 0, 
x cos 7 + y + z cos a x\m\ yim% z\m^ = 0, 
x cos |3 + y cos a + 2 Zi/ii 7/1*12 Zin 3 = 0; 
and 

Iz/i + ymi + zni Xi 7/1 cos 71 Zi cos ft = 0, 
xk + ym 2 + znz Xi cos y L yi z\ cos a t = 0, 
xl s + 7/m 3 + zn 3 Xi cos ft ?/i cos i z\ = 0. 

The system of equations (1) has a unique solution for x, y, z in 
terms of Xi, yi, z\\ and the system (2) has a unique solution for 
x\, 7/1, Zi in terms of #, T/, z. These solutions may be written down 
by means of Cramer's rule (see Section 21, p. 37). For the coeffi- 
cient determinants of these systems are oqual respectively to 



1 cos 7 cos 
cos 7 1 cos a 
cos j3 cos a. 1 



and 



1 

cos 71 

COS ft 



COS 71 COS ft 
1 COS Oil 

COS Oil 1 



and, by virtue of Corollary 2 of Theorem 3 (Section 62, page 118), 
the values of these determinants are equal to the squares of the 
volumes of the c.p/s of the points (1, 1, 1) in the two systems. 
If these volumes are denoted by v and vi respectively, we find from 
the first system, that 







Iixi+I 2 yi-+ 


Itfi 


cos 7 


cos/? 




X 


= X 


niiXi+m 2 y 


i+w 3 2i 


1 


cos a 


= 




V 


* 














niXi+n 2 yi 


+ ^3^1 


cos a 


1 




where 












i 


li COS 7 


COS 




1 


^2 


n 


= X 


?>l! 1 


COS a , 


012 = 


_ v/ 


m 2 




V 








z; 








HI cos a 


1 






Ut 




1 


Zs COS 7 


COS 






#13 


X 


m 3 1 


cos a 


J 






V 


n 3 cos a 


1 







cos 7 cos 
1 cos a 
cos a 1 



(see Theorem 8, Chapter I, Section 5, page 9). Similar results 
are obtained for y and z, the trinomial elements now appearing in 
the second and third columns. And from the system (2) we obtain 



120 



OTHER COORDINATE SYSTEMS 



the solution 



yi = - x 


I 


lix+miy+niz cos 
l^x-^m^y-i-n^z cos 


Vi 


COS ft 


hx+fthy+ 


tt;{2 1 


where now 






1 

'''21 "~ /N 


1 

COS 7! 


h eos ft 

/2 COS i 


] 


1 


COS ft 


/3 1 


V 


fc, - 1 X 


1 Hi 

cos 7! ?1 2 


cos/3i 
cos ai 


Vi 


cos ft rig 


1 



1 

cos 71 

cos/3i 


77? 1 
?/?2 

m 3 


cos |9i 

COS i 

1 



, with similar results for 



x\ and 2i. The reader should have little difficulty in writing out 
these further results. We state the following conclusion of our 
discussion. 

THEOREM 4. If *, y, z are the coordinates of an arbitrary point P 
with respect to a reference frame O-XYZ-apy, and *i, y lf *i are the 
coordinates of the same point in the reference frame O-XiYiZi-aifayi, 
of which the A>, !>, and Ziaxes make with the axes of O-XYZ angles 
whose cosines are /i, mi, /ii; / 2 , m 2 , n 2 and /a, m 3 , n 3 respectively, then 



x a\\x\ 
y = 021*1 
s = 031*1 



> and 



-f 
4- 



-f 



Here aij(i,j = 1, 2, 3) is the value of the determinant obtained 

1 cos 7 cos ft 



from the determinant v = 



cos 7 



1 



cos a 



by replacing the 



cos ft cos a 1 
ith column by //, wj,-, n,-, divided by the value of v] and the co- 
efficient by(i 9 j = 1, 2, 3) is the quotient by the value of the deter- 
1 cos 71 cos ft 



minant v\ = 



cos 71 



cos ot\ 



of the determinant obtained 



cos ft cos i 1 
from Vi by replacing the ith column by the jih column of the de- 



terminant 



Remark. The reader should write out in full the values of the 
coefficients a# and &# in the form of the determinants, which in 



ROTATION OF AXES 121 

order to save space have merely been described in the statement 
of the theorem. 

The formulas established in this theorem take a particularly 
simple form in case both reference frames are rectangular. For 

in that case, a = |ft = 7 = ai = /3i = 7i=~, hence cos a = cos ft 

2t 

= cos 7 = cos ai = cos ft = cos 71 = 0; moreover the numbers 
h> mi) ni] lz, nit, nz and Z 3 , 7/13, n^ become the direction cosines of 
the axes OXi, OY lf and OZi with respect to 0-XYZ respectively. 
The reader should have no difficulty in obtaining the formulas 
for this special case which are stated in the following theorem; these 
formulas, more than those of Theorem 4, arc the ones which we 
shall have frequent occasion to use in our further work. 

THEOREM 5. If x 9 y, z are the coordinates of a point P with respect 
to a rectangular Cartesian frame of reference O-XYZ, and x l9 y\ 9 z\ the 
coordinates of the same point with respect to another rectangular 
frame O-XiYiZ i9 of which the Xi-, Yr, and Zi-axes have in O-XYZ 
direction cosines \i, MU "i ^2, M2 ^ and X 3 , MS v* respectively, then 

x = \ixi 4- X 2 yi 4- X 3 *i, and XL = Xi* + my -f ns 9 
y = MI^I + M2ji + M3*i, y\ = X-2^ -f May H- 2S 9 

x = VLXI + ^ 2 yi 4- ^si; -i = Xa^ 4- May 4- ^* 

64. Exercises. 

1. Set up the equations for the transformation of coordinates resulting from 
translating the axes to a new origin whose coordinates in a system O-XYZ 
are -3, 5, 2. 

2. Determine the equation of the sphere x' 2 4" 2/ 2 + 2 2 = 9 with respect 
to a new frame of reference obtained by translating the original axes to the new 
origin O'(-2, -1,3). 

3. Show that the planes determined by the equations 3 x 6?/4-2z=0, 
2 x 4- y and 2 x 4 y 152 = are mutually perpendicular, and that 
they pass through the origin. Establish the formulas for the transformation of 
coordinates which results when these planes are taken respectively as the 
FI#I-, the ZiXi-, and the XiFi-pIane of a new frame of reference. 

4. Solve the same problem for the planes 2 x //4-2z = 0, x 2y 
20 = 0, 2x + 2y -z = 0. 

6. Apply the formulas obtained in Exercise 4 to determine the equations 
in the new reference frame of the loci of the following equations: 

(a) 2/ 2 4- z 2 = 3; (6) x 2 + y 2 + z 2 = 4; (c) 2 x 2 - 5 ?/ - 4 z 2 = 10; 
(d) ax 4- 6?/ -f cz 4- d =* 

6. Express the distance between two points in terms of their coordinates in 
a system of oblique axes. 



122 OTHER COORDINATE SYSTEMS 

7. Determine the volume of the coordinate parallelepiped of two points in 
an oblique frame of reference. 

8. Prove that, if t , 2 , and 3 are the angles which a line I makes with the 
axes of a reference frame 0-XYZ-apy, then 

1 cos y cos cos O l 
cos 7 1 cos a cos 2 
cos /3 cos a 1 cos 3 

COS 0i COS 2 COS 03 1 



0. 



Hint: If I coincides with one of the coordinate axes, the formula can 
readily be verified. If I does not coincide with any of the axes, take a point P 
on J, so that OP = 1 and project the closed broken line OP x P xy PO on the axes 
and on /; from the resulting equations the desired formula should follow. 

9. Show that, in case a. = 3 y = - , the formula of the preceding exercise 

& 

reduces to that given in Theorem 7, Chapter III (Section 33, page 56). 

10. Show that if the formula of Exercise 8 reduces, for every line Z, to the 
formula of Theorem 7, Chapter III, then a. = p = y = * 

i 

65. Rotation of Axes, continued. The formulas obtained in 
Theorems 4 and 5 appear to contain a large number of parameters; 
but these are not all independent parameters. For the a# and by 
of Theorem 4, and the \, ^ and ^ of Theorem 5 can not be chosen 
arbitrarily if the formulas are to represent a rotation of axes. 
This can be seen most readily if we observe that the expressions 
for the distance from to an arbitrary point P should be the same 
in any two frames of reference which have the same origin. Hence 
it follows that if the parameters in the formulas of Theorem 4 are 
properly selected, then we must have, in view of Corollary 1 of 
Theorem 3 (Section 62, page 117): 

x 2 + y 2 + z 2 + 2 yz cos a + 2 zx cos ft + 2 xy cos 7 

= Zi 2 + y\ 2 + Zi 2 + 2 y^i cos 71 + 2 z&i cos ft + 2 x^ cos 71 

for all values of z, y, and z, if for xi, yi, and z\ we substitute the 
expressions given in Theorem 4. If, in particular, both reference 
frames are rectangular, we find by using Theorem 5 that 

(XiX + Mi?/ + viz)* + (X 2 x + M 2 y + v&Y + (X 3 X + my + v&) 2 
= x* + y* + z* 

for all values of x, y, and z. 

If we carry out the squaring of the trinomials on the left-hand 
side, and equate the coefficients of like terms on the two sides, we 



ROTATION OF AXES 123 

are led to the conclusion that, if the formulas of Theorem 5 do 
indeed represent a rotation of a rectangular reference frame, then 
the parameters X,-, M, v^ i = 1, 2, 3, must satisfy the following 
relations : 

Xl 2 + X 2 * + X 3 2 = 1, Ml 2 + M2 2 + M3 2 = 1, *1 2 + "2 2 + "3 2 = 1 

and MI^I + ^2^2 + M3^3 = 0, v\\i + ^2X2 + ^3X3 = 0, 



+ X 2 M2 + X 3 /U :i = 0. 

Conversely, in case these conditions hold, the equations of 
Theorem 5 will carry a given rectangular frame over into another 
rectangular frame with the same origin. For, by virtue of the first 
three of the above relations, we can then take Xi, X 2 , Xs; jui> M2, Ms> 
and *>i, j> 2 , J>3 as the direction cosines of three lines through the 
origin; and it follows from the last three relations that these 
lines are mutually perpendicular. The equations of Theorem 5 
represent then the transformation to the new rectangular refer- 
ence frame of which these three lines are the axes. Hence we have 
established the following theorem. 

THEOREM 6. If *i, y l9 i represent the coordinates of a point with 
respect to a rectangular reference frame, the necessary and sufficient 
conditions that the equations 

xi = Xi* 4- my + viz, yi = \& -f M2y + v>s>> s\ = \& -f May + "a* 

shall represent a transformation to another rectangular reference 
frame, are that 

Xi 2 + X 2 2 + X3 2 = 1, Mi 2 -f M2 2 -f Ma 2 = 1, vi* + *2 2 + "3 2 = 1 

and that 

Ml*'! H~ M2^2 ~f" M3**3 = 0, t>i\l ~f- J/2^2 ~f ^3X3 = 0. XiMl + ^2^2 4" XajUS = 0. 

Remark 1. A transformation which satisfies the conditions of 
Theorem 6 is called, with obvious justification, an orthogonal 
transformation. 

Remark 2. With the aid of Theorem 6, it becomes easy to verify 
that the two sets of equations in Theorem 5 are equivalent. For 
if we multiply those of the second set by Xi, X 2 , X 3 respectively and 
then substitute the results in the first equation of the first set, we 
find: 

x = Xi( 
= 

= X] 



124 



OTHER COORDINATE SYSTEMS 



and the other equations of the first set are verified in similar 
manner. 

Remark 3. It should be clear that the conditions of Theorem 
6 can also be put in the equivalent form : 

Xi 2 + Mi 2 + vi 2 = X 2 2 + M2 2 + *2 2 = A 3 2 + M3 2 + "3 2 = 1 
and 



+ iw = 0. 

The equivalence of the two sets of equations in Theorem 5, 
observed in Remark 2 above, leads to another interesting result. 
Since neither set of axes consists of coplanar lines, their orientation 
determinant (the determinant formed from their nine direction 
cosines, see Exercise 12, Section 56, page 106) does not vanish. 
Hence the equations of the first set can be solved for xi, T/I, and Zi 
by Cramer's rule. If we denote the value of the orientation de- 
terminant by D, we find, for example, 

1 . . 2 " 1 1 1 M2 Ms 

D 



A, 



But this value of x\ must be identical with the value furnished by 
the first equation of the second set, for every value of x, y y and z. 
Consequently, the coefficients of x, y, and z in the two expressions 
for x\ must be equal, each to each. Hence we have 

A 2 X 3 

M2 M3 

If these equations are squared and added, we find: 

A2 AS 

M2 Ms 

But the right-hand side of this equation represents the square of 
the sine of the angle between the lines whose direction cosines are 
X2, M2, vt and X 3 , jt 3 , v 9f that is, between OFi and OZi] it is therefore 
equal to 1 (see Theorem 14, Chapter III, Section 36, page 64). 
We have therefore obtained the following result. 

THEOREM 7. The value of the orientation determinant X 2 w ^ 

Xs M3 V* 

of three mutually perpendicular directed lines is equal to -f 1 or to -1 



LINEAR TRANSFORMATION 



125 



Remark. It follows from Theorem 7 that if two of the three 
directed lines are interchanged, the value of the orientation de- 
terminant changes sign; this will also happen if the direction on 
one of the lines, or on all three, is changed. This suggests that 
whether D is +1 or 1 depends upon whether or not the three 
lines whose direction cosines are given by the elements in its rows, 
taken in the order of these rows, form a reference frame of the 
same type as the frame with respect to which their direction 
cosines are taken (see footnote on page 50). This is indeed the 
case, but a satisfactory proof of this fact can not be given without 
a more extended discussion than can find a place in this book; 
for it involves considerations of continuity. We shall therefore 
not pursue this question. 

We shall likewise omit a discussion of the conditions which the 
parameters a# and % in the equations of Theorem 4 must satisfy 
in order that these equations may represent a transformation from 
one reference frame to another with the same origin. 

66. Linear Transformation. Plane Sections of a Surface. If 
we combine the results of Theorems 2 and 5 (see Sections 61, 
page 115, and 63, page 121), we obtain formulas for the trans- 
formation of coordinates which occurs when we pass from one 
rectangular Cartesian refer- 
ence frame to another, keep- 
ing the units unchanged. For 
such a change can always be 
accomplished by a translation 
and a rotation. Suppose that, 
with reference to the frame 
0-XYZ) the coordinates of 
the new origin 0\ are a, ft, c; 
and that the direction cosines 
of the axes OiX ly OiY ly and 
OiZi are Xi, /*i, *>i, \2, M2, ^ and 

A 3 , /i3, v* respectively. Starting with 0-XYZ, we translate the axes 
to the new origin 0\\ this leads to the reference frame Oi-XtYJli 
(see Fig. 18). From this we make the transition to the frame 
Oi-XiYiZi by a rotation of axes. It follows, from Theorem 2, 
that x = x 2 + a, y = ?/ 2 + b, z = z 2 + c; and from Theorem 5, 




FIG. 18 



126 OTHER COORDINATE SYSTEMS 



that X2 = Xi^i + X 2 ?/i + Xs2i, 2/2 = MI^I + M22/i + Ma^i, 22 = 
+ v&i- We obtain therefore the following theorem. 



THEOREM 8. If *, y, * and *i, yi, * are the coordinates of an arbitrary 
point P with reference to the rectangular Cartesian frames O-XYZ 
and Oi-XiYiZi respectively, and when, with respect to O-XVZ, the 
coordinates of Oi are o, 6, c, and the direction cosines of OXi, O Yi, 

and OZi are X if MI> "u X 2> /*2, "2, and \ 3 , jua, ^ respectively, then 

Jt = Ai*i -h X2ji + Xa^i 4- > 
y = MI*I + M2ji -h M3*i -h ^ 
* = ^1^1 -f ^yi 4- fa*i -f- c. 

Remark 1. If we solve these equations for x\, yi, and zi, and 
make use of the relations established in the proof of Theorem 6 
(see Section 65, page 123), we find that 

xi = \i(x - a) + Mi(y - 6) + vi(z - c), 
2/i = X 2 (x - a) + /i2(j/ 6) + "2(2 - c), 
zi = X 3 (x - a) + Ms(2/ - 6) + "sO - c). 

It will be worth while for the reader to deduce this result by 
direct application of Theorems 2 and 5. 

Remark 2. A transformation of the frame of reference such as 
we have discussed in the preceding paragraphs will be called a 
rigid transformation. The algebraic transformation of coordi- 
nates which corresponds to it is called a transformation of the 
first degree, or a linear transformation. 

COROLLARY 1. The degree of a polynomial in *, y, *, such as/to, y, s) 9 
Is the same as that of the polynomial /i(#i, yi, s\) obtained from 
/(*, y, s) by a linear transformation. 

Proof. Since the expressions to be substituted for x, y, and z 
are of the first degree in xi, y\, and zi, it should be clear that the 
degree of fi can not exceed that of/. But since/ can be obtained 
from /i by substituting for xi, T/I, and z l the linear functions of 
x, //, and z stated in Remark 1, the degree of / can not exceed that 
of /i. Therefore the degrees of the two polynomials arc equal. 

Remark. The transformation of coordinates which corresponds 
to a rotation of axes carries a homogeneous polynomial in x, y, z 
over into a homogeneous polynomial of the same degree in x\, y\, z\. 

We are now prepared to take up a question of interest and im- 
portance, namely, to determine the character of the curve of 



LINEAR TRANSFORMATION 127 

intersection of a surface with an arbitrary plane. If one of the 
variables, let us say by way of example y, is eliminated between the 
equation of the surface f(x, y, z) = and that of the plane ax + by 
+ cz + d = 0, we obtain an equation, say F(x, z) = 0, whose 
space locus is the cylindrical surface parallel to the F-axis, which 
projects the curve of intersection of surface and plane upon the 
Z-ST-plane; and whose plane locus is the projection of this curve 
upon the ZX-plane (compare Theorems 1 and 2, Chapter IV, 
Section 40, pages 69, 71). 

When the intersecting plane is parallel to one of the coordinate 
planes the curve of intersection is congruent to its projection 
upon that coordinate plane; in that case our question can be 
answered immediately by the methods of Plane Analytical Geom- 
etry. But when the intersecting plane is in a general position, 
these two curves will not be congruent. The question can then 
be answered, as is suggested clearly by the answer in the special 
case, by first making a transformation of coordinates to a new 
reference frame, of which one of the coordinate planes is parallel 
to the given plane. How is such a transformation determined? 

Let us propose so to transform a given frame of reference 0-XYZ 
to a new frame Oi-XiYiZi that a plane whose equation in normal 
form (see Section 44) is \x + ^y + vz p = shall be parallel 
to the XiFi-plane. The necessary and sufficient condition for 
this is that the direction cosines of the Zi-axis shall be X, /u, v (see 
Theorem 7, Chapter IV, Section 44, page 78). It follows there- 
fore from Theorem 8 that the desired transformation will be ac- 
complished if we put 

2 2/i + Xzi + a, y = MI^I + ^y\ + vz\ + 6, 

+ Vtfji + VZ l + C, 



where Xi, /i j; vi, X 2 , M2, v 2 , and a, 6, c are arbitrary, save for the re- 
strictions imposed by Theorem 6. This arbitrariness in the choice 
of some of the coefficients in the equations of transformation 
corresponds to the fact that the position of the origin and that of 
the axes OiJfi, OiFi have not yet been specified. When these 
specifications have been made the equations of transformation can 
be completely determined (see the Examples below). 

In view of Corollary 1 (see page 126) the equation of the given 
surface in the new reference frame will have the same degree as 



128 OTHER COORDINATE SYSTEMS 

the original equation of the surface. Since the equation of the 
plane section of the surface is obtained from the new equation of 
the surface by replacing one of the variables by a constant, the 
degree of the equation of the curve of intersection will not exceed 
the degree of the equation of the surface. For convenience of 
reference, we record this fact as follows. 

COROLLARY 2. The degree of the plane equation of the section of a 
surface made by an arbitrary plane does not exceed the degree of the 
equation of the surface. 

Examples. 

1. To determine the curve of intersection of the sphere x' 2 -f- y~ -f 2 2 = 9 
with the plane 3 x 4 y + 12 z 2 = 0, we reduce the equation of the plane 

, ,, , , 3 x 4 ?/ 12 z 2 .. IT . 3 

to the normal form -7- ~ -f- - t - -- TTT = 0. Hence we have X = , 

lo 1J io lo lo 

M = ~ TO ' v ~ 7v> From the conditions of Theorem 6 (see Section 65, page 

lo lo 

123), it follows that Xi, MI, "i and X 2 , /* 2 , y'2 must be so chosen that 

(1) 3 Xi - 4 MI -f 12 vi = and Xi 2 -f- Mi 2 + "i 2 = 1 
and that 

(2) 3 X 2 - 4 M2 -h 12 v 2 = 0, X t X 2 -f MiM2 -f m^ = 0, 

and X 2 2 -h M2 2 -f ^2 2 = 1. 

If one of the coefficients Xi, /u t , or ^i is chosen arbitrarily and the others are 
determined so as to satisfy the equations (1), then the remaining coefficients 
X 2 , M2, ^2 are completely fixed by equations (2^ provided that Xi, MI, and ^ 
are so selected that the rank of the coefficient matrix of the first two of equa- 
tions (2) is 2. If we take v\ = 0, we find Xi : /*i =4:3 and hence \i = , 
^ = jf ; Vl =o. For the determination of X 2 , ^2, ^2 we have then the conditions 
3 X 2 4 /U2 + 12 >2 = and 4 X 2 -f 3 M2 = 0; we find therefore, by using 
Theorem 4, Chapter II (see Section 25, page 41), that X 2 : ^ v 2 = -36 : 48 : 25 
and hence that X 2 = -}j, ^ - ||, ^ 2 = jf. If, finally, we take a = 6 = c 
= 0, we obtain the following equations of transformation: 

_4xi 36 yi 3zi _3xj , 48 yi _4zi __ 5j/t , 12 2j 
X " "F " "65" +13' IJ ~ 5 "*" 65 13' 13 "*~ 13 ' 

It follows now from the discussion at the beginning of Section 65 (see page 
122) that these equations of transformation must carry the equation of the 
sphere over into the equation rci 2 -f i/i 2 -f Zi 2 = 9. The equation of the 
given plane 3x 42/ + 12z 2=0 becomes: 



and this reduces to the simple equation 13 Zi = 2. 



LINEAR TRANSFORMATION 129 

Hence the curve of intersection of sphere and plane is congruent to the 
curve in the A^Fi-plane whose equation is obtained by eliminating Zi from the 
equations Xi~ -h y\ 2 -f- ^i 2 = 9 and 13 z\ = 2, that is, congruent to the plane 
locus of the equation 169 xi 2 -f 169 yi 2 = 1517. The curve is therefore a 
circle. That the result would be a circle was evident from the fact that every 
plane cuts a sphere in a circle; the derivation which we have given serves to 
illustrate the method which can be used also in cases in which the result 
could not be so easily predicted. 

2. The curve of intersection of the plane .3 x 4 y -f- 12 z 2 = 0, used 
in Example 1, with the surface x 2 -h y 2 4- 2 z' 2 = is obtained by using the 
same equations of transformation that were used above. The equation of the 
surface now becomes 

+ *)- 9. 

If this equation is solved simultaneously with the equation of the plane, 
13 z\ = 2, we find that the curve of intersection of the surface and the plane 
is congruent with the plane locus of the equation 

(re 7 , 24 \ 2 
13 + I69 



This equation reduces to 169 a?! 2 + 194 ?/i 2 -f ~JiL l _ 1513 100 = Q; the 



240 yi 
Lu;j xi- -p is*** /yr i 

curve is therefore an ellipse. 

67. Exercises. 

1. Set up the equations for a transformation of coordinates which carries 
the plane 2 x y -\- 2 z 5 = over into the ZjAVplane. 

2. Determine the equations of transformation when the plane x + 2 y 
- 2 z -f- 4 = is to become the Fi#i-plane; the line x + 2y 2z + 4 =0, 
3 x ?/-}- z 3 = becomes the Zi-axis; and the point (?, \, 2) is the 
new origin. 

3. Set up the equations for a transformation of coordinates which will 
carry the line 3 x + 4 y 2 z = 0, = into the AVaxis. 

4. Determine the volume of the coordinate parallelepiped of the point P 
whose coordinates are x 4, y 2, z 1 in the reference frame in 
which the planes 2x 2y -f z = 0, y -\- z = and x z = are the YiZi-, 
ZiXi-, and X^Yi -planes respectively. 

6. Determine a plane equation of the curve of intersection of the plane 
6 x 2y -{- 3 z 4 =0 with each of the following surfaces: 

(a) 3 x 2 -f 2 y 2 - 3 z 2 + 18 = 0; (6) 4 x 2 - y 2 = 6 z; 
(c) 3xy -f 4yz -2 zx 10. 

6. Determine, for each of the axes of the new frame of reference introduced 
in Exercise 3, Section 64 (page 121), such a direction that the new frame is of 
the same type as the original frame; also such directions that the new frame 
is of the opposite type. 



130 OTHER COORDINATE SYSTEMS 

7. Solve the corresponding problem for the reference frame introduced in 
Exercise 4, Section 64. 

8. Set up the equations of transformation for the transition from a refer- 
ence frame O-XYZ to a new frame whose origin has in 0-XYZ the coordinates 
5, 3, 2 and whose X-, F-, and Z-axes have in O-XYZ direction cosines 
which are proportional to 4 ; 8, 1; 3, 2, 4, and to 34, 13 and 32 respectively. 
Decide how to select the directions on the new axes to make the new frame 
of the same type as O-XYZ. 

9. Interpret geometrically the transformation of coordinates which is 
determined by the equations 3 x = 2 oj ?/i 2 z\, 3 y = x\ 2 y\ -{- 2 z\ 
-6, 3 z = 2 xi + 2 ?/i 4- zi -f 3. 

10. Prove that the curve in which an arbitrary plane meets the locus of an 
equation of the second degree in x, y, and z is a conic section or a straight line. 



CHAPTER VI 



GENERAL PROPERTIES OF SURFACES AND CURVES 

Before undertaking a somewhat detailed study of the loci of 
equations of the second degree in x, y, and z, we shall consider in 
the present chapter a few general properties of surfaces and curves. 
Our attention will be restricted almost entirely to the loci of equa- 
tions of the form /(z, y, z) = 0, whose left-hand side is a poly- 
nomial in the three variables. The case in which not all three of 
the variables are actually present in the equation has been con- 
sidered in an earlier chapter (Section 40). We recall also Defi- 
nitions I and II of Chapter IV (Section 39). 

68. Surfaces of Revolution. 

DEFINITION I. A surface of revolution is a surface that can be gen- 
erated by revolving a plane curve about a line in its plane. The line a 
around which the revolution takes place is called the axis of revolution 
of the surface; the plane curve, in any 
of Its positions, is called a meridian 

curve. 

It should be clear that all meridian 
curves are congruent plane curves and 
that the planes in which they lie con- 
stitute a pencil of planes through the 
axis of revolution.* 

Every point P on the given curve, 
see Fig. 19, describes a circle, whose 
center is the projection P a of P on 
the axis of revolution and whose 
radius is P a P; these circles, which 

lie in planes perpendicular to the axis of revolution, are called 
parallel circles. 

* The reader will notice that the simplest curved surfaces studied in ele- 
mentary Solid Geometry, spheres, cones, cylinders, as well as the surfaces of 
a large number of manufactured articles in daily use, such as teacups, lamp- 
shades, hats, are, exactly or approximately, surfaces of revolution. Is there 
a possible simple reason for this? 

131 




Axis 



Meridian 



FIG. 19 



132 GENERAL PROPERTIES OF SURFACES AND CURVES 



Our principal problem is to determine the equation of a sur- 
face of revolution when we are given the equations of a meridian 
curve and those of the axis of revolution. In case the axis of 
revolution does not coincide with one of the coordinate axes, we 
can always transform the frame of reference in such a manner as 
to make one of the new coordinate axes coincide with the axis of 
revolution; and there is evidently no loss in generality if we 
assume that the meridian curve from which we start lies in one 

of the coordinate planes through 

i the axis of revolution. We shall 

therefore suppose that the given 
curve lies in the XZ-plane and 
wo shall seek the equation of the 
surface obtained by revolving 
this curve about the X-axis (Fig. 
20). Let the equations of the 
meridian curve be f(x, z) 0, 
y = 0; and let P(x, y, z) be an 
F IO 20 arbitrary point on the surface. 

The parallel circle through P 
will cut the initial meridian in a point Q(x', 0, z'). It should now 




be easy to see that x' = x, and z r = P X P = 



Since 



Q is a point on the meridian, its coordinates satisfy the equation 
f(x, z) = 0. Consequently, the coordinates of P satisfy the 
equation f(x, Vy 2 + z 2 ) = 0. Conversely, if the coordinates of 
a point P f satisfy this equation, then the coordinates of the point 
Q' in which the circle through P' with center at P x ' and radius 
P x 'P f cuts the XZ-plane, will satisfy the equations f(x, z) = 
and y = 0; the point Q' will therefore lie on the given curve and 
the point P' on the surface of revolution which this curve generates 
when revolving about the X-axis. We have therefore reached the 
following conclusion: 

THEOREM 1. The equation of the surface of revolution generated 
when a plane curve in the ZX-plane revolves about the X-aiis, is ob- 
tained by replacing z in the plane equation of this curve by Vy + * 2 
and then rationalizing the equation. 

It should be a simple matter to state similar conclusions for 
the surface of revolution that is obtained when a curve in any co- 



SURFACES OF REVOLUTION 



133 



ordinate plane revolves about either axis in that plane (see Section 
69). 
Examples. 

1. The equation of the surface generated when a parabola in the X F-plane, 
whose equations are y z = 4 ax and z = 0, is revolved about the X-axis, is 
obtained by replacing y in the plane equation of the curve by V?/ 2 -f- z 2 and 
then rationalizing. The surface is 

called a paraboloid of revolution; its 
equation is y 2 -f- z 2 = 4 ax (see Fig. 
21). 

2. Of especial interest to us are the 
surfaces of revolution generated when 
the conic sections are revolved about 
one of the axes of the curve. The 
shapes of these surfaces can easily be 

pictured. By taking the equations of FIG. 21 

the curves in the standard forms, fa- 
miliar from Plane Analytical Geometry, we obtain readily the following 
results: 




Meridian curve 
(1) x 2 -f ?/ 2 = a 2 , Circle 



Axis of Equation of sur- 
revolution fare of revolution 



X-, or F-axis 
X-nxis 



a 2 ~*~ b* + 6 2 



(5) 



Name 

Sphere. 

Ellipsoid of revo- 
lution ; prolate 

spheroid. 
Ellipsoid of revo- 
lution; oblate 

spheroid. 

Hyperboloid of 

revolution of two 

sheets. 

The surface that is obtained in this ease consists of two parts entirely 
separate from each other; they are called the two sheets (nappes) of the sur- 
face compare also the Remark on page 136. 

_, ... -2 ..2 ,2 Hyperboloid of 

revolution of one 

sheet. 

Paraboloid of 
revolution. 



a > b 

(.3) ~2-f^ 2 = *. Ellipse 
a > b 

(4) ^ ~ iT2 = L Hyperbola 



F-axis 



JT-axis 



~2 - ^2 = 1, Hyperbola 



(6) |/ 2 = 4 oar, Parabola 



(see Example 1) 

(7) 2/ 2 ra 2 * 2 = 0, Pair of 

intersecting lines. 

(8) i/ 2 m 2 = 0, Pair of 

parallel lines. 



F-axis 



-ST-axis 



.XT-axis 



2/ 2 -f- z 
m 2 x 2 y 2 



4 ax 



Circular cone. 



Circular cylinder. 



3. Suppose that we wish to determine the equation of the surface obtained 
when the line 2x + 6y 3^ + 1 =0, 3x y + 2z 3 =0 revolves about the 



134 GENERAL PROPERTIES OF SURFACES AND CURVES 

Iine2x + 6y -3*-f-l =0, x + y +z = 0. Evidently the two lines lie in the 
plane 2x + 6?/ 32-f-l =0. Therefore we determine a transformation of co- 
ordinates to a new frame of reference Oi-XiYiZi, in which this plane becomes the 
XiFi-plane and the line in which it meets the plane x+y+ *> =0, that is, the axis 
of revolution, becomes the Xi-axis. As origin Oi of the new frame, we shall take 
the point in which the two given lines meet. In the notation of Section 66, we 
have therefore a = 1, b - f , c = J. Since the XV-axis is the intersection of 
the planes 2 z 4- 6 i/ 3 2 -f 1 =0 and x + y -f z - 0, we find, by use of 
Theorem 17, Chapter IV (see Section 47, page 87), \i : MI : "i = 9 : -5 : -4; 

954 

and therefore \i = . _ , //i = -- ==, vi = -- ==. The Zi-axis is per- 
V122 V122 V122 

pendicular to the plane 2 x -f 6 y - 3 z -f 1 =0; its direction cosines are there- 
fore proportional to 2, 6, -3 (see Theorem 7, Chapter IV, Section 44, page 
78). Therefore X 3 = ?, MS = ?, "a = f. The direction cosines X 2 , M2, *2 
of the Fi-axis must therefore satisfy the conditions 9 \2 5 M2 4 1/2 = 
and 2 X 2 -h 6 M2 - 3 1/2 = (see Corollary 2 of Theorem 13, Chapter III, 
Section 36, page 64). From this we find, by use of Theorem 4, Chapter 

39 
II (Section 25, page 41) that X 2 : M2 : "2 39 : 19 : 64, and that X 2 = ~~F= . 

19 64 

=- Hence the equations of transformation are, in 



- - , 2 - 

7\/122 7V122 

view of Theorem 8, Chapter V (Section 66, page 126) 

fl*i , 39 yi . 2 ^ nr 9(x - 1) - 5(y + }) - 4(g + t) 

-- - -=- H- = - - 



x -- 7= - -=- - , i 

Vl22 7V122 7 Vl22 

5si 19 yj 6zi 



"r" l " 7V122 

4 on , 64 y 



The first of 'these two sets of formulas will carry the equations 

2z+62/-3z + l' = over into 7zi = Q 
x + y + z = Vl22 1/1 + 5 z, = 

3z-2/-f 2z -3=0 84xi-f 113i/i -3VI222! =0. 

Hence the equations of the given line may be written in the form zi = 0, 
84 x\ -f- 113 2/1 = 0; and the equations of the axis of revolution can be put 
in the form Zi - 0, 2/1 = (compare Remark 2, following Theorem 18, Chap- 
ter IV, Section 49, page 92). The equation in Oi-XiYiZi of the required 
surface of revolution is therefore obtained by rationalizing the equation 
84 xi -f 113VV -j- Zl * = 0. This gives the equation -84 2 Zi 2 -f 113 2 (2/i 2 -h*i 2 ) 
= 0. To obtain the equation of the surface in 0-XYZ (from elementary 
Solid Geometry we know that it is a circular cone), we now substitute for 
xi t 2/1, and z\ the expressions in terms of x, y, z to which they are equal by the 
second set of transformation equations. This gives us for the equation of 
the required cone 

-49-84 2 (9:c-52/-43- Y) 2 + H3 2 (39z + 190 + 64z - 5) 2 
+ 122 113* (2 x + 6 y - 3 z + I) 2 0. 



THE SHAPE OF A SURFACE 135 

69. Exercises. 

1. Prove that, if z = 0, g(x, y) = are the equations of a curve in the .XT- 
plane, then: (a) g(x, V?/ 2 -f z 2 ) =0 is the equation of the surface of revolu- 
tion obtained by revolving this curve about the A r -axis; (b) g( Vx 2 4- z 2 , y) = 
is the equation of the surface of revolution obtained by revolving the curve 
about the 7-axis. 

2. Determine the equation of the surface which is generated when the circle 
x = 0, t/ 2 + (z - 5) 2 = 9 in the FZ-plane, is revolved, (a) about the Z-axis; 
(b) about the F-axis. 

NOTE. The surface generated when a circle revolves about a line in its 
plane not through its center, is called an anchor ring, or a torus. Which of 
the surfaces described in this problem is a torus? 

3. Determine the equation of the surface obtained by revolving the lem- 
niscate (x 2 -f 2/ 2 ) 2 = 8 (x 2 ?/ 2 ), (a) about the X-axis; (b) about the K-axis. 

4. Determine the equation of the circular cylinder which is formed when 
the line z = 0, 3 x 4 ?/ -}- 2 = revolves about the line z 0, 3 x 4 y 
-8 = 0. 

6. Develop the general equation of a torus. (Take the axis of revolution 
as one of the coordinate axes.) 

6 Determine the equation of the surface generated when the equilateral 
hyperbola xy = a 2 is revolved about one of its asymptotes. 

7. Determine the equation of the surface of revolution obtained by re- 
volving (a) the parabola z 2 4 ay about the Z-axis; (6) the semi-cubical 
parabola x 2 = z 9 about the Z-axis; (c) the same curve about the X-axis; 
(d) the curve y = sin x about the X-axis. 

8. Determine the equation of the oblate spheroid obtained by revolving 

the ellipse ( * ~ 3) * + (y ~ 2) " = 1, z = about the line y = 2, z = 0. 
rr J 

9. Determine the equation of the circular cone which is generated when the 
line 8 z 4y -{- z 2 = 0, 2 x y 22 + 3=0 is revolved about the 
line 8z-4 y + z -2 = 0, z + 2i/+22-4 = 0. 

10. The hyperbola (x 4) 2 7 =1, x = is revolved about its 

asymptote with positive slope. Determine the equation of the surface of rev- 
olution which is generated. 

70. The Shape of a Surface Determined from its Equation. 
Contour Lines. One of the fundamental problems of Solid 
Analytical Geometry is that of forming a clear picture of the shape 
of the surface which is the locus of a given equation. We have 
already obtained a number of partial solutions of this problem; 
we will begin by summarizing these: 

(a) the locus of an equation of the first degree is a plane. 
(6) the locus of an equation from which one variable is absent 
is a cylindrical surface. 



136 GENERAL PROPERTIES OF SURFACES AND CURVES 

(c) the locus of an equation which can be written in one of the 
forms /(*, vV + z 2 ) = 0, f(y, Vz* + x 2 ) = or f(z, Vx* + y 2 ) 
= is a surface of revolution; the meridian curve of such a sur- 
face can then be found by the methods of Plane Analytical Geom- 
etry. 

To these we now add the following further results. 

THEOREM 2. The locus of an equation which can be reduced to the 
form (x - a) 2 -f (y - 6) 2 4- (s - c) 2 = r 2 Is a sphere whose center Is at 
(a, f>, c) and whose radius Is r. 

The proof of this theorem is left to the reader. 

DEFINITION II. A conical surface Is a surface generated by a line, 
extending Indefinitely, which moves In such a way as to pass always 
through a fixed point, called the vertex, and successively through the 
points of a fixed curve, called the directrix. 

Remark. It is a consequence of this definition that a conical 
surface consists of two parts, one on each side of the vertex and 
connected at the vertex; these parts are called the "sheets" or 
"nappes" of the surface (compare Section 68, Example 2, part 4, 
page 133). If the directrix is a plane curve and the vertex lies 
in the plane of this curve, the conical surface is a sector of this 
plane; if the directrix is a straight line, we obtain the entire plane. 
In case the directrix consists of a pair of lines which do not both 
lie in a plane with the vertex, the surface consists of a pair of 
intersecting planes; if both lines lie in a plane with the vertex, 
the surface reduces to a single plane counted twice. According 
as the directrix is a curve of the first, second, third order, etc., 
the conical surface is said to be of the first, second, third order, 
etc.; compare also Remark 1, Section 82, page 166. 

THEOREM 3. The locus of an equation In x 9 y, and Is a conical sur- 
face with vertex at the origin If and only If the equation Is homogeneous. 

Proof. Recalling the definition of a homogeneous equation 
(Section 20, page 35) we observe that if the equation f(x, y, z) = 
is a homogeneous equation of degree n, then and only then will 
all the terms in the polynomial f(x, y, z) be of the nth degree in the 
three variables jointly, and hence we know that f(kx, ky, kz) = 
k n f(x, y, z). Suppose now that P(a, , 7) is a point on the locus of 
the equation /(x, y, z) = 0, supposed to be homogeneous of degree n. 



THE SHAPE OF A SURFACE 137 

The coordinates of an arbitrary point on the line OP are given by 
the equations x = ta y y = tf$, z = ty, in which t is a parameter that 
varies from point to point along the line (see Corollary 3 of Theo- 
rem 6, Chapter III, Section 33, page 56). And it follows from the 
homogeneity of the equation that f(ta, t$, ty) = t n f(a, 0, 7) = 0. 
Consequently the entire line OP lies on the surface represented by 
the equation f(x 9 y, z) = 0, if P does. The line OP will therefore 
generate the surface if P passes successively through the points of 
a curve in which the surface is cut by a plane not through the 
origin. The proof of the converse of this theorem is left to the 
reader. 

Remark 1 . The locus of the linear homogeneous equation ax + by 
+ cz = is a plane through the origin. In view of the remark 
preceding Theorem 3 this is a special case of a conical surface 

Remark 2. If an equation is homogeneous in x o, y 6, 
z c, translation of the axes to the point A (a, b, c) as a new origin 
will reduce it to an equation which is homogeneous in x' x a, 
y' = y b, z' = z c (see Theorem 2, Chapter V, Section 61, 
page 115). Its locus is therefore a conical surface whose vertex 
is at A (a, b, c). Conversely, translation of axes to A as origin 
puts the vertex of a conical surface with vertex at A at the origin 
of the new reference frame. We can therefore state the following 
corollary: 

COROLLARY. The locus of an equation in x, y, a Is a conical surface 
with vertex at the point (a, 6, c) if and only if it is homogeneous in 
x a, y b, and z c. 

THEOREM 4. The locus of the equation /(*, y, *) = is symmetric 
with respect to the YZ-plane* if and only if the equations /(-*, y, *) 
= and /(*, y, z) = are equivalent;! the locus of this equation is 
symmetric with respect to the Z-axis* if and only if the equations 

* Two points A and B are said to be symmetric with respect to a plane (or 
line) if the segment AB is bisected perpendicularly by the plane (or line); 
they are called symmetric with respect to a point if the segment AB is bi- 
sected by that point. A surface is said to be symmetric with respect to plane 
(line, point) if with every surface point A there is associated a surface point B 
such that A and B are symmetric with respect to the plane (line, point). 

t Two equations are called equivalent if any set of values of the variables 
which satisfies either of the equations, also satisfies the other equation; 
e.g., the equations x 2 - 2 y 3 + 4 z* + y = and (-x) 2 - 2 y 3 + 4 (-z) 2 
4- y = are equivalent, also the equations x 3y + 2 z 3 = and 2 x -f 6 y 
- 4 a 3 = 0. 



138 GENERAL PROPERTIES OF SURFACES AND CURVES 

/(-*, -y, *) = and/ r, y, z) = arc equivalent; this locus is sym- 
metric with respect to the origin if and only if the equations f(x 9 
-y> -*) = and/(*, y, s) = are equivalent. 

This theorem and its obvious counterparts which assert sym- 
metry with respect to the other coordinate planes and coordinate 
axes are immediate extensions of well-known theorems of Plane 
Analytical Geometry. The reader is entrusted with proving them. 
We see, for example, that the locus of the equation z 2 2 y* 
+ 4 z 2 + y = is symmetric with respect to the F-axis, and that 
the locus of the equation x 3 ?/ + 2 z 3 = is symmetric with 
respect to the origin. Compare Exercise 3, Section 31, page 52. 

Of great value in determining the shape of a surface represented 
by an equation are the contour lines of the surfaces; these are the 
projections upon a fixed plane of the intersections of the surface 
with a series of planes parallel to this fixed plane. We shall make 
use only of sets of planes which are parallel to the coordinate 
planes; accordingly we lay down the following definition: 

DEFINITION 111. The X-contour lines of a surface are the projections 
on the YZ-plane of the curves in which the surface is met by planes 
parallel to the YZ-plane, that Is, by the planes whose equations are 
x = a; the Y-contour lines (Z-contour lines) are the projections on 
the ZX-plane (the A 1 -plane) of the curves in which the surface is met 
by the planes parallel to the ZX-plane (the XI -plane), that is, by the 
planes y = constant (the planes z = constant). 

It follows immediately from this definition, in conjunction with 
Theorem 2, Chapter IV (Section 40, page 71), that the .XT-con- 
tour lines of the surface f(x, y, z) - are the curves in the FZ- 
plane whose plane equations are /(fc, y, z) = 0, in which k is a 
parameter taking all real values. Similarly the F-contour lines 
and the Z-contour lines are given by the equations f(x, k, z) = 
and f(x, y, k) = 0, respectively. 

If we plot a set of contour lines, affixing to each of them the 
value of the parameter k to which it corresponds, we obtain a 
diagram which we shall call a contour map of the surface; such 
contour maps should be of considerable aid in forming a mental 
picture of the locus of an equation. The contour maps published 
by the U. S. Geological Survey are approximately Z-contour 
maps of portions of the earth's surface, if the level plane at some 
point in the region is taken as the XT-plane. 



THE SHAPE OF A SURFACE 139 

Remark. In engineering practice various other methods are 
used for representing space configurations in a plane drawing. 
Of these we mention Descriptive Geometry and Perspective Draw- 
ing. These subjects are treated in books especially devoted to 
them. We shall not attempt to discuss them here; the interested 
reader will have little difficulty in getting access to such books. 

Example. To determine the J-contour lines of the locus of the equation 

x 2 ?/ 2 2 2 

-r - -f- ^ = 1, we consider the set of curves in the F-plane whose plane 

4 y ID 

2 2/2 ^2 

equations are -r + TT; 1. There are three cases to be distinguished: 
4 y 10 

k 2 
(a) \k\ < 2, that is, -2 < k < 2. In this case k 2 < 4, so that 1 - y > 0. 

The equations of the X-contour lines for these values of the parameter k can 
therefore be put in the form 



= 1. 



From this we conclude that these contour lines are hyperbolas whose center 
is at the origin; that the transverse axis is along the Z-axis and the conjugate 
axis along the F-axis. The semi-axes of the hyperbola are a = 2\/4 k 2 

and b = ^ ~~ ; hence y- = 41 and all the hyperbolas of the set have the 
2 06 

4 V .5 

same asymptotes, namely, the lines z = ^ , and the same eccentricity, '-. 

(5\/4 k 2 \ 
0, 0, =b - ^ - y 

(b) \k\ =2, that is, k = 2. The equation of the contour lines now 

reduces to - -f ~ =0; the locus of this equation consists of a pair of inter- 
y JLO 

secting lines, namely the asymptotes common to the hyperbolas discussed 
under (a). 

(c) \k\ > 2, that is, k > 2, or k < -2. In this case k 2 > 4, so that 

k 2 y 2 z 2 

I - < and the equation takes the form / , - r -- -^- - r- = 1. 

4 KT-O 16 (T-0 

The contour lines are now hyperbolas with center at 0, the transverse axis 
along the Y-axis and conjugate axis along the -axis. The semi-axes are 



a = g ~ 4 and 6 =2 Vfc^=T; the foci are at the points (o, =fc 5 ^ " 4 , 
on the F-axis; the asymptotes are the same as those of the hyperbolas in (a), 
namely, the lines z = -~ , and the eccentricity is again equal to -? 

o 4 




140 GENERAL PROPERTIES OF SURFACES AND CURVES 

The X-contour map of this surface is sketched in Fig. 22, the plane of the 
drawing being the ZF-plane. It follows from Theorem 4 that the surface is 
symmetric with respect to the three coordinate planes, the three coordinate 
axes and the origin. The part of the surface suggested by this contour map, 

which lies to the right of the 
ZF-plane is therefore duplicated 
by a symmetric part to the left 
of this plane. 

It should be easy for the 
reader to show that the Z-con- 
tour lines of this surface are 
also hyperbolas and that the 
Z-contour map has the same 
general character as in Fig. 22. 
The F-contour lines are ellipses; 
the reader should construct the 
F-contour map. The surface 
represented by this equation 
is called a hyperboloid of one 
Fi(i. 22 sheet. Of the axes of sym- 

metry, the X- and Z-axes meet 

the surface in real points; they arc called the transverse axes of the sur- 
face. The F-axis does not meet the surface in real points; it is called the 
conjugate axis of the surface. 

71. Some Facts from Plane Analytical Geometry. In prepara- 
tion for the construction of the contour maps of other surfaces, 
we summarize in this section some facts from Plane Analytical 
Geometry with which the reader should be familiar. 

(1) the locus of the equation ~ + ~ = 1, a > fc, is an ellipse, 

whose major and minor axes are along the X- and F-axis respec- 
tively. The center is at the origin; the vertices at the points 
(a, 0) and the foci at the points (dbc, 0), where c 2 = a 2 fc 2 . 

The eccentricity e is equal to - , which is < 1 ; the directrices are 

the lines x - . The ratio of the distances of any point on the 

ellipse from a focus and from the corresponding directrix is equal 
to e; the sum of the distances of any point on the ellipse from the 
two foci equals 2 a. 



The locus of the equation T^ + ~=l,a 



6, is obtained from 



SOME SPECIAL SURFACES 141 

the ellipse just discussed by interchanging the roles of the X- 
and F-axes. 

x 2 y* 

(2) The locus of the equation -^ ^ = 1 is an hyperbola with 

center at the origin, transverse axis along the X-axis and conju- 
gate axis along the F-axis. The vertices are the points (ita, 0), 
the foci at (dbc, 0), where c 2 = a 2 + 6 2 . The eccentricity e is 

(* CL 

equal to - , which is > 1 ; the directrices are the lines x = - 
H a' e 

/j |y 

The asymptotes are the lines - ~ = 0. The ratio of the dis- 
tances of any point on the hyperbola from a focus and from the 
corresponding directrix is equal to e\ the difference between the 
distances of any point on the hyperbola from the two foci is equal 
to db 2 a. 

/Jr2 n 2 

The properties of the locus of the equation -5 p = 1 are 

analagous to those of the curve just discussed. The vertices and 
foci lie on the F-axis, which is therefore the transverse axis; 
the X-axis is the conjugate axis. The asymptotes, the number c 
and the eccentricity are the same as the corresponding elements of 
the first hyperbola. The two curves are called conjugate hyper- 
bolas. 

(3) The locus of the equation y 2 = 4 p(x a) is a parabola 
whose axis is on the X-axis. The vertex is at the point (a, 0) ; 
the focus at the point (a + p, 0). The curve will therefore open 
in the direction of the positive or the negative X-axis, according 
as p is positive or negative. The directrix is the line x = a p. 
The ratio of the distances of any point on the parabola from the 
focus and the directrix is equal to 1. 

72. Some Special Surfaces. We should now be able to con- 
struct readily the contour maps of the loci of a number of special 
equations. The work will only be sketched; the reader should 
work out carefully the details of the constructions. 

X 2 IV 2 2 

(a) the equation -5 + fj + -5 = 1 ; a > b > c.* 

CL U C 

* The case in which two or more of the numbers a, 6, c are equal has been 
discussed in Example 2, Section 68 (page 133). 



142 GENERAL PROPERTIES OF SURFACES AND CURVES 



The Z-contour lines are given by the equation 



6 2 ( 1 - 



= 1. They are ellipses whose semi-axes are equal 



to 



aVc 2 - fc 2 



and 




aVc 2 - 
The foci are at the points 



The eccentricity is equal to 
and therefore independent of fc. 



Va 2 - 6 2 Vc 2 - fc 2 



, 0, o). 



For 



fc = 0, the semi-axes have their largest values, a and 6; as |fc| in- 
creases from to c, the semi-axes decrease; for fc = c, they are 
and the ellipses shrink down to the points (0, 0, dbc) as fc tends 
towards c; when |fc| > c, the ellipses are imaginary. We con- 
clude that the contour map consists of a series of similar ellipses. 
In Fig. 23 the contour lines for the part of the surface above 

the .XT-plane are sketched. 
The surface is symmetric 
with respect to the XY- 
plane; the contour map for 
the part of the surface be- 
low the -XT-plane is there- 
fore identical with the one 
here drawn. This surface 
is called an ellipsoid; its 
shape is suggested in per- 
spective by Fig. 24. The segments a, 6, and c are called the semi- 
axes of the surface; they are equal to one half of the segments 
which the surface cuts off on the axes of symmetry. 




FIG. 23 



(6) the equations 



Z 2 J/ 2 


2 2 


Z 2 


2/ 2 z 2 


2 + fe 2 


C 2 


' a 2 


6 2+ c 2 


2+^ = 


1. 







1 ; and 



An equation of the second of these types has already been dis- 
cussed in the Example of Section 70. We shall therefore leave the 
discussion of these equations to the reader. Surfaces represented 
by an equation of this form are called hyperboloids of one sheet; 



SOME SPECIAL SURFACES 



143 




FIG. 24 




FIG. 25 



144 GENERAL PROPERTIES OF SURFACES AND CURVES 



the segments a, 6, and c are its semi-axes. Figure 25 suggests the 
appearance of an hyperboloid of one sheet. 

(c) the equations^ - fL - 1 = i ; _-+L_!L = i 



The loci of these three equations are of the same general charac- 
ter. Interchanges among the coordinate axes will carry one of 
them over into the others; we shall therefore make a discussion 
of the first of these equations only, and we shall suppose that 
6 > c.* The X-contour lines are determined by the equation 

t/ 2 z 2 
___ _j = i They are ellipses, whose semi- 



the eccentricity is 



, . bVk* - a 2 . cVk* - a 2 
axes are equal to and ; 



equal to 



Vb* - 



and is 




therefore independent of k. 
For \k\ < a, the semi-axes 
are not real; therefore no 
points of the locus are 
found between the planes 
x = a and x = a. For 
k = =ta, the ellipses re- 
duce to the points (a, 
0, 0). For \k\ > a, we 
obtain a series of similar 
ellipses, which increase in- 
definitely in size as \k\ 
increases. The surface 
consists therefore of two 
sheets, each of which 
extends indefinitely, one 
in the direction of the 

positive X-axis, the other in the direction of the negative X-axis. 

The surface is called an hyperboloid of two sheets (see Fig. 26). 
* In case b c, the locus is a hyperboloid of revolution of two sheets, see 

Example 2, Section 68, page 133. 



FIG. 26 



SOME SPECIAL SURFACES 145 

Of the three axes of symmetry, the X-axis meets the surface in 
real points; it is called the transverse axis. The other axes of 
symmetry which do not meet the surface in real points are called 
the conjugate axes. 

.2 yZ L 

(d) the equation ~2 + r^ + 1 = 1 . 

(JL U C 

It should be clear that there are no real points whose coordinates 
satisfy this equation ; its locus is called an imaginary ellipsoid. 

.2 y2 g2 .2 yZ g 2 

(e) the equations-; + fr s == and -5 ffa = 0. 

\ ' ^ a 2 o 2 c 2 a 2 o 2 c 2 

It follows from Theorem 3 (Section 70, page 136) that the locus 
of these equations consists of conical surfaces of the second order. 
One contour map consists of similar ellipses, the other two of 




FIG. 27 

similar hyperbolas. In case the denominators in the two terms 
with like signs are equal, the locus is a circular cone (see Example 2, 
Section 68, page 133); otherwise it is an elliptic cone (see Fig. 27). 

(/) the equation ^ + j + 5 = 0. 

The only real point on the locus of this equation is the origin; 
the surface is called an imaginary cone. 

# 2 i/ 2 
(0) the equation -^ + ~ = 2 pz 9 a > b. 



146 GENERAL PROPERTIES OF SURFACES AND CURVES 



The Z-contour lines are the similar ellipses 



+ 



tr 



1, 




Jpfca 2 ' 2pfc6 2 
whose semi-axes are equal to a V2 pk and 6 V2 pk, and whose 

eccentricity is equal to . For fc = 0, the ellipse reduces 

to the origin. If p > 0, the 
ellipses are real f or fc > and 
imaginary for fc < 0; if p < 0, 
the situation is reversed. Con- 
sequently the surface lies en- 
tirely on one side of the -XT- 
plane and extends indefinitely 
on that side, on the side of 
the positive Z-axis if p > 0, 
on the side of the negative 
Z-axis if p < 0. The Z-con- 
tour lines and the F-contour 
lines are parabolas. The sur- 
face is called an elliptic para- 
FIG. 28 boloid (see Fig. 28) . It should 

be clear that the equations 

v 2 z 2 z 2 x 2 

~ + -5 = 2 px and -^ + -5 = 2 py also represent elliptic parabo- 

c/ c c a 

loids. 

(h) the equation -g p = 2 pz. 

The ^T-contour lines are the parabolas in the FZ-plane which are 

y 2 fc 2 
represented by the equation ~ = -5 2 pz, which can be written 

<fc 2 \ 
z ^ 2 ) The axes of these parabolas 
z pa / 

(fc 2 \ 
0, 0, ^ r ). 
^ pa / 

Therefore, if p > 0, the parabolas will all extend in the direction of 
the negative Z-axis, whereas their vertices will move upward along 
the Z-axis as fc increases. But, if p < 0, the parabolas extend in 
the direction of the positive Z-axis, and their vertices move down- 



SOME SPECIAL SURFACES 



147 



ward along the Z-axis as k increases, 
map has been drawn, for the case 
p > 0, of the part of the surface 
which lies on the right side of 
the FZ-plane; since the surface is 
obviously symmetric with respect 
,to the FZ-plane, the X-contour 
map of the other part of the sur- 
face is identical with the one 
drawn in this figure. 

The F-contour map consists of 
the parabolas 



In Fig. 29, the Jf -contour 



It should be clear that, if p > 0, the 
contour lines are upward extending 
parabolas which move downward 
as k increases; and if p < 0, down- 
ward extending parabolas which 
move upward as k increases. 

The Z-contour lines are hyperbolas. 




Fio. 29 
The surface is called an 




FIG. 30 

hyperbolic paraboloid (see Fig. 30). It should be clear that the 
equations ~ = 2 

D C 

>lic naraholoids. 



2 2 ~2 

, <> a/ 

and -5 5 = 

c 2 a 2 



148 GENERAL PROPERTIES OF SURFACES AND CURVES 

73. Exercises. 

1. Show that the locus of each of the following equations is a sphere; de- 
termine for each of them the radius and the coordinates of the center: 

(a) z 2 + y 2 + * + 4 x - 6 y + 4 z - 8 - 0; 

(6) z 2 + 2/ 2 + z 2 - 3 x -f 4 ?/ - 2 z + 3 - 0; 

(c) * 2 -f 2/2 -f z 2 4. 6 s -f 4 T/ + 4 2 + 17 = 0; 

(d) z 2 + y 2 + * 2 - 8 z - 6 ?/ - 4 z + 30 = 0; 

(e) x 1 4- ?/ 2 + z 2 4- 2 az + 2 by -f 2 cz + d = 0. 

2. Prove that the locus of the pair of equations 2 x 6?y + 3z 5 = o 
and x 2 4- y 2 -f- z 2 4 x 6 ?/ 4- 8 z 7 = is a circle. Determine the 
radius of this circle and the coordinates of its center. 

Hint: The center of the circle is the foot of the perpendicular dropped on 
the plane from the center of the sphere; the length of this perpendicular, the 
radius of the circle and the radius of the sphere form a right triangle. 

3. Show that the points common to the two spheres x 2 -f 2/ 2 4- z 2 6 x -f 

6 y - 8 z 4- 3 = and z 2 4- 2/ 2 4- z 2 - 3 z 4- 2 ?/ - 5 z 4- 4 = lie in a 
plane. 

Hint: Write the equation of the pencil of spheres through the points com- 
mon to the two given spheres and show that the pencil contains a plane. 
(Compare Remark 4, after Theorem 18, Chapter IV, Section 49, page 92.) 

4. Show that the pencil of spheres 

ki(x* + ?/ 2 4- z 2 4- 2 aix 4- 2 b t y -f- 2 c& -f di) + k 2 (x* -f ?/ 2 -f z 2 
- 2 b*y + 2 c 2 z + 4) = 



contains a plane and prove that this plane is perpendicular to the line which 
joins the centers of the two spheres, given by fci= 0, fc a = 1 and fci= 1, fc 2 = 0. 
6. Prove that the curve determined by the equations x 7/+z = Oand 
3 _|_ ^3 = 2 s j s symmetric with respect to the origin but not symmetric with 
respect to any of the coordinate axes or coordinate planes. 

6. Set up the conditions under which the plane ax + by + cz -f- d = 
meets the sphere x 2 + y z + z 2 -f 2 aix -f 2 6iy -h 2 cis + di = in real 
points. 

7. Determine, supposing that the conditions of Exercise 6 are satisfied, 
the radius and the coordinates of the center of the circle in which the plane and 
the sphere meet. 

8. Determine the equation of the conical surface of the second order whose 
vertex is at the origin and whose directrix is the circle 2x ?/ 3z-f7 = 0, 
* + y s + **-2-4y + 22-5 3S 0. 

Hint: Two methods for solving this problem should be considered: First 
method The required equation must be homogeneous of the second degree; 
moreover it must be of the general form k\S\ -f &2$2 = 0, where $1 = and 
$2=0 represent respectively the plane and the sphere, where k\ is a linear 
function of x, y, and z, and where fe is a constant. Second method If 



THE INTERSECTIONS OF A SURFACE AND A LINE 149 

P(x, y, z) is an arbitrary point on the conical surface with vertex at the origin, 
then there must exist a factor of proportionality r, such that rx, ry, rz satisfy 
the equations of the plane and the sphere. Elimination of r leads to the re- 
quired equation. 

9. Find the equation of the circular conical surface whose vertex is at 
F( 3, 2, 1) and whose directrix is the circle 3 x -\- y 2 z -f 4 =0, 
x 2 + ?/ 2 4- z 2 + 4 x - 6 y - 4 z - 7 = 0. 

10. Determine the equation of the conical surface with vertex at the origin 
whose directrix is the curve of intersection of the plane x -}- 2 y 2 z 4=0 

and the ellipsoid ^ + ~ + ^ = 1. 
4 u 

11. Write the equation of the circular conical surface whose vertex is at 
V(a, /3, 7) and whose directrix is the curve of intersection of the plane ax -f by 
+ cz + d = and the sphere x 2 + y z -f z 2 -f 2 a^x + 2 b } y + 2 r,z + 6?! = 0. 

12. Construct 

(a) the F-contour map of the locus of the equation xy = 2 z. 

(b) the ^-contour map of the same surface. 

(c) the ^-contour map of the surface xy + yz -f- zx 0. 

74. The Intersections of a Surface and a Line. To find the 
points in which a line meets a surface, we have to solve simul- 
taneously three equations; two of these equations are linear, 
namely, the equations of the line, the third may be of any degree. 
The natural way to attack this algebraic problem would probably 
seem to be to solve the linear equations for two of the variables in 
terms of the third, to substitute these values in the third equation 
and to solve the resulting equation for the third variable. For 
example, if we wished to determine the points in which the line 
2 x T/ + 2 4 = 0, + 2?/ 3z + 2 = meets the surface 
x 2 2 xz + 7 y = 0, we would derive from the first two equations 
that z = 5 x 6 and y ~ 7 x 10 ; substitution in the third 
equation would then lead to the equation 9 x 2 + 61 x 70 = 0. 
If this equation were solved for x, the coordinates of the desired 
points could readily be found. 

This method, although frequently useful in special cases, lacks 
symmetry in the treatment of the variables, because it involves 
the selection of one variable (in the example above this was x), 
in terms of which the others are expressed; it casts the three vari- 
ables in non-interchangeable r61es. It has been found that if 
symmetry is maintained, greater elegance and clarity is intro- 
duced in the treatment of algebraic problems. Without going 
any further into a discussion of mathematical esthetics, to which 



150 GENERAL PROPERTIES OF SURFACES AND CURVES 

we are led quite naturally at this point, we shall proceed with the 
problem of finding the intersections of a surface and a line in a 
more symmetric form. 

For this purpose, we take the equations of the line in the para- 
metric form of Theorems 15 or 16, Chapter IV (Section 47, page 
86). Let the equations of the line be 

x = a + \s, y = + fjis t z = 7 + vs 

and let the surface be represented by the equation f(x, y, z) = 0, 
where /(#, y, z) represents a polynomial of degree n in x, y, and z. 
To determine the points common to the line and the surface we 
have now four equations in the four variables x, y, z and s. To 
solve them, we substitute the expressions for x, y y and z given by 
the equations of the line in the equation of the surface. This leads 
to the equation 

(1) f(a + Xs, + p.s, y + PS) = 0, 

which has to be solved for s. When this solution has been ac- 
complished, the coordinates of the required point can be deter- 
mined at once. If we treat the example of the preceding para- 
graph by this method, we derive first the parametric equations 
of the line (see the Examples in Section 47, pages 88 and 89); 
using the form of Theorem 16, Chapter IV, we find for them 

x = 2 + t, y = 4 + 7t, ^==4 + 5 t* 

Substitution of these values of x, y, and z in the equation of the sur- 
face gives the quadratic equation 9 1 2 25 t 16 = 0. From this 

.. . . 25 VI201 , , 61 + VI201 

equation we obtain t = ^ and hence x\ = 

lo 

247 + 7A/1201 197 + sVlloT 




247 - 7\/1201 197 - . . . . . 

2/2 = 75 1 22 = TQ The points of mter- 

lo lo 

* If we put t -7= i we obtain the parametric equations in the s-form 

of Theorem 15 (see Corollaries 2 and 3 of Theorem 10, Chapter III, Section 
34, page 60). In numerical problems, the J-form of the parametric equations 
is usually the more convenient; for theoretical discussions the s-form is 
usually to be preferred. 



DIGRESSION ON TAYLOR'S THEOREM 151 

section of the line and the surface are approximately at PI (5.3, 27.2, 
20.6) and P 2 (1.5, .2, 1.3). 

75. Digression on Taylor's Theorem. The discussion of the 
solution of equation (1) of the preceding section in the case of a 
general polynomial, can be made in a very direct way by means 
of Taylor's theorem,* which takes on a particularly simple form 
for polynomials. This theorem tells us that if F(x, y, z) is a 
polynomial, then 

F(a + h, b + k, c + 1) = F(a, 6, c) + [hFi(a, 6, c) + kF 2 (a, 6, c) 
+ lF*(a, fe, c)] + i (h*F u (a, 6, c) + fc 2 F 22 (a, 6, c) + / 2 F 33 
(a, b, c) + 2 klFn(a, 6, c) + 2 lhF u (a, b, c) + 2 hkF 12 
(a, 6, c)] + - + [AFi(a, 6, c)+kF 2 (a, 6, c)+ZF,(a,6, c)p. 

In this formula Fi(a, &, c), F 2 (a, 6, c) and F 8 (a, &, c) are abbrevi- 

dF I dF I 

ations for the partial derivatives , , 

OX |x=a, y=b, zc vlj \ x = at ys -b t 2==c 

dF I 

respectively; F n (a, b, c), F&, F^, F&, F n , F 12 repre- 

OZ \x=a,y=*b t z=c^ 

sent the second partial derivatives, for example, Fi 2 (a, 6, c) = 

d 2 F I 
--- ; similarly F 3 i2(a> &> c) will be used to designate 

=a f y=6, z=*c 

-a,y-6,i-c' 

We observe that the expression in the second paif of brackets in 
the formula closely resembles the square of that in the first pair 
of brackets, which involves only the first derivatives of F. For 
we see that [hFi(a, 6, c) + kF 2 (a, 6, c) + lF 3 (a, 6, c)] 2 = ft 2 Fi 2 
(a, 6, c) + fcW(a, 6, c) + Z 2 ft 2 (a, 6, c) + 2 WF,(a, 6, c) ft(a, 6, c) 
+ 2 toF 8 (a, 6, c)Fi(a, 6, c) + 2 WfcFi(o, 6, c)F 2 (a, 6, c). The second 
order terms are obtained from this square if we replace F i 2 by FU, 
FiF% by /^i 2 , etc., where these symbols designate second partial 
derivatives in accordance with the notation explained above. 

* For a fuller discussion of this important theorem the reader is referred to 
hooks on the Calculus. 

t Throughout our further work we shall use for partial derivatives the sub- 
script notation which has here been introduced for the arbitrary function 
F(*, V, *) 



152 GENERAL PROPERTIES OF SURFACES AND CURVES 

We shall therefore denote the set of second order terms by the 
abbreviated notation [hFi(a, 6, c) + kF 2 (a, &, c) + lF 9 (a, 6, c)] (2) ; 
and we shall call this a "symbolic square." Now the further terms 
on the right side of the formula which states Taylor's theorem for 
polynomials are the "symbolic cube" the " symbolic fourth power" 
etc., to the "symbolic n-th power" of hFi(a, 6, c) + kF 2 (a, b, c) + IF 3 
(a, b y c), divided by 3 !, 4 ! etc., n\ respectively. They are obtained 
from the ordinary cube, fourth power, to nth power of this ex- 
pression, if products of first partial derivatives are replaced by corre- 
sponding higher partial derivatives; for example, F^FzF* should 



be replaced by ^^ , etc. 

Remark. We notice that each of these symbolic powers are 
homogeneous functions of A, fc, and /. 
Let us consider as an example the function 

F(x, y, z) = ** - 3 x 2 y + 2 yz* + z 2 - 5 x + 3 y - 4. 
Here 



Fu = 6x-Qy, FM = 0, Fzz = 4 y + 2, F 23 = 4 z, F 31 
= 0, F 12 = -6 x; Fm = 6, /^ 2 2 2 = F = 0, F m = ~6, 

Fu3 = F 2n = F 223 = F 33 i = ^231 = 0, /^ 3 3 2 = 4. Hence 
+ kF 2 + IFt = (3x 2 - bxy - 5)A + (-3o: 2 + 2z 2 
+ (4^ + 2 z)Z, 
+ Jfc^j + IF*) = (6 x - 6 ?y)/* 2 + (4 T/ + 2)Z 2 + 8 



kF 2 + ZF 8 )^ = 6 /i 3 - 18 A 2 fc + 12 Z 2 fc. 
We conclude therefore that 

(a + fc) 3 - 3 (a + h)*(b + k) + 2 (6 + fc) (c + Z) 2 + (c + J) 2 
-5(a + h) + 3 (b + fc) - 4 = a 3 - 3 a 2 6 + 2 6c 2 + c 2 + 
3 6 - 5 a - 4 + [(3 a 2 - 6 a6 - 5)h + (-3 a 2 + 2 c 2 +3) 
k + (4 be + 2 c)q + [(3 a - 3 fe)fe 2 + (26 + l)i! 2 + 4 cfcZ 
-6a^] + ^~3/i 2 fc + 2Z 2 fc. 

76. The Intersections of a Surface and a Line, continued. 
We are now prepared for a further discussion of equation (1) of 
Section 74 (see page 150). If we apply Taylor's theorem to the 
left-hand member of this equation, it takes the following form: 



THE INTERSECTIONS OF A SURFACE AND A LINE 153 
/(, ft 7) + [Xtfi(a, ft 7) + fi(a, ft 7) + ?tfa(, ft 7)] + X 



2 > =0. 

But since each of the symbolic powers is a homogeneous function 
(see Remark in Section 75), this equation reduces to an algebraic 
equation in s of degree n : 

/(a, j8, 7) + [Vi + M/2 + /], *, 7 + ^ [X/i + M/2 + "/ 3 P, 0, 7 

+ + ^ ' [\/l + M/2 + /*] (w) ,0,7 = 0, 

wliere the notation a, 0, 7 in the subscript position after a bracket 
serves to indicate that a, ft and 7 are to be substituted for the 
variables x, y, and z respectively within this bracket. We shall 
state our result in the following form : 

THEOREM 5. The parameter values of the points in which the locus 
of the equation /(*, y, s) = 0, whose left-hand side is a polynomial, Is 
met by the line x = a 4 Xs, y = ft 4 M, = T + , are the roots of the 
algebraic equation 

(1) a s n -f ais n ~ l 4- 4- aks n ~~ k 4 ... 4- a M _i* 4 a = 0, 

where 

(3) a n =/(a, 0, T)> n-i = [X/i 4 M/2 4 */j] af/8f7 ... a n -A = ^j 
X [\/i + M/.+ ^C, 7 . , ao = ^ [X/i 4 Mfs + ^li%, 7 ' 

/19/29 and/3 are the partial derivatives of /(*, y, z) with respect to *, y, and 
z respectively, and the symbolic powers of the trinomial are obtained 
from the ordinary, non-symbolic powers by replacing products of the 

dP+q+rf 

form/i/ > / 2 9/3' by the partial derivatives d Pd qd , ' 

COROLLARY 1. A straight line has at most n points in common with 
a surface which is the locus of an equation of degree n in *, y, and 2, 
unless the entire line He in the surface. 

For the equation (1), being of degree n in s, has at most n real 
roots unless it be satisfied by every value of s; and to every real 
root of (1) there corresponds one point that is common to the line 
and the surface. 



154 GENERAL PROPERTIES OF SURFACES AND CURVES 

DEFINITION IV. A surface of order n Is the locus of an equation of 
degree n in *, y, and *. 

Remark. It follows from Corollary 1, in combination with this 
definition, that if a surface is met by no line in more than n points, 
its order is at most equal to n. The number of different points 
which a line actually has in common with a surface depends on 
the number of distinct, finite, real roots of equation (1); in every 
numerical case this number can be determined by the methods 
developed in the theory of algebraic equations. If two or more 
roots of the equation are equal (let their common value be *) 
we say that this number of points of intersection of the line and 
the surface coincide at the point (a + Xs*, ft + jus*, y + vs*). Con- 
versely, the algebraic meaning to be attached to the statement that 
"a line x = a + Xs, y = ft + ps, z = y + vs meets a surface 
f(x y y,z) = in two or more coincident points" is that the equation 
(1) has a double, or a multiple root. To complex roots of the 
equation (1) correspond complex points of intersection of the line 
and the surface, that is, points whose coordinates are complex 
numbers. To values of X, /*, and v for which a = 0, corresponds 
an infinite root of the equation. 

COROLLARY 2. The line whose equations are x = a + \s, y = ft + us, 
s = y 4. vs lies entirely on the surface whose equation is/(*, y, s) = 
if and only if 

/(a, ft 7) = 0, X/i(a, ft y) + M/i(a, 0, y) + tf,(a, ft 7) = 0, 

[X/l + M/2 + y f*\ahy = 0, . . . , [X/i -h M/2 + "/sll%, y = 0. 

This corollary is an immediate consequence of Theorem 5. 
77. Tangent Lines and Tangent Planes. Normals. 

DEFINITION V. A line is tangent to a surface at a point A(a, ft 9 y) if 
at least two of the points of intersection of the line with the surface 
coincide at the point A. 

This definition leads, by way of Theorem 5 (Section 76, page 
153), to the following theorem: 

THEOREM 6. If the point A(a, /s, 7) lies on the surface /(*, y, z) = 0, 
the line * = a + Xs, y = + /*, s 7 4- vs will he tangent to this 
surface at A if and only if the direction cosines x, M v satisfy the con- 
dition: 

(1) A/i(, A 7) + /i(, ft 7) + /a(a, ft 7) = 0. 



TANGENT LINES AND TANGENT PLANES 155 

Proof. Since the point A lies on the surface, the coefficient 
a in equation (1) of Section 76 vanishes; hence s = is one root 
of this equation and to this value of s corresponds the point A on 
the line. The line is tangent to the surface at A if and only if at 
least one other of its intersections with the surface falls at A, 
that is, if the equation has as a root of multiplicity 2 at least 
(see Remark in Section 76). But this is equivalent to the re- 
quirement that a rt _i = [X/i + /i/2 + "/3] a ,&7 = 0; the theorem is 
therefore proved. 

We ask next for the locus of all points P such that the lines PA 
which join P to a fixed point A on the surface shall be tangent to 
the surface at A. If the coordinates of P are x, y, z, the direction 
cosines of AP are proportional to x a, y /3, and z 7 (see 
Corollary 1 of Theorem 6, Chapter III, Section 33, page 56). The 
equation of the required locus is therefore obtained if x a, 
y /3, z 7 are substituted for X, /*, v in condition (1) of Theorem 
6. But the resulting equation is linear in x, y, and z; its locus is 
therefore a plane. The discussion shows that the plane defined 
in the next definition actually exists, if the function /(z, y t z) 
possesses partial derivatives at the point A (a, 0, 7),* not all of 
which are zero. 

DEFINITION VI. A plane Is tangent to a surface at a point A If e?ery 
line in the plane which passes through A is tangent to the surface at 
A, and if every line which is tangent to the surface at A lies in this 
plane* A line through a point A on a surface is normal to the surface 
at A if It is perpendicular to a plane tangent to the surface at A. 

We can now state the following results: 

THEOREM 7. The plane tangent to the surface /(*, y 9 ) = at the 
point A(a, 9 y) on the surface is the locus of the equation 

(x - )/,(, ft r) 4- (y - //.(, ft 7) + (z - 7) Ma, ft 7 ) = 0. 

* The possibility of non-existence of the partial derivatives of a function, 
which is suggested here, need not disturb the reader at the present stage. 
For every function considered in this book, we shall assume that partial deriva- 
tives exist at every point. It must however be recognized that this involves 
an assumption and that the reader is accepting its justifiability on faith. 
For further treatment of this question the reader is referred to books on the 
Theory of Functions of a Real Variable. 



156 GENERAL PROPERTIES OF SURFACES AND CURVES 

COROLLARY. The equations of the normal to the surface /(*, y, *) 
= at the point on the surface A(a, ft, 7) are 

x - a _ y ft = g 7 
/i(i fa 7) /2(f ft T) /a(f ft 7) 

and the parametric equations of the normal are 

ft 7) ', ^ = 7 +/(, A 7) ' 



78. Exercises. 

1. Determine the points in which the surface 2 a: 2 - y 2 -f 4 z 2 + 3 1/2 + 6 zz 
-}- 4 x?/ 2x-\-y 4tZ-}-\ is met by the lines: 

(a) x 9 - 4 , ?/ = 1 - , 2 - -7 + 3 J; 

(6) x = -3 + 2 J, = -13 + 6 J, z = -3 + ; 
(c) x = -3 - 2 , ^ = 9 -f- 5 , z = 7 + 4 J. 

2. Determine the points which are common to the surface y' 2 10 xz 8 yz 
- 12 x - 17 y + 16 z + 30 = and the lines: 

(a) x = 3 - 2 J, y = -1+ 2, 2 = 2 - /; 

(6) x = -1 -f 3 t, y = 3 - t, z = 2 ; 

(c) x = 1 + 3 *, 2/ = -6 , 2 = -3 - 2 ; 

(d) x = 1 - 2 f, ?/ = 1 -h 2 , 2=2-^. 

3. Proceed similarly with the surface 3x 2 4 z* + 3 yz -\- 2 xy -}- 4 x 2y 
+ 4 z + 2 = and the lines: 

(a) a; = 1 + 3 , ?/ = 5 + 4/, 2=6 + 6 1; 

(b) x = -2 + , y = 1 -*, 2 = 2; 

(c) x = 10 *, ?/ = 2 t, z = I7t. 

4. Expand by use of Taylor's theorem: 

(a) 2 (x + h)* - 4 (x -f ft) (y + fc) + 6 (2 + Z) 2 - 3 (x + *) + 5; 

(6) 3 (x - hY -h 5 (x - A) (2 - I) - 4 (y - fc) + 2 (x - A) + 3 (2 - I) - 7. 

6. Determine the condition which the direction cosines of a line through 
the point A(l, - 1, - 1) on the surface 3 x* - 4yz + 2 z 2 - 4x + 2 y + 5=0 
must satisfy in order that the line may be tangent to the surface at A. Write 
the equations for each of two mutually perpendicular tangent lines to the sur- 
face at A. 

6. Set up the conditions which the direction cosines of a line through the 
point A(-l, 4, 3) on the surface 3 x 2 - 2y* -f 3 z 2 - 24 x - 4 y - 12 2 
f 30 = must satisfy in order that the entire line may lie on the surface; 
determine two lines through A which lie on the surface. 

7. Determine the tangent plane and the normal to the surface 2 x 2 y 2 
+ z* 3 zx -\-4xy-\-3x 2y 4 = 0at each of the following points on 
the surface: 

A(l, 1, 1), B(0, 0, -2), C(0, -2, 2). 



THE SHAPE OF A CURVE IN SPACE 157 

8. Write the equation of the tangent plane and the equations of the normal 
to the surface: 

x 3 + 7/ 3 4- z 3 = 1, at the point A(-l, 1, 1). 

9. Determine the condition which the direction cosines of a line through 
the origin must satisfy in order to lie entirely upon the conical surface z 2 4 i/ 2 
-f 4 z 2 2 yz -f- 2 zx + 4 xy = (that is, to be a generator of this surface); 
determine the generators of this surface which lie in the coordinate planes. 

10. Prove that the origin lies on every plane that is tangent to a conical 
surface whose vertex is at the origin. Extend this proposition so as to show 
that the vertex of a conical surface lies on every plane that is tangent to the 
surface. 

79. The Shape of a Curve in Space. To determine the locus 
of a plane curve in space, the methods developed in Chapter V 
arc sufficient. For, by means of them we can transform the frame 
of reference in such a manner as to make the plane in which the 
curve lies one of the coordinate planes in the new frame. When 
this has been accomplished the curve can be studied by the methods 
of Plane Analytical Geometry. 

For curves, whose points do not all lie in one plane (such curves 
are usually called twisted curves, in French courbes gauches) a 
representation by means of a plane drawing can be obtained by 
the use of Descriptive Geometry or of Perspective Drawing. Some 
idea of the shape of the curve can also be obtained from drawings 
which show the projections of the curve on the three coordinate 
planes. 

Certain general properties of the curve can be detected by the 
principles stated in Theorem 4 (Section 70, page 137). There 
should be no difficulty in seeing, for example, that the curve de- 
termined by the equations -j- + ~ + ~ = 1 and z 2 + y 2 = 1, 

4 o y 

which is the intersection of an ellipsoid and a circular cylinder, is 
symmetric with respect to the three coordinate planes, the three 
coordinate axes and the origin. The Z-projection of this curve is 
given by the equations x 2 + y 2 = 1, z = 0; the F-projection by 

x 2 2 z 2 
the equations + = 1, y = 0; the -Sf-projection by the 

1U 1O 

equations-^- - ~ = 1. 
For a more detailed study of the properties of curves and sur- 



158 GENERAL PROPERTIES OF SURFACES AND CURVES 

faces of general character, the reader is referred to treatises on 
Differential Geometry. In the next Chapters we shall take up 
the study of the loci of equations of the second degree in x, y, 
and z. 



CHAPTER VII 
QUADRIC SURFACES, GENERAL PROPERTIES 

Surfaces which are loci of equations of the second degree in 
#, y> and z are surfaces of the second order, see Definition IV, 
Chapter VI, Section 76, page 154; they are usually called quadric 
surfaces or conicoids. A number of such surfaces have already 
been discussed, as to their shape, in Chapter VI (see Sections 68 
and 72). In proceeding to a more detailed study of these surfaces, 
we shall apply to them the results obtained in Sections 76 and 77. 

80. The Quadric Surface and the Line. We shall write the 
general equation of the second degree in the form 



Q(x, y, z) = a n x 2 + any* + a^z 2 + 2 a^yz + 2 a^zx + 
+ 2 dux + 2 a^y + 2 a 34 z + a 44 = 0. 



We shall find it convenient to use the symbol Q(x, y, z), or simply 
Q, throughout as an abbreviation for this general form of the 
function of the second degree in x, y, and z; we shall also use Q 
to designate the quadric surface which is the locus of the equation 
Q(x, y f z) 0. The notations a& and a 32 , si and a J3 , a u and 021, 
an and 4i, 24 and a 42 , a 34 and a will be used interchangeably. 
The partial derivatives of Q will be denoted by the subscript no- 
tation, as already agreed upon in Section 75 (see footnote on page 
151). Moreover we shall use q(x, y, z), or q, to designate the part 
of Q which is homogeneous of the second degree; and q with sub- 
scripts will be used for the partial derivatives of q. Thus we 
have 

q(x, y, z) 



2(a n x + a&y + a 33 3 + a 34 ); 
2(a n x 



Qn = 2 an, q& = Q 2 2 = 2 032, #33 = Qsa = 2 a 33 , 
= 623 2 025, &i = Qsi = 2 ou 9 0* = Qi2 = 2 a u . 
159 



160 QUADRIC SURFACES, GENERAL PROPERTIES 

Furthermore we shall find it useful to use Q 4 as an abbreviation 
for the linear expression 2(a^x + a^y + a 43 z + a 44 ), and corre- 
spondingly # 4 as an abbreviation for the linear homogeneous part 
of Q 4 , that is, for 2(a 41 x + a^y + a^z)\ and the derivatives of 
these expressions will be denoted by the use of a double subscript 
on Q or on q. We observe that the partial derivatives of Q and 
of q of order higher than the second are identically zero. 

From Theorem 5, Chapter VI (Section 76, page 153), we derive 
therefore the following results. 

THEOREM 1. The parameter values of the points In which the line 

x = a + Xs, y = ft -f /us, z = y -f vs meets the quadrlc surface Q(x, y, ) 
= are the roots of the equation Los 2 + 2 LIS -f L 2 = 0, where 

/ - o( H ^ r - 

Lt = Q(a, (3, y), Li = 

= X(a u a -f 0120 -f Oi 

-f v(a z ict -f a 32 /3 -f 0337 4- 834), 

and Lo = ^(X, /i, y) = flnX 2 + 22M 2 -f ass*' 2 -I- 2 a23Aw + 2 asii'X -f 2 a^X/x- 

COROLLARY 1. The line * = a -f Xs, y = /3 -f /us, * = 7 -f ?s will (a) 
meet the quadric surface Q in two distinct real points, if and only if 
Li 2 - Lota > 0; (b) be tangent to Q if and only If Li 2 = L L 2 ; (c) not 
have any real points In common with the surface if and only if 
Li 2 - LoLs < 0. 

COROLLARY 2. The line * = -fXs, y = ft + ius 9 z=y + vs will lie 
entirely on the quadric surface Q if and only if L = L t = L 2 = 0. 

Remark. Values of X, M, v for which L = q(\, n, v) = Q give 
rise to at least one infinite root of the equation. Such values cor- 
respond to lines which meet the surface in one or more infinitely 
distant points. A line for which L\ = L = 0, but L 2 4 1 meets 
the curve in two infinitely distant points. We lay down now the 
following definition. 

DEFINITION I. A direction determined by values of X, /* v which sat- 
isfy the equation q(X, ^, v) =0 is called an asymptotic direction of the 
quadric surface Q. A line which meets a quadric surface in two in- 
finitely distant points is called an asymptote of the surface. 

We can now state a further corollary. 

COROLLARY 3. The necessary and sufficient conditions that the line 

x = a + \s, y = ft + us, s = y + vs be an asymptote of the quadric 
surface Q are that ?(x, ^, v) = and that \Qi(<x 9 , 7) + 
7) = 0. 



TANGENT LINE; TANGENT PLANE; NORMAL, ETC. 161 

81. Tangent Line; Tangent Plane ; Normal; Polar Plane. If 

the discussion of Section 77 be applied to the quadric surface Q, 
the following results will appear without difficulty. 

THEOREM 2. If the point A(a 9 (3, y) lies on the quadric Q, the line 
x = 4- \s 9 y = -f M> * = T -f " will be tangent to the surface at ^ 
if and only if its direction cosines x, //, v satisfy the equation \Qi (, 0, 7) 

+ /i<M 0> 7) -f *<?3(a, 0, 7) = 0. 

Remark. This result can also be obtained from Corollary 1 of 
Theorem 1. 

THEOREM 3. The equation of the plane tangent to the quadric sur- 
face Q at the point A(a 9 p, y) on the surface is (x - a) Q\(a 9 0, 7) 
+ (y - 0) <M A 7) + (* - 7) <M 0, 7) = 0. 

COROLLARY. The equations of the normal to the quadric surface Q 
at the point A(<* 9 0, 7) are x = a -f (),( 0* 7) * y = + (M A 7) t, 

* = 7 + <?i( A 7) ' t. 

The equation of the tangent plane to the quadric surface, given 
in Theorem 3, can be put in a form which is very convenient for 
application in numerical cases. We observe first that Q(a, ($, 7) 

= q(a, j3, 7) + 2 a^a + 2 034^ + 2 a^y + a 4 4 and that therefore 
Qi = q l + 2 ai 4 , Qz = q* + 2 a 24 , Q 3 = g s + 2 a 34 . 

And it follows from the notation introduced in the first paragraph 
of Section 80 that Q 4 = q* + 2 a 44 . Furthermore, since q is a hom- 
ogeneous function of the second degree, it follows* that aqi(a, 0, 7) 
+ Pq*(ct f ft 7) + yq*(<* 9 ft 7) = 2 q(a, ft 7). If we add 
2(2 a i4 a + 2 a 24 + 2 a 84 7 + 044) = 2 g 4 (a, ft 7) + 2 a 44 = 2 Q 4 (a, ft 7) 
2 a 44 to both sides of the last equation, we find that 

oQi(a, ft 7) + 0Q 2 (a, 0, 7) + 7<Me*, ft 7) + Q 4 (a f ft 7) 
= 2<2(a,ft T ). 

The left-hand side of the equation of the tangent plane, as given 
in Theorem 3, may now be transformed as follows, remembering 

* We are here making use of the following theorem, known as Euler's 
theorem on homogeneous functions: If F(x, y, z) is a homogeneous function 
of degree n in x, y, and z, then xFi -f- yF* + ^F 3 = nF. The proof of this theo- 
rem may be made as follows: The homogeneity of F(x, y, z) tells us that 
F(kx, ky, kz) = k n F(x, y, z) for every k. Differentiation with respect to A; 
gives xFi(kx, ky, kz) + yF 2 (kx, ky, kz) + zF^kx, ky, kz) = nk~ l - F(x, y, z), 
from which the formula of the theorem is obtained if we substitute 1 for k. 



162 QUADRIC SURFACES, GENERAL PROPERTIES 

that Q(, ft 7) = 0: 

(x - )&(, ft 7 ) + (y - 0)Q 2 (a, ft 7) + (* - 7)Qs(, ft 7) 



0220 + 0237 + ^24) + 2(a 3 ia + 320 + ^337 



If we remember the convention concerning the coefficients a#, 
which was made in the first paragraph of Section 80, we have the 
following result : 

THEOREM 4. The tangent plane to the quadrlc surface Q at the point 
A (a, 0, 7) on the surface, has the equation 

ana* 4- 0220y 4- 0337* 4- 023(0* + yy) 4- 031(7* 4- *) 4- oiz(ay -f fix) 
+ ai 4 (* -f a) 4- a 2 4(y + /3) -h 34( -f 7) 4* 44 = 0. 

Remark. The reader will observe that this form of the equation 
of the tangent plane is obtainable from the equation of the quadric 
surface by a process which consists in distributing the coefficients 
of the equation among the variables x, y, z and the constants 
a, ft 7 in equal shares; the technical name of this process is 
" polarization. " 

COROLLARY 1. If , 0, 7 are selected entirely arbitrarily, we have the 
relation 

<?i(f V* T) 4- <? 2 (, 0, 7) 4- 7<?( 0, 7) 4- <M> ft 7) = * (>(> ft 7). 



The equation in Theorem 4 represents a plane whether A (a, 0, 7) 
is on the quadric surface Q or not. For a general position of the 
point A, this plane is called the polar plane of A with respect to 
the surface. 

DEFINITION II. The plane which Is represented by the equation 

auax 4- <*vpy 4- a337* 4- 23(/3* 4- yy) 4- 031(7* 4- as) 4- ai 2 (ay 4- px) 4- 
OH(* 4- a) 4- 24(y 4~ 0) 4~ 034 (* 4~ 7) 4- 044 = Is the polar plane of the 
point A(a, 0, 7) with respect to the quadric surface Q; the point Is 
called the pole of the plane with respect to the surface. 

This definition enables us to state a further corollary of Theorem 
4, namely, 

COROLLARY 2. The tangent plane to the quadric surface Q at the 
point A (, 0, 7) on the surface coincides with the polar plane of A 
with respect to the quadric surface. 

Thus we have obtained a geometrical interpretation of the polar 
plane of a point on the surface Q with respect to this surface. We 



TANGENT LINE; TANGENT PLANE; NORMAL, ETC. 163 

proceed next to inquire as to the geometrical significance of the 
polar plane with respect to the surface Q of a point which is not 
on this surface. 

We observe first that, in view of the proof of Theorem 4, of 
Corollary 1 of this theorem, and of Definition II, the equation of 
the polar plane can be written in the form 

xQv(a, ft 7) + yOt(, 7) + *Qi(, ft 7) + <M, ft 7) = 



or in the equivalent form 

(1) (x - a)Qi(a, ft 7) + (if - &(, ft 7) + (* - 7) 
Qi(a, ft 7) + 2 Q(a, ft 7) = 0. 

Let us now consider an arbitrary line I through the point 

A (a, ft 7): 

x = a + \s, y = ft + ps, z = 7 + j/s; 

let the points where the line i meets the quadric surface be A\ and 

Az ; and let its point of intersection with the polar plane (1) of A 

with respect to the surface be 

A' (see Fig. 31). It follows 

then from the geometrical 

meaning of the parameter s in 

the equations of the line I (see 

Corollary 2 of Theorem 10, 

Chapter III, Section 34, page 

60) that the directed segments 

AAi and AA 2 are equal to the 

roots of the equation Lo 2 + 

2LiS + L 2 = 0, established in FIG. 31 

Theorem 1; and that the di- 

rected segment A A' is equal to the root of the equation 

XQi(, ft 7) + /*&(, ft 7) + *&(, ft 7) + 2 Q(a, 0, 7) = 0, 




obtained by eliminating x, y, and z between the equations of the 
line I and the equation of the polar plane (1). Now we have to 
recall the meaning of the coefficients L , LI and L 2 stated in 
Theorem 1, and also the relations between the coefficients and the 

roots of an algebraic equation (the sum of the roots of the 
equation ax 2 + bx + c = is equal to , their product is equal 



164 QUADRIC SURFACES, GENERAL PROPERTIES 

to - Y With the aid of these tools, we find the following result: 



AA' = - 



+ M& + 



2L 2 L 2 /L 



: = 2 



+ AA 2 

From this relation we derive two consequences: 

(a) If we divide both sides by two and take the reciprocals of 

211 
both sides of the equation, we find that -r-j, = -r-r- + -r-r- , or that 



1111 



AA' AA 1 AA 2 ' 

This informs us that, independently of the direction of the line I, 
the reciprocals of the segments AAi, A A', q,nd AA 2 form an arith- 
metical progression, that is, the segments AAi, AA', and AA 2 form 
a harmonic progression, so that A A f is the harmonic mean between 
AAi and AA 2 . 

(6) If we transform the relation by writing AA f = AAi + A\A' y 
clearing of fractions and carrying out the indicated operations, we 
find that 



+ AA l AiA' + AA 2 A^' = AA l AA 2 ; 

that is, 

AAi(A 2 A + AA 1 + A,A') + AA 2 A,A' = 0, 
A 2 A' + AA 2 - A,A' = 0, 



fi i 
or finally --. 



This equation expresses the fact that the points A and A r divide 
the segment AiA 2 in ratios which are equal numerically, but oppo- 
site in sign. This is what is meant by the statement that A' and 
A are harmonic conjugates with respect to the points AI and A 2) 
according to the following definition: 

DEFINITION III. If A, B, C, and D are collinear points and so situ- 
ated that the ratio of the segments CA and AD, in which A divides the 
segment CD, is equal numerically but opposite in sign to the ratio 
of the segments CB and BD in which B divides the segment CD, then 
A and B are called harmonic conjugates with respect to C and D. 



POLAR PLANE AND POLE 



165 



It should be clear from the preceding discussion that if A and 
B are harmonic conjugates with respect to C and D, then the 
segment AB is a harmonic mean between the segments AC and 
AD, and conversely. Furthermore, if B is the harmonic conjugate 
of A with respect to the intersections of the line AB with the 

2 Q 1 

quadric surface Q(x, y, z) =0, then AB = ^ ; ^ ; 7r 



in which X, /*, v are the direction cosines of the line AB. 
from this that the coordinates of B are 
2 XQ 



It follows 



+ 



+ 



ya = P - 

2* = 7 ~ 
Consequently 



+ M? 2 + 



+ 



(, ft 7) 



+ 



+ MQ 2 + 



,ft7)+ (*B - 7)Q 3 (, ft 7) 
= -2Q(a,0, 7), 



from which we conclude that the coordinates of B satisfy the equa- 
tion (1) of the polar plane. We can summarize our discussions 
by the following theorem. 

THEOREM 5. The polar plane of a point A with respect to a quadric 
surface is the locus of the harmonic conjugates of A with respect to 
the intersections of the surface with the lines through A. 

82. Polar Plane and Pole. Tangent Cone. Preliminary to a 
discussion of some further properties of the polar plane we raise 
the question whether it is possible for a plane to have more than 
one pole with respect to a quadric surface, that is, whether it is 
possible for two different points A (a, 0, 7) and A* (a.', &', y') to 
have the same polar plane with respect to the surface. In view of 
Definition II this amounts to the question whether the equations 
(ana + a^ft + a^y + a M )x + (a u a + a<np + 0237 + ^24) y + 

044) 



= 0, 

(ana' + ai 2 /3' + 0137' + 
y + (aisa' + 0280' + 
+ 044) = 



+ (a^af + 0%$' + a^y' + 024) 
+ 034)2 + (ana' + <h&' + 



166 QUADRIC SURFACES, GENERAL PROPERTIES 

can represent the same plane. This will be the case if and only if 
the coefficients of x, y, and z and the constant terms in the two 
equations differ by a factor of proportionality, k (see the footnote 
on page 78), that is, if and only if the numbers ka a', kft /3', 
ky y' and k 1 satisfy the following system of linear homoge- 
neous equations: 

an(ka - a') + a 12 (/c/3 - 0') + a 13 (fc T - T') + u(fc - 1) = 0, 
a 12 (ka - a') + a*(W - (?) + a*(ky - 7') + a,,(k - 1) = 0, 

- cf) + a 23 (/c/3 - 0') + a,,(ky - 7') + <to(k - 1) = 0, 

- a') + 024(fc0 - 0') + au(ky - 7') + a u (k - 1) = 0. 

This system of equations will or will not possess a nontrivial 
solution, according as its coefficient determinant has a value that 
is equal to or different from zero (see Theorem 2, Chapter II, page 
38). In the latter case, the only solution is the trivial one, so that 
we find ka a' k/3 /3' = ky y' = k 1 = ; from this we 
find k = 1, a = a', = 0', 7 = 7', so that no plane can have more 
than one pole. The discussion leads to the following definitions 
and conclusions: 

DEFINITION IV. The determinant of the symmetric square matrix 

1 1 ay 1 1, i, j = 1, 2, 3, 4, dij = aji, is called tht discriminant of the 
quadrlc surface Q; we shall use A to designate the value of this 
determinant. 

DEFINITION V. A quadric surface whose discriminant vanishes is 
called a singular quadric surface. (Compare Definitions II and III on 
page 43.) 

THEOREM 6. No plane has more than one pole with respect to a 
non-singular quadric surface. 

Remark 1. The homogeneous equation of the second degree in 
x, y, and z represents a conical surface of the second order; it is 
called a quadric cone. The quadric cone is a singular quadric 
surface, for in its equation a J4 = 24 = 34 = #44 = 0; consequently 
the value of the discriminant is zero. 

Remark 2. While Theorem 6 asserts that no plane has more 
than one pole with respect to a non-singular quadric surface, it 
does not say that there exists a pole for every plane in space. The 
question depends on whether the system of equations 

n + #120 + ia7 + i4 = ka, 

O>12<X + 220 "T" #237 + 24 = kb, 



0240 + 0347 + 044 = 



POLAR PLANE AND POLE 167 

does or does not have a solution for a, /3, 7, and k for every set of 
values of a, 6, c, and d. Since for a non-singular quadric the aug- 
mented' matrix of this system is always of rank 4, it follows from 
Theorems 1 and 8 of Chapter II (see Sections 21 and 27, pages 36 
and 44) that a plane will have a single pole or none with respect to 
a non-singular quadric surface, according as the value of the 



determinant 



0,12 #22 #23 



is different from or equal to zero. 



#13 #23 #33 
#14 #23 #34 

We derive now a number of further consequences from the defi- 
nition of the polar plane. 

THEOREM 7. If A(a 9 /3, 7) lies on the polar plane of A'(a' 9 ', 7') with 
respect to a quadric surface, then A' lies on the polar plane of A with 
respect to the same surface. 

The proof of this theorem is left to the reader. 
THEOREM 8. If A(a 9 0, 7) lies on its own polar plane with respect to 
a quadric surface, then A lies on the surface; and conversely. 

The first part of this statement becomes evident when we sub- 
stitute the coordinat >l of A in the equation of the polar plane of A ; 
the second part follows from Corollary 2 of Theorem 4 (Section 
81, page 162). 

Suppose now that from a point A not on the quadric Q tangents 
be drawn to the surface ; let the points of contact of these tangents 
be A', B', etc. Then, since A lies in the planes tangent to the 
surface at A', B' y etc., A lies in the polar planes of A', B', etc., with 
respect to the surface. It follows therefore from Theorem 7, that 
the polar plane A passes through A 1 ', J5', etc. Conversely, if A' 
is a point in which the polar plane of A meets the surface, then the 
tangent plane to the surface at A' passes through A, that is, the 
line A A' is tangent to the surface. We have therefore the follow- 
ing theorem: 

THEOREM 9. The points of contact of a quadric surface Q with the 
tangents drawn from a point A not on the surface, are the points in 
which the surface is met by the polar plane of A with respect to Q. 

On the basis of this result we introduce the following definition: 

DEFINITION VI. The tangent cone from a point A to a non-singular 
quadric surface which does not contain A is the cone whose vertex is at 
A and whose directrix is the curve common to the surface and to the 
polar plane of A. 



168 QUADRIC SURFACES, GENERAL PROPERTIES 

Remark. This tangent cone is a quadric cone ; for the curve in 
which a plane meets a quadric surface can not be of higher order 
than the second; and since A does not lie on the surface, it 'can not 
lie on its own polar plane. 

To obtain the equation of the tangent cone, we shall use two 
methods. 

(a) In the first method, we translate the frame of reference to 
A as a new origin, set up the equation of the cone in the new system 
and then return to the original frame. Since the transformation of 
coordinates has been discussed in Chapter V, we shall suppose that 
the translation has already been carried out and we shall ask 
therefore for the tangent cone from the origin to a quadric surface 
Q which does not pass through the origin. 

It follows from Theorem 9 that the cone passes through the points 
common to the surface and the polar plane of the origin; conse- 
quently its equation can be written in the form 

(1) kiQ(x, y y z) + fc 2 (a 14 x + a^y + a 34 z + a 44 ) = 



(see Remark 4, Section 49, page 92). The multipliers ki and k% 
have not been restricted in any manner as yet; since we are look- 
ing however for a quadric cone, they have to be selected in such a 
way that equation (1) shall be a homogeneous equation of the sec- 
ond degree in x, y, and z. Therefore we put k\ k and k% 
= Ix + my + nz + p and we write down that the constant term 
and the coefficients of the first degree terms in equation (1) must 
vanish; this leads to the following conditions: 

(2 k + p)a l4 + I(i 44 = 0, (2 k + p)a 24 + ma 44 = 0, 
(2 k + p)a 34 + nau = 0, fca 44 + pa 44 = 0. 



Since the origin does not lie on the surface Q, a 44 4= 0; hence the 
last equation gives p = k. Substituting this value for p in the 
other equations, we find Za 44 = fca i4 , ma 44 = fca^ and na 44 
= fca 34 . We are therefore free to choose an arbitrary non-zero 
value for fc, as could be expected from the fact that equation (1) 
could have been divided through by k\ without essentially chang- 
ing anything. Taking k = a^, we obtain I = a\\, m = 024, 
n = 034, p = a 44 . Consequently the tangent cone from the 
origin to the non-singular quadric Q is represented by the homo- 
geneous equation a,uQ(Xj y, z) (dux + a^y + a^Z + a^) 2 = 0. 



POLAR PLANE AND POLE 169 

(6) In the second method w suppose that P(z, ?/, z) is an arbi- 
trary point on the tangent cone. Then the line AP is tangent to 
the surface and its direction cosines must therefore satisfy the 
condition of Corollary 1 of Theorem 1 (Section 80, page 160), 
namely: 

1 [XQi(, 19, 7) + M<Ma, 18, 7) + KM, ft 7)] 2 = Q(, ft 7) X 
(an\ 2 + a 2 2M 2 + 33^ 2 + 2 a 2 3M^ + 2 a 3 i^X + 2 Oi 2 X/i). 

But for the line -AP, we have \:p:v x a :y ft :z y ; and 
since the condition which we have just written down is homoge- 
neous of the second degree in X, /z, and v, we may omit the factor of 
proportionality. We obtain therefore for the required tangent 
cone the following equation homogeneous of the second degree in 
x dj y ft, and z 7 : 

\ K* - )Qi(, ft 7) + (If - PXMa, ft 7) + (s - 7XM, ft 7)] 2 
= Q(, ft 7) [au(x - a) 2 + a 22 (7/ - 0) 2 + 033(2 - 7> 2 
+ 2 023(0 - 0) (2 - 7) + 2 031(3 - 7) (x - a) + 
2 a w (x - a) (T/ - j8)]. 

Remark. For a: = = 7 = 0, this last equation should be 
equivalent to the equation obtained by the first method. 

83. Exercises. 

1. Determine the equation of the tangent plane and the equations of the 
normal for the surface 4 x- 6 xy -f 5 y 2 -f- 4 yz 3 z 2 -\- '2 zx 4 x -{- 3 y 
4- 2 z -h 4 = at the point A(l, -1, 2). 

2. Set up the condition which the direction cosines of a line through 
P(2, 1,1) must satisfy in order to be tangent to the surface 3 x\ 2 t/ 2 + 
5 zx -4 y 4-6^-3=0. 

3. Set up the condition on X, ju, v under which the line x 1 + Xs, 
y = 2 + jus, ^ = 2 4- J>s will lie entirely on the surface 4 z 2 6 y 2 -f- 8 z 2 
= 12. 

4. Determine the polar plane with respect to the ellipsoid 3 x 2 -f- 2 ?/ 2 
-f 4 z 2 = 20 of the points A(-2, 2, 1), B(5, 5, 0), C(0, 4, -3), D(0, 0, 0). 

5. Find the pole with respect to the surface 3 x 2 2 xy -\- y 2 -f 4 yz 6 x 
-f 2 i/ + 7 = of the plane (a) z + y + - 3 = 0; (&)2s-t/-r-2z-f- 
3=0. 

6. Derive the equation of the tangent cone to the surface 4 x 1 -f 3 y 2 - 12 z 
from the points A(0, 0, -6), (-4, 5, 3), C(0, 0, 4). 

7. Determine the equation of the tangent cone from an arbitrary point 

x 2 v 2 z 2 
P(a, /3, 7) to (a) the ellipsoid i +#5 +-5 = 1; (b) the hyperboloid of one 

x 2 v 2 z 2 a o c ^ 2 2 

sheet -5 -h jr, i = lJ ( c ) the elliptic paraboloid + p = 2 pz. 



170 QUADRIC SURFACES, GENERAL PROPERTIES 

8. Show that the equation for the tangent cone obtained by the second 
method of the last part of Section 82 reduces to the result found by the first 
method if we put a = ft = 7 = 0. 

9. Determine the asymptotic directions of the hyperboloid of one sheet 

x 2 y 2 z* 

7- TT + |T = 1, which lie in the plane 3 x 2 y = 0. 

10. Show that the asymptotes of an hyperbola are also asymptotes of the 
hyperboloid of revolution of one sheet which is obtained when the hyperbola 
is revolved about the conjugate axis. 

11. Show that the ellipsoid of Exercise 7 does not have any real asymptotic 
directions. 

~2 ? .2 2 

12. Determine whether the hyperboloid of two sheets ~ ~ - 2 =1, 

a o c 

x 2 y 2 
the elliptic paraboloid of Exercise 7 or the hyperbolic paraboloid ~ = 2 pz 

have real asymptotic directions. 

84. Ruled Quadric Surfaces. We have already met a few 
examples of surfaces which contain every point of a line (see, for 
example, Exercise 3 in Section 83). We proceed now to a sys- 
tematic study of the question which quadric surfaces have straight 
lines on them. We saw in Corollary 2 of Theorem 1 (Section 80, 
page 160) that the line x=*a + \s 9 y = P + ns,z = v + vs will 
lie entirely on the quadric surface Q if and only if L = q(\, M> v) 
= flnX 2 + &22M 2 + 33^ 2 + 2 OKIJLV + 2 a 3 i*>X + 2 a^X/x = 0, 

Ja = i [XQi(a, ft 7) + MQ 2 (, ft 7) + "Qs(, ft 7)] = 
and L 2 = Q(a, ft 7) = 0. 

Suppose now that we have a point A(a y ft 7) on the surface, so 
that the condition L 2 = is satisfied. To determine lines through 
A which lie entirely on the surface, X, /*, and v have to be so se- 
lected that LI = Z/o = 0. These equations are homogeneous in 
X, M, and v and of degree 1 and 2 respectively; if we solve them 
for two of the variables, say X and M> in terms of the third variable 
v, we shall in general be led to a quadratic equation, giving rise 
to two values for the ratios X : n : v, that is, to two lines on the 
surface through A. These two values may be real and distinct, 
coincident or imaginary. We want to learn under which condi- 
tions these different situations will arise. 
We begin by proving the following auxiliary theorem. 

THEOREM 10. A non-singular quadric Q contains no point at which 
Qi = <?2 = Q* = 0. 



RULED QUADRIC SURFACES 171 

Proof. It follows from Corollary 1 of Theorem 4 (Section 81, 
page 162) that at a point at which Q Qi = Q 2 = Q 3 = 0, we 
must also have Q 4 = 0. Hence there would be at least one set of 
three numbers a, 0, 7 which satisfy the four linear non-homogeneous 
equations a^a + a^ + 8 7 + 0*4 = 0, i = 1, 2, 3, 4. If the 
rank of the coefficient matrix of this system of equations is 3, a 
solution exists if and only if the rank of the augmented matrix is 
also 3 (see Theorem 8, Chapter II, Section 27, page 44). But the 
determinant of this augmented matrix is the discriminant of the 
quadric surface (see Definition IV, Section 82, page 166); hence 
if the rank of the augmented matrix is 3, the surface is singular. 
On the other hand, if the rank of the coefficient matrix of the 
system of linear equations is less than three, then the cofactors of 
the elements in the last column of the discriminant are all equal 
to zero and the value of the discriminant is therefore zero. Thus 
we have seen that if a quadric surface contains a point at which 
Qb Qz an d Qz all vanish, then it is singular. 

We shall have frequent occasion to refer to a point on a quadric at 
which these conditions hold and we introduce therefore a name for it. 

DEFINITION VII. A vertex of a quadric surface Q is a point at which 
Q = <?i = <? 2 = <? 3 = 0. 

In the terminology of this definition, we can then state the 
following corollary: 

COROLLARY. A non-singular quadric surface has no vertex; a singu- 
lar quadric surface may have one or more vertices. 

We proceed now with the problem of determining lines through a 
point A (a, 0, 7) on a quadric which shall lie entirely on the surface; 
and we shall divide our discussion of this question in two parts: 
CASE I. The point A (a, 0, 7) is not a vertex. 

In this case at least one of the partial derivatives of Q is different 
from zero at A ; let us suppose that Qi(a, p, 7) =t= 0. We can then 
solve the equation LI == for X in terms of M and v and substitute 
the result in the equation L = 0. This will lead us to the follow- 
ing quadratic equation in /* and v:* 



(1) 



dll 012 Ql 
012 022 Ql 

Qi Q 2 



011 012 Ql 
013 23 Q 
Ql <?2 



011 013 Ql 
013 033 Q3 

Qi Q 3 



= 0, 



* In order not to interrupt our main argument too much at this point, we 
relegate the proof of this statement to the Appendix, I (see page 296). 



172 



QUADRIC SURFACES, GENERAL PROPERTIES 



(2) 



in which the partial derivatives Q\ 9 Q 2 , and Qz have the arguments 
a, 0, and 7; it will be understood that this is the case throughout 
our further argument, unless the contrary is definitely specified. 

We observe now that the coefficients of v?, 2 pv, and j> 2 in this 
equation are equal respectively to the minors of the elements 
033, 23> and 022 in the determinant 

011 01 2 013 *fcl 

01 2 22 023 Qz 

013 023 033 Qs 

If the value of this determinant is denoted by A^(Q) and the co- 
f actors of its elements a# by Aij(Q), it follows from Theorem 18, 
Chapter I (see Section 16, page 29), that the discriminant of equa- 
tion (1) is equal to ^4 23 2 (Q) A^(Q) X A^(Q) 

\A^(Q) A&(Q) 3 Qi 

Consequently the roots of the quadratic equation (1) will be real 
and unequal, real and equal or complex according as the value of 
the determinant (2) is positive, zero or negative. We will show 
now that this determinant can be reduced to a simpler form. If 
we add to the last column the products of the 1st, 2nd, and 3rd 
columns by 2 a, 20, and 27 respectively, its elements will 
become 2 an, 2 024, 2 as4 and 2 aQi 2 0Q 2 2 7^3. Next, we add 
to the last row the products of the 1st, 2nd and 3rd rows by 
2 a, 20, and 27 respectively; this transforms the elements 
of the last row into 20 U , 2a 24 , 2034 and 2aQi 20Q 2 27Q 3 
4 0i4 4 00554 4 7034. This last element, in the lower right- 
hand corner, is equal to 2(aQi + 0Q 2 + 7^3 + Q\) + 4 a 44 
= 4(044 Q), by use of Corollary 1 of Theorem 4 (Section 81, 
page 162). Therefore the determinant (2) has been reduced to 



011 012 


013 


2 


014 






0n 


012 


013 


2014 


01 2 022 


023 


2 


024 






012 


022 


023 


20 24 


013 023 


033 


2 


034 






013 


023 


033 


2034 


2014 20 24 


2034 


4(044 


-Q) 




2014 


2024 


2034 


4044 




011 


012 


013 

















01 2 


022 


023 

















013 


023 


033 

















2014 


2024 


2034 




-4Q 











RULED QUADRIC SURFACES 173 

Therefore, if, in agreement with our general notation, we designate 
by ^44 the cofactor of the element a 4 4 in the discriminant A of the 
quadric surface, we obtain the interesting formula 

(3) A 9 (Q) = 4A - 4^44 -Q(, 0,7). 

In the particular case which we are having under consideration, 
Q(ot y fi, 7) = 0, since the point A (a, 0, 7) lies on the surface and 
therefore A 3 (Q) = 4 A. We state this preliminary result of our 
discussion in the following theorem and corollary. 

THEOREM 11. If the point A(a 9 , 7) lies on the quadric surface <?, 
then the value of the determinant A*(Q) Is independent of the position 
of A on the surface and equal to four times the value of the discrimi- 
nant A of the surface. 

COROLLARY. The matrices of the determinant A t (Q) and of the dis- 
criminant A have equal rank, if the point A (a, 0, y) lies on the quadric 
<? 

Proof. This Corollary follows from Theorem 14, Chapter I, in 
view of the fact that if Q(a, 0, 7) = 0, the matrix of the determi- 
nant Az(Q) is transformed into that of the discriminant A by 
means of elementary transformations (see Definition XIV, Chap- 
ter I, Section 10, page 18). 

The further discussion of our problem depends on the rank of 
the matrix of the discriminant A ; henceforth we shall denote this 
matrix by the symbol a 4 and its rank by r 4 . We consider now the 
following possibilities: 

(a) n = 4, that is, A 4= 0. 

It follows from our discussion that in this case the quadratic 
equation (1) will have two distinct roots, which are real if A > 
and complex if A < 0; to each root of the equation (1) there 
corresponds a set of direction cosines of a line through A which will 
lie entirely on the surface. We can conclude therefore that 
through every point on a quadric surface for which A > 0, there 
pass two different real lines which lie entirely on the surface ; and 
that through no point on a surface for which A < there are lines 
which lie on the surface. 

(b) n = 3. 

Asa result of the Corollary of Theorem 11, the rank of the 
matrix of the determinant (2) will also be equal to 3 in that case. 



174 QUADRIC SURFACES, GENERAL PROPERTIES 

We can show now that the coefficients of equation (1) can not all 
vanish, by showing that if they did, then every three-rowed minor 
of the determinant (2) would vanish.* Consequently in this case 
the equation (1) has two coincident roots and through every point 
of the surface, which is not a vertex, there will pass two coincident 
lines which lie entirely on the surface. 

(c) r 4 = 2 or 1. 

It follows now from the Corollary of Theorem 11, that all the 
coefficients of the equation (1) vanish. Consequently every line 
through the point A whose direction cosines satisfy the condition 
Li = lies entirely on the surface. But this carries with it that 
every line through A which lies in the plane (x a)Qi(a, ft 7) 
+ (y - 0)Q 2 (a, ft -V) + (* - 7)Q 3 O, ft 7) = must lie on the 
surface. We conclude that the surface contains every point of this 
plane; in virtue of Corollary 1 of Theorem 4 (Section 81, page 
162) and because Q(a, ft 7) = 0, the equation of this plane may 
also be written in the form xQi(a, ft 7) + yQ 2 + zQ s + $4 = 0. 
CASE II. The point A (a, ft 7) is a vertex of the surface. 

In this case, which can arise only on a singular quadric (see 
Corollary of Theorem 10, page 171), the equation LI is satis- 
fied by every set of direction cosines. And we shall show that the 
condition L = q(\ M> v) = is satisfied by the direction cosines of 
any line which joins A (a, ft 7) to another point A'(a f , /3', 7') on 
the surface and by no others. For, in virtue of Taylor's theorem 
(see Section 75, page 151) we have 

Q(', 0', 7') = <?( + [' - ], + W - ft], y + [V ~ 7]) 
= Q(, ft 7) + [(' - )0i(a, ft T) + (0' - j8)Q a (a, ft 7) 
-7XM, ft 7)] + (' - a, 0' - ft 7' - 7). 



Therefore, if A (a, ft 7) is a vertex of the surface and if A'(a x , ($', 7') 
is an arbitrary second point on the surface, then q(a f a, 0' ft 
7' 7) = 0- But since the direction cosines of the line A A ' are 
proportional to a' a, /3' ft and 7' 7, and since q is a homo- 
geneous function, it follows that q(\, /z, v) 0. And it should be 
an easy matter to show that this will not be the case for the di- 
rection cosines of a line which connects the vertex A with any other 
point in space. 

* The proof is given in the Appendix, II (see page 296). 



RULED QUADRIC SURFACES 175 

The results of the discussion of our problem may now be sum- 
marized as follows. 

THEOREM 12. Through every real point A(a, 0, 7) on a non-singular 
quadric surface with positive discriminant, there pass two and only 
two lines which He entirely on the surface; through no real point on a 
non-singular quadric surface with negative discriminant is there any 
line which lies entirely on the surface. Through every point on a 
singular quadric for which the rank of the matrix of the discriminant 
Is 3, and which is not a vertex of this surface, there pass two coincident 
lines which lie entirely on the surface, and no others. Through every 
point on a singular quadric for which the rank of the discriminant 
matrix Is less than 3, and which is not a vertex of this surface, there 
passes a plane which belongs entirely to the surface. The lines joining 
a vertex of a singular quadric to any other point on the surface lie en- 
tirely on the surface. 

COROLLARY 1. A singular quadric surface which possesses a vertex 
is a conical surface; it is a quadric cone. 

For, from the last part of Theorem 12, it follows that the surface 
may be generated by a line through a vertex which moves so as to 
pass through the points of the surface cut out by any plane which 
does not pass through the vertex. 

Remark. A vertex of a singular quadric is also a vertex of the 
quadric cone which it represents. 

COROLLARY 2. If the rank of the discriminant matrix of a quadric 
surface is less than 3, the locus of the equation consists of two planes; 
It is a degenerate quadric. (See Definition V, Chapter IV, Section 46, 
page 83.) 

Proof. For it follows from Theorem 12 that in this case there 
is at least one plane all of whose points lie on the surface. Let 
the left-hand side of the equation of this plane be E\ and let 
Q = E Ei + R, where R is a function of y and z alone. Ob- 
viously if R does not vanish identically we can determine particular 
values of y and z for which R 4 1 0; and it will also be possible in 
general to associate with these values of y and z a value of x such 
that these values of x y y, and z cause E to vanish. But for these 
values, we will have Q =f= 0; and therefore, we would have a point 
on the plane E = which does not lie on the surface. This con- 
tradicts our hypothesis. Consequently, R must vanish identically, 
and Q = E EI. It is now easy to see that E\ is also a linear func- 
tion and therefore we conclude by use of Theorem 10, Chapter IV 



176 QUADRIC SURFACES, GENERAL PROPERTIES 

(Section 46, page 83) that the locus of Q = consists of two 
planes. 

85. The Centers and Vertices of Quadric Surfaces. 'Among 
the particular quadric surfaces with which we have already be- 
come familiar are the sphere, the elliptic cylinder and the circular 
cone. The center of a sphere is usually defined as the point from 
which all the points on the sphere are equally distant; for an 
elliptic cylinder, and even for a circular cylinder such a point does 
not exist. If we take for the center of the sphere however the 
property that it bisects every chord which passes through it, we 
observe that every elliptic cylinder has points which possess the 
same property, namely, the points on its axis. The axis of such a 
cylinder could then be called a line of centers. But even on this 
definition of a center, the cone, the elliptic paraboloid and 
other quadrics do not possess any centers. We undertake there- 
fore in the present section the inquiry as to the conditions under 
which a quadric has a center; and we shall seek to develop con- 
venient methods for the location of centers in the cases in which 
they exist. Our work will be based on the following definition: 

DEFINITION VIII. A center of a quadric surface Is a point which bi- 
sects every chord drawn through It;* a proper center is a center which 
does not lie on the surface, an improper center of a surface lies on the 
surface. 

It follows from this definition that, if A (a, 0, 7) is a center of the 
quadric surface Q 9 then the two roots of the equation 

L 2 = 



which was established in Section 80, must be equal numerically 
but opposite in sign for all admissible values of X, pt, and v (see 
the Remark, following Theorem 7, Chapter III, Section 33, page 
56) ; hence the sum of these roots must equal 0. 

If Lo =1= 0, that is, if the line through A (a, 0, 7) does not have an 
asymptotic direction (see Definition I, Section 80, page 160), the 

O J 

sum of the roots is equal to p- 1 ; it will be equal to zero therefore 

JL/o 

* A chord of a surface is a line which joins two of its points; it follows from 
Corollary 1 of Theorem 1, Section 80, page 160, that a chord of a quadric surface 
does not have any other points in common with the surface besides the two 
points which it joins, unless it lie entirely on the surface. 



THE CENTERS AND VERTICES OF QUADRIC SURFACES 177 

if and only if LI = 0. And if L = 0, so that one of the roots is 
infinite, the condition requires that the other root be also infinite, 
which leads again to the condition LI = 0. Conversely, if LI = 0, 
the two roots of equation (1) are equal numerically but opposite 
in sign. Therefore the necessary and sufficient condition that 
A(dj 0, 7) be a center is that L\ for all admissible values of 
X, ju, v. In particular we must have LI s= XQi(a, , 7) +AtQ2(, , 7) 
+ vQz(a, |8, 7) = for the sets of values 1,0, 0; 0, 1, and 0, 0, 1 
of X, Hj v] these special sets lead to the conditions Qi(a, 0, 7) = 0, 
Q 2 (a, 18, 7) = 0, Q 3 (a, )8, 7) = 0. Moreover it is easily seen that 
if these conditions are fulfilled, then LI will vanish for every ad- 
missible set of values of X, /*, v. We have therefore obtained the 
following theorem : 

THEOREM 13. The necessary and sufficient conditions that a point 
shall be a center of the quadric surface Q Is that its coordinates satisfy 
the three linear equations Qi(x 9 y 9 a) = 0, Qt(x 9 y 9 a) = 0, Qa(x 9 y 9 a) = 0. 
If and only if the coordinates satisfy moreover the condition Q(x 9 j, 3} 
=t= 0, the point is a proper center. 

COROLLARY. An Improper center of a quadric surface Is a vertex of 
the surface, and conversely. 

For the further discussion of our problem we observe in the first 
place that, in view of Corollary 1 of Theorem 4, the condition 
Q = of Theorem 13 may be replaced by the condition Q 4 4= 0. 
Consequently a proper center is a point common to the three 
planes 

~ = anx + a w y + aisz + M = 0, ~ 



+ 24 = 0, y = a 13 X + 0237/ + O& + 34 = 0, 

but not on the plane 

~ = a u x + any + 0342 + 044 = 0; 

and a vertex is a point common to the four planes. 

We shall denote the coefficient matrix of the first three equations 
by as and its rank by r 3 ; the value of the determinant of a 3 has 
already been designated by An (see equation (3), Section 84, page 
173). The augmented matrix of the first three equations is 



178 QUADRIC SURFACES, GENERAL PROPERTIES 



lijll i = 1, 2, 3; j = 1, 2, 3, 4; we shall denote it by b. The 
coefficient matrix of the equations of the set of four planes is 
\\dij\\, i = 1, 2, 3, 4; j = 1, 2, 3; we shall denote it by b'; and the 
augmented matrix of this set of four equations is the discriminant 
matrix of the quadric, which we have already designated by a 4 
(see page 173, proof of Corollary of Theorem 11) and whose de- 
terminant is denoted by A. 

We fall back now on Theorems 20, 22, 23, 24 of Chapter IV and 
their Corollaries (Sections 51 and 54, pages 95, 101, and 102); ap- 
plication of these theorems shows that if there is to be any center 
the matrices a 3 and b must have the same "rank, and if there are 
to be any vertices, the matrices b 7 and a 4 must have the same rank. 

Now it should be clear: (1) that the ranks of the matrices b 
and b' are equal, since either of these matrices is obtained from 
the other if we write the columns as rows and vice versa; (2) that 
the rank of a 3 can not exceed that of b, which in turn can not ex- 
ceed the rank of a 4 ; (3) that the rank of a 4 can not exceed the 
rank of b by more than 1 ; and (4) that the rank of b can not ex- 
ceed the rank of a 3 by more than 1. Moreover, (5) if the ranks of 
b and a 4 are equal, then the ranks of b and a 3 are equal. 

An algebraic proof of this last statement may be somewhat 
lengthy. It can be deduced very readily however from the theorems 
of Chapter IV referred to above. For if the ranks of b and a 4 
are equal, the four planes have at least one point in common; 
consequently the first three planes have at least one point in 
common and therefore the c.m. and the a.m. of the first three 
equations have the same rank, i.e., the ranks of a 3 and b are equal. 

We conclude from (3) and (4) that r 4 and r 3 can differ by 2 at 
most. If r 4 r 3 = 2, then the rank of b is different from either. 
If r 4 r 3 = 1, it follows from (5) that the rank of b is equal to 
r 3 . And, if n = r s , the rank of b is of course equal to the same 
number. It should be clear that the existence of proper centers 
and vertices depends on the difference r 4 r 3 . If we draw on the 
further content of the Theorems of Chapter IV, which were cited 
above, we obtain the following result : 

THEOREM 14. If the ranks of the matrices a 4 and as are equal the 
quadric surface has a unique vertex, a line of vertices or a plane of 
vertices, according as this common rank Is 3, 2 or 1. If the ranks of 
these matrices differ by 1, the quadric surface will have a single 



THE CENTERS AND VERTICES OF QUADRIC SURFACES 179 



proper center, a line of proper centers or a plane of proper centers, ac- 
cording as the lower of these ranks is 3, 2 or 1. If the ranks of these 
matrices differ by 2, the quadric surface has no center at all. 

Remark 1. The content of this theorem may conveniently be 
put in the following tabular form : 



7-3 


r 4 


The quadric surface has 


3 


4 


a single proper center 


3 


3 


a single vertex 


2 


3 


a line of proper centers 


2 


2 


a line of vertices 


1 


2 


a plane of proper centers 


1 


1 


a plane of vertices 


n - r 3 > 1 


no center 



Remark 2. The reader should convince himself that the cases 
indicated in this table include all possible cases for the ranks of 
the matrices a 3 and a 4 and that therefore the conditions of Theorem 
14 are sufficient as well as necessary. 

Remark 3. A quadric surface with a single proper center is 
called a central quadric. A quadric surface with a single vertex 
is called a proper quadric cone. 

Remark 4. A non-singular quadric surface is either a central 
quadric or else a surface without any center. 

We record moreover the following corollaries. 

COROLLARY 1. The rank of the matrix a 4 can not exceed the rank of 
the matrix a 3 by more than 2. 

COROLLARY 2. The necessary and sufficient condition that a quadric 
surface be a conical surface is that the ranks of the matrices a 4 and a 3 
be equal. 

We are able furthermore to complete in an essential way the 
result contained in Corollary 2 of Theorem 12 (Section 84, page 
175), as follows: If a quadric surface has a plane of vertices, it 
consists of this plane, counted doubly. For, if it contained a point 
A outside this plane it would have to contain every line which 
connects a point of the plane with A (compare the last sentence 
in Theorem 12, Section 84, page 175). This is obviously impos- 
sible; therefore the surface can not contain any point outside the 
plane of vertices. And it should be a simple matter to show that 



180 



QUADRIC SURFACES, GENERAL PROPERTIES 



this plane muse be counted doubly. And if a quadric surface has 
a line of vertices, it must consist of two planes through this line. 
For, if A is any point of the surface outside the line I on which the 
vertices lie, then the plane determined by I and A must be entirely 
contained in the surface; and the argument used in the proof of 
Corollary 2 of Theorem 12 (page 175) shows that then the surface 
consists of two planes. If these two planes were coincident planes 
the equation of the surface would be Q = (ax + 677 + cz + d) 2 
= 0, from which we could conclude that r 3 = r 4 = 1, and there- 
fore the surface would have a plane of vertices. We can therefore 
state the following result : 

COROLLARY 3. A singular quadric surface Is a proper quadric cone 
if and only If r a = r 4 = 3, a pair of Intersecting planes If and only If 
rj = r4 = 2, a pair of coincident planes If and only If n = r< = 1. 

After the existence of centers or vertices has been established, 
their position can be determined by solving the equations Qi = 0, 
$2 = 0, Q 3 = 0. In the case of a central quadric, these equations 
have an unique solution which is given by Cramer's rule (see 
Theorem 1, Chapter II, Section 21, page 37). The solution may 
be written in the following form: 



012 013 



x :y : z : 1 = - 

034 

011 012 014 
12 022 024 
013 023 034 



011 014 013 



012 024 023 

013 034 033 

011 012 013 

012 022 023 

013 023 033 



It should be easy to see that the terms on the right are equal to 
the cofactors of the elements in the last row of the discriminant A. 
If these cofactors are designated in the usual manner, we have 

x : y : z : 1 = An : A 2 * : AM : An. 

COROLLARY 4. The coordinates of the center of a central quadric 
surface are equal to ^ 4 , ^ 4 , ^- 

A\\ Au A\i 

Examples. 

1. The coordinates of the possible centers of the surface 5 x- 4- 5 y 2 -f 8 z 2 
Szx 2xy -{- 12x 12 y -f 6 =0 must satisfy the equations 



THE CENTERS AND VERTICES OF QUADRIC SURFACES 181 



|i = 5s-2/ + 4z+6=0, 



6=0, 



Q* 



4 
Moreover, ~ = 6 x < 



The rank of the matrix a 3 = 



5 -1 4 
-1 54 

4 48 

3rd row is equal to the sum of the first two rows, whereas the two-rowed minor 
in the upper left-hand corner does not vanish. And the rank of the matrix 
5-14 6 
4 -6 
8 
6 

sum of the 1st and 2nd rows, whereas the matrix contains several non-vanishing 
three-rowed minors. We conclude therefore that the surface has a line of 



a 4 = 



-1 5 
4 4 
6 -6 



is readily found to be 2; for the 



is found to be 3; for the 3rd row is equal to the 



Q 



centers in the line of intersection of the two planes ~ 5x y + 4z 

and ~cT Z-J-5T/ + 42 6 = 0. The parametric equations of this line 

are found to be x = 3 + tf, y 5 + t, z = 4 t. 

2. To determine the possible centers of the surface 2 x 2 3 y 2 -f 4 ^z 
5 zx + 4 x 3 ?/ + 5 = 0, we set up the equations Qi =4a; 52-f-4=0, 
#2 = -6 y + 4 2 - 3 = 0, Q 3 = -5 x + 4 1/ = 0, and Q 4 = 4 x - 3 1/ -f- 10 
= 0. The determinant of the matrix as has the value: 



4 -5 

-6 4 

-5 4 



43 



and the discriminant 



A== I6 X 



40-54 

0-64 -3 

-5400 

4 -3 10 



861 



Therefore r s = 3 and n = 4; consequently the surface has a single proper 
center; its co6rdinates are A, *\, Jjj. 

3. For the surface 2 x 2 + 20 1/ 2 4- 18 z 2 - 12 yz + 12 xy + 22 x + 6 y - 
2 2 2 = 0, the matrices as and a 4 are 



260 
6 20 -6 
-6 18 



and 



respectively. We find that r s = 2 and r 4 
has neither a proper center nor a vertex. 



2 6 11 

6 20 -6 3 

-6 18 -1 

11 3 -1 -2 

4; we conclude that the surface 



182 QUADRIC SURFACES, GENERAL PROPERTIES 

86. Exercises. 

1. Show that through every point of the surface 4 x 2 6 y 2 12 z there 
pass two real distinct lines which lie entirely on the surface. 

x 2 y' 2 z 2 

2. Prove that there are no real lines on the ellipsoid -- -f 75 + - = 1- 

CL C 



3. 



Show that the ellipsoid ~ + ~ + ~ - 1, the hyperboloid of one sheet 

x 2 y 2 z 2 x 2 y 2 z 2 

I ~h r 2 2 = 1 an( l ^ c hyperboloid of two sheets ^ ^ 1 are 

central quadric surfaces. 

4. Show that through every point of the hyperboloid of one sheet of Exer- 
cise 3 there pass two real lines which lie on the surface; and that no such lines 
exist through any point of the hyperboloid of two sheets of Exercise 3. 

6. Show that the locus of the equation : /- -f = is a proper quad- 

a 2 b 2 c 2 

ric cone; and prove that every tangent plane of this surface passes through 
the vertex. 

6. Determine the conditions which the direction cosines of a line through 
the point A( 1, 1, 1) on the surface x 3 ?/ 3 -h 2 3 = 1 must satisfy in 
order that the line may lie entirely on the surface. 

7. Determine the centers, proper centers or vertices, of each of the follow- 
ing surfaces: 

(a) x 2 + 5 y 2 - 2 z 2 + 6 yz + 8 xy - 4 x + 6 y - 6 z -f 6 = 
(6) 9 x 2 -f 49 y 2 -f 4 z 2 - 28 yz + 12 zx - 42 xy - 24 x -f 56 y - 16 z 
+ 16-0 

(c) 3 x 2 -f 5 y 2 -f- 9 z 2 + 2 yz -f 8 zx - 4 xy - 6 x + 4 y - 4 z + 3 = 

(d) 5x 2 - y 2 - 16 z 2 - 20 yz + 4 ar - 8 xy - 6 * + 2 y - 8 z + 2 = 
(c) 4 x 2 -f y 2 + 9 z 2 - 6 yz + 12 zz - 4 xy + 6 x - 3 y -f 9 2 - 4 = 
(/) G x 2 - 2 y 2 - 2 z 2 + 5 ys - zz - 4 jy - 10 x - 6 y -f 9 z - 4 = 
to) 3 :c 2 -f 3 y 2 + 3 z 2 - 2 yz - 2 zz - 2 x?y + 8 x - 4 2 + 6 = 

(/O 2 x 2 -f- 5 ?/ + 2 z 2 - 6 2/2 + 4 ac - 6 xy -f 2 3 - 4 y + 2 z -f 2 = 0. 

8. Show that the tangent lines from a point A (a, p, 7) to the elliptic cylinder 

x 2 y 2 

+ p = 1 lie on a pair of planes through a line parallel to the Z-axis. 

9. Prove that if a quadric surface has a plane of centers the surface consists 
of a pair of parallel planes. 



10. Prove that the elliptic paraboloid ^-}-~^ = < 2pz and the hyperbolic 

x z ? ,a a 

paraboloid -5 ~-^ 2 pz do not possess a center. 

87. The Asymptotic Cone. If C(a, 0, 7) is the center of the 
central quadric Q, then Qi(a, ft 7) = 0, Q 2 (a, 0, 7) = 0, Q 3 (, ft 7) 
= and Q(a, 0, 7) =t= 0. It follows that the equation of the 



THE ASYMPTOTIC CONE 183 

tangent cone from C to the surface reduces from the form given at 
the end of Section 82 (page 169) to the simpler form: 

(1) a n (x - aY + a*(y - ft 2 + (te(z - 7) 2 + 2 a n (y - ft) 
(z - 7) + 2a 3 i(z -y)(x- a) + 2a 12 (rc - a) (y - 0) = 0. 

Since the direction cosines of a generating line on this cone are 
proportional to x a, y p, and z y, where x, y, z are the 
coordinates of some point on the cone, it follows that the direction 
cosines X, ju, v of such a line satisfy the equation anX 2 + a^v? 
+ a^v 2 + 2 023M" + 2 anv\ + 2 a l2 \^ = 0, that is, the equation 
q(\ fjLj v) = 0. Since moreover they evidently satisfy the equation 
L! = \Qi(a, P, 7) + M& + J>#3 = 0, the generators of this cone are 
asymptotes of the surface (compare Corollary 3 of Theorem 1, 
Section 80, page 160). 

DEFINITION IX. A cone of which every generator is an asymptote of 
a surface is called an asymptotic cone of the surface. 

We can therefore say that the center of a central quadric sur- 
face is the vertex of an asymptotic cone of the surface. The same 
argument shows that a proper center of any non-degenerate quad- 
ric surface is the vertex of an asymptotic cone. And we raise the 
question whether any other points, besides proper centers, can be 
vertices of such cones of non-degenerate quadrics. If A (a, ft 7) 
is such a point, we know from Corollary 3 of Theorem 1 (Section 
80, page 160) that the equations q(\, /i, v) = and XQi(X, /i, v) 
+ t*Qz + vQs = must have an infinite number of solutions which 
are admissible values of X, /i, and p. Let us suppose now: 

(a) that A (a, ft 7) is not a center of the surface. We can then 
suppose that Qi(a, ft 7) 4= and proceed as in Section 84. The 
quadratic equation (1) which was discussed in that section will 
have more than two roots if and only if the rank r 4 of the dis- 
criminant matrix is less than 3 (compare (b) on page 173), that is, 
if the quadric surface is degenerate (see Corollary 2 of Theorem 
12, Section 84, page 175). Hence for a non-degenerate non- 
singular quadric a center is the only point which can be the vertex 
of an asymptotic cone. And we suppose: 

(b) that A (a, ft 7) is a vertex of the surface. In this case 
Q, Qi, Qz, and Q 3 all vanish for x = a, y = ft z = y. If we make 
use once more of Taylor's theorem as in Case II on page 174, we 



184 QUADRIC SURFACES, GENERAL PROPERTIES 
find that 

Q(*> y, *) = Q(* + [*-], + [y - ffl, 7 + b - 7]) 

= Q(*, ft 7) + (x - )&(, ft 7) + fa - j8)Q 2 (a, ft 7) 
+ (2 - 7)03(, ft 7) + q(x - , 2/ - ft ^ - 7), 

so that the equation of the surface reduces to the equation q(x a, 
y ft z 7) =0, which is the equation of the asymptotic cone. 

If we observe furthermore that it follows from the Taylor's ex- 
pansion written above that the equation of the asymptotic cone of 
a central quadric can also be written in the form Q(x, y, z) 
Q(ct, ft 7) = 0, we can put our results in the form of the follow- 
ing theorem. 

THEOREM 15. A non-degenerate quadric surface Q has an asymp- 
totic cone if and only if It has a center. If It has a proper center at 
A(a 9 9 y) the equation of the asymptotic cone is Q(x 9 y, z) Q(a 9 /3, 7) 
= 0; if it has a vertex the asymptotic cone is identical with the surface 
itself. 

Obviously there is no further interest in considering the asymp- 
totic cone of a surface which has vertices; therefore there remain 
for consideration the non-degenerate quadrics which have proper 
centers, that is, the cases in which r 4 and r 3 are 4 and 3, or 3 and 2 
respectively. 

CASE 1. 7*4 = 4, r 3 = 3. In this case there is a single center 
and therefore a single asymptotic cone, which is a proper quadric 
cone. 

CASE 2. r 4 = 3, r 3 = 2. In this case there is a line of cen- 
ters determined by the equations Qi(x, y, z) = 0, Q*(x 9 y, z) = 
and Q 3 (#, y> z) = 0. The c.m. of these equations is as. We 
shall henceforth denote the cof actors of the elements a#, i,j = 1, 2, 
3 of this matrix by a#, i, j = 1, 2, 3. Since r 3 = 2, not all of these 
cofactors vanish. Let us suppose a 33 4= 0; then the direction 
cosines of the line of centers are proportional to ai 3 , 2 3, and a 3 s 
(see Theorem 17, Chapter IV, Section 47, page 87). Therefore, 
if a, ft 7 is an arbitrary point on the line of centers, the parametric 
equations of this line may be put in the form x = a + ant, 
y = ft + awl, 2 = 7 + asrf; and the equation of the asymptotic 
cone which corresponds to an arbitrary center can be put in the 
form: 

Q(*> y, *) - Q(<* + irf, ft + "23*, 7 + aasO = 0. 



matrix of this equation is 



since in this case 



DIAMETRAL PLANES OF A QUADRIC SURFACE 185 

If the second term on the left-hand side is expanded by Taylor's 
theorem, the equation reduces to 

Q(x, V, *) ~ Q(, ft, T) ~ <[awQi(, ft 7) + 23 Q 2 (a, p, 7) 
+ 0:3363(0;, ft 7)] - t 2 q(au, a 23 , ass) = 0. 

The coefficient of t is obviously zero; and it is shown in the Ap- 
pendix* that the coefficient of t 2 also vanishes. Consequently the 
equation of the asymptotic cone, which corresponds to an arbi- 
trary center, is independent of t] that is, there is only one asymp- 
totic cone. If its equation is written in the form q (x a, y ft 
* ~~ T) = 0, and we translate the axes to the point (a, ft 7) as 
origin (see Theorem 2, Chapter V, Section 61, page 115), the 
equation takes the form q(x', y', 2') = 0. The discriminant 

i 2 ai3 

022 #23 

013 023 33 

0000 
r 3 = 2, the rank of this discriminant matrix is also 2, and there- 
fore, the asymptotic cone consists of a pair of intersecting planes 
(compare Corollary 3 of Theorem 14, Section 85, page 180) ; that 
is, the asymptotic cone degenerates into a pair of asymptotic 
planes. We have now obtained the following amplification of the 
last theorem. 

THEOREM 16. If the ranks r 4 and r 3 are equal to 4 and 3 respectively, 
the quadrlc surface Q has a single proper quadrlc cone as asymptotic 
cone; this cone may be real or Imaginary. If these ranks are equal to 
3 and 2 respectively, the surface has a pair of asymptotic planes, which 
may be real or imaginary. In either case the equation of the asymp- 
totic cone may be written In the form Q(x 9 y, *) - <?(, /?, 7) = 0, where 
, /3, 7 are the coordinates of a center of the surface. 

88. The Diametral Planes and the Principal Planes of a Quadric 
Surface. We return once more to the equation in Theorem 1 
(Section 80, page 160) and inquire for the locus of points which 
are midpoints of chords drawn through them in a fixed direction 
given by the ratios X : n : v. The argument which led us to 
Theorem 13 in Section 85 shows that if A (a, 0, 7) is a point of this 
locus, then XQi(a, j8, 7) + vQz + vQ 3 must vanish for the specified 
values of X, JJL, and v. Since d, Q, and Qs are linear functions of 

* See III, page 297. 



186 QUADHIC SURFACES, GENERAL PROPERTIES 

a, P, y, we conclude that the locus is a plane. We obtain therefore 
immediately the following theorem. 

THEOREM 17. The locus of all points which bisect chords of the 
quadric surface Q whose direction cosines are X, /*, v 9 is the plane 
x(M* y ) + M<M* r> *) + "(M* j, *) = o. 

DEFINITION X. The plane which is the locus of the midpoints of a 
set of parallel chords is called the diametral plane of the direction of 
these chords. 

If we write out in full the expressions for Qi, Q 2 , and (? 3 in the 
equation of the diametral plane and collect the terms in x, y y and 
z, the equation of this plane takes the form : 



= 0, 

or, using the notation introduced in Section 80, 

<7i(X, fjL 9 v)x + </ 2 (X, n, v)y + </ 3 (X, M, v)z + </ 4 (X, /x, ^) = 0. 

For all values of X, ju, and *>, this equation represents a plane 
(for those values for which #i(X, /*, i/) = g 2 (X, /i, v) = g 3 (X, /i, ^) = 0, 
and g 4 (X, ju> v) 4= 0, this plane is the "plane at infinity/' see 
Remark 2, Section 41, page 73), except for such values as cause 
% #2, Qsj and #4 to vanish simultaneously; but this can not happen 
for admissible values of X, ju> and v unless the rank of the matrix 
b is less than 3 (see Theorem 2, Chapter II, Section 22, page 38; 
compare also Section 85, page 178). We can therefore state the 
following corollary. 

COROLLARY. In a quadric surface for which the rank of the matrix 
b is equal to 3, there exists a diametral plane corresponding to every 
direction; In a quadric surface for which the rank of this matrix is 
less than 3, this correspondence falls for those directions for which 

|i(X f M> v) = q* = q 3 = q* = 0. 

The correspondence between systems of parallel chords and di- 
ametral planes which has been established for quadric surfaces, is 
an extension to three-space of the correspondence between con- 
jugate diameters in the theory of conic sections; for either of two 
conjugate diameters is the locus of the midpoints of chords parallel 
to the other. We recall that in the ellipse and the hyperbola 
there is one pair of mutually perpendicular conjugate diameters, 
namely, the axes of these curves. On account of the importance 



PRINCIPAL PLANES OF A QUADRIC SURFACE 187 

of these lines in the theory of these curves, we are led to inquire 
whether there are directions in a quadric surface which are per- 
pendicular to the corresponding diametral planes. To facilitate 
the discussion, we introduce the following definition. 

DEFINITION XL A principal direction of a quadric surface is a di- 
rection which is perpendicular to the corresponding diametral plane; 
a diametral plane which corresponds to a principal direction is called 
a principal plane. * 

According to Corollary 3 of Theorem 7, Chapter IV (Section 
44, page 79), the angle between any direction X, /*, v and the cor- 
responding diametral plane, when this is a "plane at finite dis- 
tance," is given by the formula: 



- 

v qi* + g 2 2 + </s 2 

Therefore the necessary and sufficient condition that the diametral 
plane which corresponds to the direction X, M, v shall be a plane at 




FIG. 32 

finite distance and perpendicular to this direction, is that the 
equation 

Xffi(X, At, v) /igsCX, fJL, v) vq*(\ M, v) _ , 

2k "*" 2k ~*~ 2k 

shall be satisfied by admissible values of X, M> and v, such that 
fc = ^ 2 + g2 2 + ^ 2 ^ Q (gee Fig 32) ^ If wc multiply thia 



* If fc = 0, we must have #1 = q* #3 = 0, so that wc would be dealing 
with the plane at infinity if there were a plane at all; and if the diametral 
plane were the plane at infinity, we would have q\ = 5-2 = q* and therefore 
k = 0. Consequently the non-vanishing of fc is a necessary and sufficient 
condition that the diametral plane shall be a plane at finite distance. 



188 



QUADR1C SURFACES, GENERAL PROPERTIES 



equation by 2 and subtract the result from the sum of the equations 

x 2 + ^ + ^ = i and rr 2 + rn + r& =1 ' we obtain the 

4 fc 2 4 fc 2 4k 2 
condition 



which in turn is equivalent to the three equations 

#i(X, /x, v) = 2 fcX, q 2 (\, M, ") = 2 fc/i, g 3 (X, M, v) = 2 fci/; 
that is, to the equations 

(1) (on ~ k)\ + a 12 M + auv = 0, Oi 2 X + (o 22 fc)/* + 
= 0, Oi 3 X + o 23 /* + (a 33 fc> = 0*. 



The condition for the existence of a principal plane, stated 
above, is therefore equivalent with the condition that there exist 
admissible values of X, /x, v which satisfy the equations (1) and for 
which k 4= 0. But, since these equations may be looked upon as 
linear homogeneous equations in X, /z, and v y it follows from 
Theorem 2, Chapter II (Section 22, page 38) that their coefficient 
determinant must vanish (since the trivial solution of these equa- 
tions does not lead to admissible values of X, /x, and *>); that is, 
we must have: 

^12 O 22 fc O23 

2l3 23 033 

This equation is a cubic in fc; therefore it has 3 roots. To every 
root fc*, for which the rank of the corresponding matrix 

On fc* Oi 2 



(2) 



=0. 



(3) 



Ol 2 



22 ~ fc* 



3 3 fc 



is 2, there corresponds (compare Corollary of Theorem 4, Chapter 
II, Section 25, page 42) a single infinitude of values of X, /z, v deter- 
mining uniquely the ratios X : n : v, and hence a single principal di- 
rection. If the rank of the matrix (3) is 1, the three equations (1) 

* It should be clear that these equations can be derived, independently of 
the formula for sin 0, from the equation of the diametral plane by means of 
Theorem 7, Chapter IV, Section 44, page 78. The derivation followed in the 
text has the advantage of giving significance to the variable k. 



PRINCIPAL PLANES OF A QUADKIC SURFACE 189 

are equivalent; there is therefore only one linear condition on X, M, 
and v t so that an arbitrary admissible value may be assigned to one 
of these variables. Hence, to a value k* of k, for which the rank of 
the matrix (3) is 1, there corresponds a single infinitude of principal 
directions. If there is a k* which causes the rank of the matrix 
(3) to become 0, then the direction cosines X, ju, v are entirely un- 
restricted,t and hence every direction is a principal direction. In 
order to facilitate the statement of our results we introduce the 
following definition. 

DEFINITION XII. The equation (2) is called the discriminating equa- 
tion of the quadric surface Q, the discriminating numbers of the 
surface Q are the roots of the discriminating equation. 

THEOREM 18. Every quadric surface has three discriminating num- 
bers; to each of these corresponds a single principal direction, a 
single infinitude of principal directions, or all directions, according 
as it gives the matrix (3) the rank 2, 1 or 0; with every discriminating 
number which is different from zero, there Is associated a principal 
plane at finite distance. 

Remark. The direction cosines of a principal direction, associ- 
ated with the discriminating number ki will be denoted by X,-, ju,-, 
and vi, i = 1, 2, 3; they are found by solving any two of the 
equations (1) for the ratios X,- : //, : v{. 

Since q(x, y, z) is a homogeneous function of the second degree 
in x, y, z, we find by application of Euler's theorem (see footnote 
on page 161) that 

2ff(Xi, MI, Vi) = A#i(X,-, /i,-, Vf) + Maftw &, vj) + ^(A;, ^, v$ 

= 2 fc,'(V + tf + V f) = 2 ki. 

This leads us to the following useful corollary. 

COROLLARY. If X;, & , \>i are the direction cosines of a principal direc- 
tion t corresponding to the discriminating number ki of the quadric 

f That such a situation may arise should be clear geometrically from the fact 
that in a sphere every plane through the center bisects the chords which are 
perpendicular to it, so that every direction is a principal direction. In an 
ellipsoid of revolution, every plane through the axis of revolution bisects the 
chords perpendicular to it, so that every direction perpendicular to this axis 
is a principal direction; this furnishes an example of a surface in which the 
direction cosines of a principal direction are subject to only one linear condi- 
tion. 

J It is to be understood here that more than one principal direction may bo 
associated with a single discriminating number. 



190 



QUADRIC SURFACES, GENERAL PROPERTIES 



surface Q 9 then 



Mi> vi) 
2, 3. 



2 fa I'M 



89. The Discriminating Equation. We proceed now to a fur- 
ther discussion of the discriminating equation 

11 k 012 



012 022 



023 



013 

023 

#33 fc 



= of a quadric surfacef 



THEOREM 19. A root fc* of the discriminating equation is a single, 
double, or triple root according as the rank of the matrix 

fill ~ /C* Oi di2 



(2) 



is 2, 1 or 0; and conversely. 



k* 



k* 



Proof. It follows from Theorem 19, Chapter I (Section 17, 
page 32) that, if the left-hand side of equation (1) is designated by 
A(k) and its derivatives with respect to k by means of accents, 
then 



-1 


ft 11 k (l\2 $13 




A' (k) = i2 022 k 


023+0-1 




013 023 033 ~~ k 013 023 033 ~~ 


- k 




0n k (i 12 


013 


f 1 7 










022 A,* 2 3 




-)- 


012 022 K> 


023 


= - n ] 









-1 


[1 023 033 k 






0n k 013 




a n ^ 012 |1 




~*~ 


013 033 k 




012 022 A' |p 





A"(k) = 2[( n - k) + (, 2 - A-) + (033 - fr)], A'" (k) = -G. 

f It should be clear how equations analagous to the one written above can 
!>e formed for every symmetric square matrix ||i;||, i, j = 1, 2, . . . , n of any 
order. Such an equation is usually called the characteristic equation of the 
matrix. The equation treated in the text is therefore the characteristic 
equation of the matrix a 3 . The characteristic equation of a matrix plays a 
very important role in the theory of matrices. Many of the properties devel- 
oped in the text for the characteristic equation of the matrix a 3 hold, with 
appropriate changes, for the characteristic equation of the general symmetric 
square matrix; and the methods of proof here used are readily adaptable to 
the general case. 



THE DISCRIMINATING EQUATION 191 

Suppose now 

(a) the rank of the matrix (2) is 0. Then clearly, A(k*) = 0, 
A 9 (fc*)'= 0, and A 11 (fc*) = 0; therefore k* is a triple root of the 
equation A (k) 0. 

(b) the rank of the matrix (2) is 1. In this case A(k*) = 0, 
A'(k*) = 0, but A"(k*) =f= 0. For, the rank of the matrix being 
1, every two rowed-principal minor vanishes; that is, (a# k*) 
(a,jj k*) = a# 2 ^ 0, i, j = 1, 2, 3; f 4 1 j. It follows from this 
that if two of the elements in the principal diagonal of (2) vanish, 
then all the elements outside the principal diagonal vanish also; 
hence, the remaining element in the principal diagonal must be 
different from zero and therefore A"(k*) 4 1 0. And it follows also 
from this relation that any two principal diagonal elements of (2) 
which do not vanish must be of the same sign, so that A"(k*) 
can vanish only, if each of its terms vanishes, which has been shown 
to contradict the hypothesis that the rank of the matrix (2) is 1. 
Consequently fc* is a double root of the equation (1). 

(c) the rank of the matrix (2) is 2. In this case we can apply 
the Corollary of Theorem 7, Chapter II (Section 26, page 44); 
and we conclude that A' (A;*) ^ 0. It follows therefore that k* is 
a simple root of equation (1). 

The converse follows from the fact that the three cases in the 
hypothesis and in the conclusion both represent all possibilities. 
For example, if k* is a double root of equation (1), the rank of the 
matrix (2) is 1 ; for if the rank were not 1 , it would be 2 or 0, and 
7c* would therefore be a simple root or else a triple root of the 
equation. 

If we expand the polynomial A (k) according to Taylor's theorem, 

we find that 

fr 2 fr 3 

A(k) = A(0) + k A'(0) + ~ - A" (0) + ~ - A'" (0). 

From the formulae developed in the proof of the theorem, we 
find that 

A (0) = |a y |, i, j = 1, 2, 3; that is, A (0) = A H (compare 
page 173), 

A f (o) = ~ n ^ a23 + an ai3 + an ai2 1 1 

[J #23 #33 a !3 ^33 12 &22 | J ' 

= [n + <*22 + 0:33], (compare page 184) 
A" (0) = 2(a n + a 22 + a 33 ), A"' (0) = -6. 



192 



QUADRIC SURFACES, GENERAL PROPERTIES 



We shall now use the following abbreviations: 

Ti = an + 022 + #33, T 2 = 11 + 22 + 33- 

Our discussion yields then the following corollary. 

COROLLARY 1. The expanded form of the discriminating equation 
of the quadric surface Q is 

(3) fc 3 - 7W 4- T*k - An = 0. 

For convenience of future reference we record here also the 
following results obtained by applying Theorem 18, Chapter I 
(see Section 16, page 29; see also the Corollary of Theorem 5, 
Chapter II, Section 26, page 43) and Theorem 7, Chapter II (see 
Section 26, page 44) to the determinant A^. 

COROLLARY 2. Between the value of the determinant An, the ele- 
ments In its principal diagonal and the elements aij of its adjoint, the 
following relations hold: 



COROLLARY 3. If the determinant A vanishes, then those of its 
principal two-rowed minors which do not vanish are of like signs. 

THEOREM 20. The discriminating numbers of a real quadric surface 
are real. 

Proof. We will show first that, if the coefficients a# in the equa- 
tion of the surface are real, then the discriminating equation can 
not have a root which is a pure imaginary. Suppose that iq is a 
root of equation (1). Then, according to a well-known theorem 
of algebra, iq must also be a root of this equation; that is, we 
will have 

a u +iq 
= 0, and 



012 



#23 



012 



012 

022+ iq 
023 



023 



= 0. 



012 022 1 

013 023 033 iq 

But the product of the two determinants on the left-hand sides of 
these equations must then also vanish; application of Theorem 16, 
Chapter I (Section 14, page 26) gives us therefore the further 
result 



^01*03* 



= 0. 



THE DISCRIMINATING EQUATION 193 

This last equation is of the same general form as equation (1) and 



is obtained from it if we substitute q 2 for A; and 



for ay. 



Corollary 1 of Theorem 19, enables us to state therefore that the 
expanded form of this equation is 

where 



and 183 = 



It should be clear that S s = A^ y and therefore that S\ and S 3 are 
both non-negative. To show that the same thing is true of S 2 , we 
observe that to each determinant in S 2 we can apply the Lemma 
preceding Theorem 14, Chapter III (Section 36, page 64) ; thus 2 
is transformed in to the sum of the squares of 9 two-rowed de- 
terminants. Consequently equation (4) has no negative coeffi- 
cients and therefore, considered as a cubic in g 2 , no solutions for 
which q 2 is positive. Therefore equation (2), the discriminating 
equation of the quadric surface, can have no root of the form iq, 
where q is real, unless q = 0. 

Suppose next that equation (1) has a root of the form p + iq; 
then the new equation obtained from (1) by replacing an, 022, 033 



194 QUADRIC SURFACES, GENERAL PROPERTIES 

by an p, a 2 2 P, 3'i P has the root iq. But the new equation 
is of the same form as equation (1) and therefore it can not have a 
root iq. Our theorem has therefore been proved. 

THEOREM 21. Not all the discriminating numbers of a quadrlc sur- 
face can lie equal to zero. 

Proof. If all the roots of equation (3) were zero, then we would 
have A 44 = 0, T<> = 0, and 7\ = 0. But from the last two of these 
equations we could then conclude that 

TV - 2 T 2 = (a u 2 + 22 2 + a 33 2 + 2 a 22 a 33 + 2 a 33 an + 2 a u a 22 ) 
2 (a 2 2fl33 23 2 ) 2(a 3 3n a J3 2 ) 2(a u a 2 2 i2 2 ) 
= an 2 + 22 2 + 33 2 + 2 aas 2 + 2 a 13 2 + 2 a 12 2 = 

and from this it would follow that an = ^22 = ^33 = 23 = is = 012 
= 0, which would mean that the equation of the surface contains 
no terms of the second degree. 

COROLLARY. The functions T and T 2 can not both vanish; If Ti = 0, 
then T 2 < 0. 

The first part of this corollary follows also from Theorem 20 
(page 192), for, if TI = T 2 = 0, the discriminating equation re- 
duces to fc 3 = A 44, which has only one real root. 

90. Principal Planes and Principal Directions. The results of 
the two preceding sections enable us to establish some further 
properties of principal planes and principal directions. 

THEOREM 22. Every quadrlc surface has at least one real principal 
plane at finite distance. 

This theorem is an immediate consequence of Theorems 18 (last 
part) and 21. 

THEOREM 23. The principal directions which correspond to two 
distinct discriminating numbers are mutually perpendicular. 

Proof. Let ki and k% be two discriminating numbers of the 
quadric surface and let ki = Afe. Then, in the notation of the 
Remark following Theorem 18 (see page 189) and in virtue of the 
Corollary of this theorem, we have 



Mi, Vi = iMi, (foi, Mi, vi = 

M2, V 2 ) = fc 2 X 2 , ^2(X 2 , M2, 1*2) = ^2M2, ^(Xo, M2, ^2) = 



PRINCIPAL PLANES AND PRINCIPAL DIRECTIONS 195 

Now the reader should have no difficulty in showing, by writing 
out the expressions in full, that 



Ml, V\) ^2?si, Ml, ^l = l<?l'2, M2 



Therefore, if we multiply the equations of the first set written 
above by X 2 , M2, and v 2 respectively and those of the second set by 
Xi, MI, vi respectively, we find that 

(ki fe) (XiX 2 + MiM2 + VM) = 0. 

Since we supposed that fci =(= fe, we conclude that XiX2 + MiM2 
+ KM = 0, from which the theorem follows in virtue of Corol- 
lary 1 of Theorem 13, Chapter III, Section 36, page 64. 

THEOREM 24. If a quadric surface has three distinct discriminating 
numbers, then there exist three mutually perpendicular principal 
directions for the surface; if there are two distinct discriminating 
numbers, one principal direction is defined, and the second and third 
principal directions are any directions perpendicular to the first; if 
there is only one discriminating number, the principal directions are 
entirely arbitrary. 

This theorem follows from Theorems 18, 19, and 23 and the ob- 
vious facts that if the three discriminating numbers are distinct, 
each of them is a simple root of the discriminating equation; if 
there are two distinct discriminating numbers, one of them is a 
simple root and the other a double root; and if there is only one 
discriminating number it must be a triple root of the equation. 

91. Exercises. 

1. Determine for each of the following surfaces, wnether or not an asymp- 
totic cone exists; set up the equation of this cone in the cases in which one is 
present, and indicate whether the cone is proper or degenerate, real or imagi- 
nary: 



W) + | = 2z; (h) x*=pz. 

2. Discuss the asymptotic cone for the surfaces in parts (a), (b), (c), and (d) 
of Exercise 7, Section 86. 



196 QUADRIC SURFACES, GENERAL PROPERTIES 

3. Determine the diametral planes of the surface 



which correspond to the following directions : 

(a) X : M : v = 1 : 1 : 1, (6) X : M : v = 4 : -1 : 8, (c) X : M : v = 6 : - 
2 : -3. 

4. Are there any directions with which no diametral plane of the surface in 
the preceding exercise is associated? Are there directions for which the asso- 
ciated diametral plane is not at finite distance? 

6. Determine for each of the following surfaces the directions for which 
there is no diametral plane or no diametral plane at finite distance: 

(a) 2z 2 -f 20?/ + 18z 2 - 12 yz + 12 xy + 22x + y - 2 z - 5 = 0; 

(6) 2 x 2 + 20 1/ 2 + 18 z* - I2yz + I2xy + 6 x + 1677 + 62 - 3 = 0; 

(c) 36 x* -f 4 2/ 2 -f z 2 - 4 yz - 12 zx -f 24 xy + 4 s -f 16 y 26 z -f- 1 = 0; 

(d) 2 x 2 - 7 7/ 2 -f 2 z 2 - 10 yz - 8 zz - 10 xy + 6 x + 12 y - 6 z + 5 = 0. 

6. Prove that a quadric surface which has a single center (proper center or 
vertex) has a diametral plane at finite distance associated with every direction. 

7. Prove that a quadric surface with a line of centers has a diametral plane 
at finite distance for every direction except one. 

8. Prove that for a quadric surface with a plane of centers there exists 
an infinite number of directions with which no diametral plane at finite 
distance is associated. 

9. Set up the discriminating equation for each of the following surfaces: 

(a) 2* 2 -72/ 2 -f 2z 2 - Wyz-8zx- Wxy + 4x - 2y + 3z - 7 = 0; 

(6) 2z 2 +22/ 2 + 2* 2 +4a* + 42/-f-5 = 0; 

(c) 2x 2 +22/ 2 +22 2 -h27/z + 22x-h2xi/ + 4x-}-42/-f 4z + 3 =0; 

(d) x 2 -f 4 y 2 + z 2 - 4 xy - 12 yz -f 6 zx - x -f 2 y -f 5 z = 0. 

10. Determine three principal directions for each of the following surfaces: 

(a) 2* 2 -f 2?/ + 2z 2 + 2^ + 2*r + 2z2/ + 4o;--4y-h4e-f-3 = 0; 
(6) x 2 + if + 2z* - 4xz + 2xy -f 1 = 0; 

(c) x 2 + i/ 2 - 2 z 2 -f 4 7/2 + 4 zx + 8 ZT/ - G x -f 5 y - 4 z -f 6 = 0; 

(d) x 2 -f z 2 -f 2 XT/ -f 2 zz - 2 t/z - 2 x + 4 y - 4 = 0; 

(e) 13 x 2 -f 13 2/ 2 -f 10 z 2 + 4 yz + 4 zx -f 8 x?^ - 3 x - 4 1/ 4- 2 z - 6 0; 

-5 = 0. 



CHAPTER VIII 
CLASSIFICATION OF QUADRIC SURFACES 

92. Invariants. The properties of quadric surfaces which were 
discussed in the preceding chapter, such as the existence of an 
asymptotic cone, of straight lines on the surface, of centers and of 
diametral planes, do not depend in any way on the frame of 
reference which is used; they are intrinsic properties of the surface. 
The algebraic magnitudes and relations involving the coefficients 
of the equations of the surfaces, by means of which these proper- 
ties were characterized, must therefore be preserved when a new 
reference frame is introduced. This fact is expressed by the 
statement that these expressions and relations are invariant with 
respect to the transformation of coordinates which carry us over 
from one reference frame to another. Conversely, it is to be ex- 
pected that an expression or a relation involving the coefficients 
of the equation of a surface which remains unchanged under such 
a transformation of coordinates, has an important bearing on the 
intrinsic geometrical properties of the surface. Indeed the search 
for such expressions and relations furnishes a method for the 
systematic study of these properties; it is therefore of fundamental 
importance in the entire field of Analytical Geometry. We shall 
undertake now to prove the existence of a number of such expres- 
sions and relations. But before doing so, we shall give an exact 
definition of the concept of invariance and we shall illustrate it by 
a few familiar examples. 

DEFINITION I. An expression Involving the coefficients In the equa- 
tion of a surface in Cartesian coordinates, and numbers which depend 
on these coefficients, are called invariants of the surface with respect 
to a transformation of coordinates which leads to another Cartesian 
reference frame, If they remain unchanged when the coefficients of 
the equation are replaced by the corresponding coefficients of the 
equation obtained from the given one by such a transformation of 
coordinates; relations between the coefficients which are preserved 
under such a transformation are called invariant relations with respect 
to the transformation. 

197 



198 



CLASSIFICATION OF QUADRIC SURFACES 



Examples. 

1. The unsigned distance from the origin to the plane ax + by + cz -f d = 
d 



is given by 



Va 2 



(compare Corollary 1 of Theorem 7, Chapter IV, 



Section 44, page 79). This distance remains the same no matter what 
Cartesian reference frame is used, so long as the origin is not changed. Con- 
sequently, if we put 

x = \iXi -h X 2 i/i 4- X 3 Zi, y = MiZi + M22A + MsZi, ^ = "1*1 + "22/1 -f "aZi 

in the equation ax -f- by -f cz + d = and if we suppose that the equation of 
the plane is thereby transformed into a'xi + b'yi -f c'zi -f d' = 0, then it must 



be true that 
I d 



Va' 2 + 6' 2 -f 



rf 



6 2 



If this is so, the expression 



\Va 2 + 6 2 -f c 2 



will be called an invariant of the plane with respect to the 



linear homogeneous transformation indicated above. To verify this fact, we 
observe that 

a' - a\i -f km + cvi t b f = oX 2 + &M2 + ^2, c' = aX 3 + &M3 + cs, d f =d. 
Therefore 



-f- V* + c'* = (Xx 2 
c 2 -f 2 (MH>I -f 



X 3 2 )a 2 



-f- M2 



+ 2 



"2X2 -f v 3 X 3 )ca -f 2 (Xi/*i 



But it follows from Theorem 6, Chapter V (see Section 65, page 123) that the 
coefficients of a 2 , 6 2 , and c 2 are each equal to unity and that the coefficients of 



be, ca, and ab vanish; hence a' 2 -f- b' 2 -f- c' 2 
d 



-f- fc 2 + 



Thus we have 



proved that 



is an invariant of the plane with respect to ro- 



Va 2 + 6 2 -f c 2 

tation of axes. Properly speaking this expression is an invariant with respect 
to rotation of axes of the configuration consisting of the plane and the origin. 
2. The angle between the planes a& -f biy + c\z + d\ =0 and a z x -f 6 2 ?/ 
-f c 2 z -h ^2 = is independent of the reference frame that has been used to 
represent these planes. Therefore the expression for the cosine of this angle, 
given in Theorem 9, Chapter IV (Section 46, page 82), should be an invariant 
with respect to a transformation from one rectangular Cartesian system to 
another. Such a transformation is given by the equations 



(See Theorem 8, Chapter V, Section 66, page 126.) 
If these expressions transform the equations of the given planes to 
ai'si + 6/2/1 + ci'zi -f di' = and a*'xi + bfa + c^ + d/ = 

we have 

a\ = aiXi 4- bim -f- civi t bi f 01X2 + 61^2 -h CM, c\ = 01X3 

efi' == aip -f 619 -f cir, 
as' = a 2 Xi + bzfjii -h cji^i, 62' = ^2X2 



INVARIANTS OF A QUADRIC SURFACE 199 

It follows from example 1 therefore that 

oi 2 + bi* + fi 2 = fli' 2 + 61" -f <V 2 and a 2 2 + b-. 2 + c-> 2 = a 2 ' 2 + V 2 + c^. 
Moreover 



iV -f &iV + d'c 2 ' = OiCuCV + X 2 2 + X 3 2 ) + M 2 <V + M2 2 + Ms 2 ) -f 
eic 2 0i 2 -f- ^ 2 2 -f "3 2 ) + (aik -f 0261) (Xi/ii + X 2M2 + X 3/ u 3 ) + (6ic z -f 



Consequently 

Oia 2 -f bj) 2 -f 



Vfl!* + V + c, 2 X Va 2 * -f 6 2 2 + r 2 * 

is indeed an invariant with respect to the linear transformation of coordinates. 
In particular we notice that the relation a\Qv -f- bj) 2 -f- r i^ 2 = is invariant; 
the reader should see the geometric significance of this fact. 

93. Invariants of a Quadric Surface with respect to Rotation 
and Translation of Axes. We proceed now to the following im- 
portant theorems. 

THEOREM 1. The functions T t , T 2 , A and A of the coefficients of a 
quadric surface are invariant under translation of axes. 

Proof. Translation of axes is accomplished by means of the 
transformation x = x' + 7?, y == y' + q, z = z' + r (see Theorem 
2, Chapter V, Section 61, page 115). The equation of the quadric 
surface Q(XJ y, z) = with respect to the new reference frame is 
therefore Q(x f + p, y' + q, z' + r) = 0. But, in virtue of Section 
75 and the notation introduced in Section 80, this equation may 
be written in the form: 

(1) Q(P, q, r) + x'Q^p, q, r) + y'Q z (p, q, r) + z'Q,(p, q, r) 
+ q(x', y', z') = 0. 

It follows from this that the coefficients of the second degree terms 
in the new equation of the surface are the same as those of the 
corresponding terms in the original equation; that is, if we differ- 
entiate between the new and the original equation by the use of a ', 

ll' = Oil, 012' = 012, Ol3 7 = 13, ^22' = 22, O^ = O 2 3, 

033' = 033. 
Moreover, we see at once that 

OM' = I ' Qi(p, q, r), ^ = \ Q,(p, g, r) a 34 ' = \ Qs(p, g, r\ 
o 44 ' = Q(p, g, r). 



200 



CLASSIFICATION OF QUADRIC SURFACES 



Therefore (see Section 89, page 192) 



1 = 


On + 2 2 + 3 3 = On + 022 + 33 = 1 i] 


T / 

2 


022' 023' 


^11 #13 . Ctn di2 O22 O 2 3 . On Oi3 




023' 033' 


Ois 033 Oi2 O22 O23 033 OM (133 


+ 


On Oi2 


3=5 ^; 






Oi2 O22 








Oi/ Oi 2 ' Oi 3 ' 




OH 012 Ois 




144' = 




n f n ' n ' 
Ol2 O 22 O23 


= 


(1 12 O-22 O23 


= ^44- 






/ t f 

Oi3 O 2 3 033 




Oi3 O 2 3 ^33 





It remains to show that A' = A. From what we have already 
proved, it follows that 

Qi(p, q, r) 



A' = 



On 



r) 



2 

o, 9, r) 



ii **M l *'J.> Q 

\ f f Q(P>V,r) 
Since , L = a,-ip + a,- 2 (/ + a; 3 r + a,- 4 , i == 1, 2, 3, 4, and also 

t 



since 2 Q(p, 5, r) = pQi(p, q, r) + gQ 2 (p, 3, r) + rQ 3 (p, q, r) 
+ Q4(P> 3> r ) ( see Corollary 1 of Theorem 4, Chapter VII, Section 
81, page 162), it follows as in Section 84, page 172, that if to the 
last row of this determinant are added the products of the first 
three rows by p, 3, and r respectively, and then to the last 
column, are added the products of the first three columns by 
p, 3, r respectively, then this determinant reduces to the 
discriminant A of the surface. This completes the proof of our 
theorem. 

COROLLARY 1. The discriminating numbers of a quadric surface are 
Invariant under translation of axes. 

This theorem follows from the fact that the coefficients of the 
discriminating equation, namely, 1, 7\, T 2 , and A 44, are in- 
variant under translation of axes. 



COROLLARY 2. The rank of the matrix a 3 is invariant under transla- 
tion of axes. 



INVARIANTS OF A QUADRIC SURFACE 201 

COROLLARY 3. The rank of the matrix a 4 is invariant under trans- 
lation of axes. 

Proof. The proof of the invariance of A shows that the matrix 
a/ is obtained from the matrix a 4 by means of elementary trans- 
formations (see Definition XIV, Chapter I, Section 10, page 18) ; 
it follows therefore from Theorem 14, Chapter I, that these two 
matrices have the same rank. 

THEOREM 2. The functions T i9 T 2 , and A* are invariant under rota- 
tion of axes. 

Proof. The proof of this theorem and of the next could be 
made by the direct method followed in the proof of Theorem 1, 
which consists in first expressing the coefficients of the new equa- 
tion in terms of those of the given equation and then substituting 
these expressions in the function whose invariance we wish to 
prove. But this method, besides being laborious, does not give 
us any further insight into the geometric meaning of the theorem. 
We shall therefore follow a method of proof which is apparently 
less direct and which may impress the reader as being rather so- 
phisticated, but which has the merit, apart from greater elegance 
and brevity, of penetrating more deeply into the problem under 
consideration. 

We consider the function q(x, y, z, fc) defined as follows: 
q(x, y, z, k) = q(x, y, z) - k(x 2 + y 2 + z*). 

Let an arbitrary rotation of axes carry the function q(x, y, z) over 
into the function q'(x', y', z'). Since the expression x 2 + y 2 + z 2 
represents the square of the distance from the origin to the point 
(x, ?/, 2), it is invariant under rotation of axes; that is, x 2 + y 2 + z 2 
x' 2 + y' 2 + z' 2 . Hence, if the same rotation of axes changes 
the function q(x, y y z, k) to q'(x' y y', z', fc), we have 

q'(x', y', z', fc) = '(*', 2/', *') - k(x' 2 + y' 2 + z' 2 ). 

The equation q(x, y, z, k) = 0, being homogeneous in x, y, and 
z, for every value of fc, represents a quadric cone; this will be a 
degenerate quadric cone (that is, a pair of planes) if and only if 
k has such a value fc* that the determinant 

(in fc* 012 013 

012 022 ~~ fc* 023 

013 023 033 fc* 



the value of the determinant 



202 CLASSIFICATION OF QUADRIC SURFACES 

vanishes. If this determinant vanishes, the equation q(x, y, z, 
fc*) = represents a pair of planes; therefore the equation 

q'(x f , y', z', k*) = represents a pair of planes and consequently 

/ ?/.* ^ t si ' 

11 K Cti2 #13 

12 7 ' O22' k* 023' 

/ / ~ t TLsfc 

is also equal to zero. And it should be clear that the same argu- 
ment holds in the opposite direction. From this we conclude, in 
view of Corollary 1 of Theorem 19, Chapter VII (see Section 89, 
page 192), that the two equations 
fc 3 - Tik* + T 2 k - Au = and fc 3 - Ti'k* + T 2 'k - AJ = 

have the same roots, that is, TI = TI, T 2 = TV, Au = A^', thus 
our theorem is proved. 

If r 3 , the rank of the matrix 83, is equal to 3, Au 4= 0; therefore 
Au ^p and 7-3' = 3. Ifr 3 = 2, sothat-A 4 4 = 0, then^ 4 / = Oand 
r 3 ' < 3. If r 3 = 1, all the two-rowed minors of a 3 vanish and 
therefore 5T 2 = 0; hence, it follows from Theorem 2, that A^ ~ 
and TV = 0, and thence by use of the Corollary of Theorem 7, 
Chapter II (Section 26, page 44) that r 8 ; < 2. If r 3 = 0, the 
function Q(x, y, z) is of the first degree; therefore the function 
Q*%&9 y r > 2') is also of the first degree (compare Corollary 1 of 
Theorem 8, Chapter V, Section 66, page 126) and r/ = 0. But 
this entire argument can be applied equally well to the transfor- 
mation which carries Q' back into Q. It follows therefore that if 
r 3 ' = 2, then r 3 < 3; if r 3 ' = 1, then r 3 < 2; and if r 3 ' - 0, then 
r 3 < 1. We have therefore obtained the following important 
corollary. 

CQHOLLARY. The rank of the matrix a 3 is invariant with respect to 
rotation of axes. 

94. Invariance of the Discriminant of a Quadric Surface with 
respect to Rotation. We shall begin by proving the following 
theorem. 

THEOREM 3. The singularity of a quadric surface is not affected by 
rotation of axes. 

Proof. In view of Definition V of Chapter VII (Section 82, 
page 166) the statement of this theorem is equivalent to the 
statement that if A = 0, then A ; = 0, where A' is formed from 



INVARIANCE OF THE DISCRIMINANT 



203 



the equation obtained from Q(x, y, z) = by rotation of axes; or 
again, to the statement that if r 4 < 4, then r 4 ' < 4, where r 4 ' 
designates the rank of the matrix a 4 ' formed from this same equa- 
tion. All the cases in which r 4 < 4 have been specified geometri- 
cally in Remark 1, following Theorem 14, Chapter VII (Section 
85, page 179), excepting the case in which r 4 = 3 and r 3 = 1. In 
this case, we know on the basis of the Corollary to Theorem 2 
(Section 93) that r 3 ' = 1 and hence, in view of the discussion 
preceding Theorem 14 of Chapter VII, that r/ can not exceed 3. 
We conclude therefore that in every case in which r 4 < 4, we must 
also have r/ < 4. Our theorem is therefore proved. 

THEOREM 4. The discriminant of a quadric surface is invariant with 
respect to rotation of axes. 

Proof. The method of proof is similar to that used in the proof 
of Theorem 2. We consider now the auxiliary function 

Q(x, y, z, k) = Q(x, y, z) - k(x* + y* + z* + 1). 
Rotation of axes will carry this function over into 

QV, 7/, z', k) = QV, '/, ') - k(x'* + y'* + z'* + 1). 

A value fc* of k for which the locus of the equation Q(x, y t z, k) = 
is singular will, in virtue of_the preceding theorem, also be a vaidc 
of k for which the surface Q'(x', y', z', k) = is singular. Hence 
the roots of the equation 



= 



0ii - k 


012 


013 


014 


012 


022 k 


023 


024 


013 


023 


033 k 


034 


014 


024 


034 


ot 44 k 



will also be roots of the equation 

11 - k 

012' 

013' 

014' 



f 

012 


013' 


f 
0J4 


22 ' ~ k 


023' 


024' 


023' 


033' k 


034' 


024' 


034 7 


a 44 ' k 



= 0, 



and vice versa. These equations are therefore equivalent. Now 
it should be obvious that they have the form 

fc 4 + . . . + A = and k 4 + . . . + A' = 0; and therefore 
that A = A'. This proves our theorem. 



204 



CLASSIFICATION OF QUADRIC SURFACES 



It will be worth while to consider in further detail the equation 
A (k) = 0, which is quite similar in form to the discriminating 
equation considered in Section 89. If we use again Theorem 19 
of Chapter I (see Section 17, page 32), we find 



A'(fc) = - 



#22 k #23 #24 




# n -fc 


#13 #14 


#23 #33 k #34 
#24 #34 #44 k 




#13 
#14 


#33 ~~ k #34 
#34 #44 k 


#11 k #12 #14 




#11 -fc 


#12 #13 


#12 #22 ~~ k #24 





#12 


#22 k #23 , 


#14 #24 #44 k 




#13 


#23 #33 k 


| #M #44% 


+ 


#22 ~~ k #24 
#24 #44 ' 


c H 


#11 k #14 

#14 #44 ~~ k 


#22 k #23 
#23 #33 k 


+ 


#11 ~fc #13 
#13 #33 k 


+ 


#11 fc #12 ll 

#12 #22 fc 1 J 



A" ' (k) = 6 [(#11 k) + (#22 ~ k) + (#33 A:) + (# 44 fc)], 
A" " (fc) = 24. 

Since, moreover, A(fc) = A(0) + A'(0) X fc + A" (0) X ~ 

+ A" ' (0) X || + A" " (0) X ^, the equation A (fc) = 

can be written in the form 

fc 4 - D x fc 3 + # 2 fc 2 ~ D 8 k + A = 0, 



D 2 = 



6 

A" (0) 



= #n + # 22 + #33 + #44, 
4 



and 



- 2 



It will be observed that D\, D 2 , and D 3 are respectively the sums 
of the one-rowed, the two-rowed, and the three-rowed principal 
minors of the discriminant A. 
We have now the following Corollary of Theorem 4. 

COROLLARY. The sums of the one-rowed, of the two-rowed, and of 
the three-rowed principal minors of the discriminant of a quadric 
surface are invariant with respect to rotation of axes. 



INVARIANTS OF A QUADRIC SURFACE 205 

It follows moreover from Theorem 4 that, if r 4 = 4, then r 4 ' = 4; 
and that if r 4 = 3, then r 4 ' < 4. If r 4 = 2, all the three-rowed 
minors -of a 4 vanish and therefore Z) 3 = 0; the corollary enables 
us then to conclude that Z> 3 ' = and the Corollary of Theorem 7, 
Chapter II (Section 26, page 44) establishes then the fact that 
r\ < 3. Similarly it can be shown* that if r 4 = 1, then r/ < 2; 
and it should be clear that if r 4 = 0, then r/ = 0. Moreover the 
argument can be made equally well from the rank of A' to that of 
A. We have therefore obtained the further result, stated in the 
following theorem. 

THEOREM 5. The rank of the matrix a 4 is invariant with respect to 
rotation of axes. 

The results obtained in Sections 93 and 94 may be summarized 
in the following statement : 

The values of the expressions A, /i< 4 , 7\, T 2 and the ranks of 
the matrices a 3 and a 4 are invariant with respect to translation 
and rotation of axes; the expressions A>,, />>, and D 3 are invari- 
ant under rotation of axes. 

95. Exercises. 

1. Prove that the condition under which three planes have a single point in 
common is invariant with respect to translation of axes, and also with respect 
to rotation of axes. 

2. Prove that the distance from the plane ax -f- by + cz -f d = to the 
point P(XI, 2/1, 21) is invariant with respect to translation and rotation of axes. 

x 2 ?/ 2 z 2 

3. Show that for the surface -f ~ -f 1 =0, the sum A, of the 

three-rowed principal minors of the discriminant is not invariant with respect 
to translation of axes. 

X 2 y 2 

4. Show that for the surface 5 ^ 1 =0, the sum D 3 is invariant with 

a 2 b 2 

respect to translation of axes; also that the sum 7) 2 of the two-rowed prin- 
cipal minors of the discriminant and the sum D\ of its one-rowed principal 
minors are not invariant with respect to this transformation of coordinates. 
6. Show that for the surface x 2 = a 2 , the sums D 3 and D> 2 are invariant with 
respect to translation of axes; also that the sum A is not invariant. 

6. Prove that if the axes are translated to the new origin P(a, 0, 7), the 
sum ZY for the new equation is equal to Z> 3 plus multiples of three-rowed 
minors of the matrix b (compare Section 85, page 178). 

7. Prove that under the conditions of Exercise 6, the sum TV for the new 
equation is equal to D 2 plus multiples of two-rowed minors of the matrix b. 

* See Appendix, IV, p. 297. 



206 CLASSIFICATION OF QUADRIC SURFACES 

96. Two Planes, We have already met a number of instances 
in which the equation Q(x, y, z) = represents two planes. In 
the present section we undertake a more detailed study <tf these 
cases; and we begin with the following theorem. 

THEOREM 6. The necessary and sufficient condition that a quadrlc 
surface consist of two planes is that the rank of r 4 of the discriminant 
matrix a be less than 3. 

Proof. The sufficiency of this condition has already been 
proved in Corollary 2 of Theorem 12, Chapter VII (Section 84, 
page 175). To prove the necessity of the condition, we observe 
that if the locus of the equation Q(x, y, z) = consists of two planes 
then, by the argument made in the proof of this corollary, the 
function Q(x, ?/, z) must be factorable in two linear factors; that 
is, there must exist numbers a, 6, c, d and ai, 61, ci, di such that 



Q(x, y, z) = (ax + by + cz + d) (aix + biy + c\z + di). 

If this is the case, the coefficients of the function Q can be ex- 
pressed as follows: 

an = aai, a 22 = 661, ass = cci, a 44 = ddi, 2 a i2 = a&i + a\b, 
2 ais = aci + aic, 2 a H = adi + aid, 2 a 23 = bc\ + &ic, 
2 a 24 = &di + &id, 2 a 34 = cdi + Cid. 

Consequently the discriminant A is given by the equation 



16 A = 



2 aai ab\ + a\b ac\ + a\c ad\ + aid 

ob\ + a\b 2 bhi bc\ + b\c bd + 6id 

aci + aic 6ci + b\c 2 cci cdi + Cid 

adi + aid &di + 6id cdi + Cid 2 ddi 



It is not difficult to show that this determinant and its three- 
rowed principal minors vanish (the details of this proof will be 
found in Appendix, V, page 298). But since the matrix of this 
determinant is symmetric, we can then conclude by use of Theorem 
6, Chapter II (see Section 26, page 43) that the rank of the matrix 
a 4 is less than 3. 

It was proved in Corollary 3 of Theorem 14, Chapter VII (Sec- 
tion 85, page 180), that if r 4 = r 3 = 2, the locus of the equation 
Q(x y i/, z) = consists of two intersecting planes. We shall now 
determine the equations of these planes. 



TWO PLANES 207 

Since ra = 2, it follows that at least one of the two-rowed prin- 
cipal minors of the matrix a 3 is different from zero; let us sup- 
pose that asa = flute 0i2 2 4= 0. From Theorems 12 and 13 of 
Chapter I (see Section 7, page 13), we derive then the following 
equalities : 

^ <*33013_ = 0, 1 3 012 + 2 322 +..#33023 = ^j 
#23023 + #33033 = 0, #1 3 014 + #23024 + #33034 = 0. 

And if we denote by 0# the cofactors of the elements a# in the de- 
an 012 



terminant 



012 022 023 



, we have also (notice that 3 4 = # 33 4 1 0) 



014 024 034 

the equalities 

013011 + 023012 + 034^14 = 0, 013012 + 023022 + 034024 = 0, 
013013 + 023023 + 034034 = 0, 01 3 0H + 023024 + 034044 = 0. 

If we multiply the equalities of the first of these sets by x, y, z, and 
1 respectively and add, we obtain i 3 Qi + 0:23^2 + 33 Q 3 = 0; sim- 
ilarly, we find from the second set the result 0i 3 Qi + 02sQ2 + 034$4 
= 0. And from these equations we conclude (remembering that 
3 4 = <* 33 4= 0), that 

2 



But, a 33 y 01232023 = (011022 012 2 )2/+ (011023 ~ 012013)^+ (011024 

012014) = 011^2 012^1 ,' 

and a33^-~Ctl3^ 013= (011022 012 2 )^+(022013~012023)2+(0 22 014 

012024) = 022Ql 012^2- 

Therefore, 2 a 33 Q(a;, y, z) = (anQ 2 ai2<2i)Q2+ (022^1 - 



Since the discriminant of this quadratic function of Qi and Q 2 is 
equal to 4 33 and is therefore different from zero, we conclude that 
the equation Q(x, y,z) =0 is equivalent to the two linear equations 
Qz XQi = and Q 2 /iQi = 0, where X and n are the distinct 
roots of the quadratic equation a^t 2 2 12 ^ + 022 = 0. The 
two planes represented by the equation Q = in this case are there- 
fore distinct planes through the line of intersection of the planes 
Qi = and Q 2 = (compare page 180). It follows moreover 
that if 33 > 0, the two roots of the quadratic equation ant 2 2 a 12 t 



208 CLASSIFICATION OF QUADRIC SURFACES 

+ 0-22 = are complex, whereas they are real if #33 < 0. Finally, 
we observe that the sign of 33 is the same as that of the invariant 
7 7 2 , which is equal to the sum of the two-rowed principal minors of 
the matrix a 3 , in virtue of Theorem 7 of Chapter II (see Section 
26, page 44). We summarize the result in a theorem. 

THEOREM 7. If the ranks of the matrices a 4 and a, are both 2, the 
locus of the equation Q(x, y, s) = consists of two planes; these 
planes are real if the invariant T 2 is negative, and imaginary if T 2 is 
positive. 

We consider next the case in which r 4 = 2 and r 3 = 1 ; it was 
shown on page 178 that in this case the rank of the matrix b 
is also equal to 1. Consequently the three rows of this matrix 
are proportional; and if we suppose that a n ^ 0,* we can write 



and Q 3 = !. Moreover Q 4 = q, + 2 



#11 (hi #11 

_i_ o a ' 4 v (n 9 \ _L o a nQi 2(a 14 2 

+ 2 (7 4 4 = -- X (Ql 4 flu) + ^ #44 = 

an #11 #11 

Therefore 



2(a H 2 



#n 

and the equation Q(x, y, z) = is equivalent to the equation 
d 2 = 2(#i4 2 - 011044). 

It is shown in Appendix, VI (page 299) that if r 4 = 2, the two- 
rowed principal minors of the matrix a 4 can not all vanish and that 
those which do not vanish all have the same sign and therefore the 
sign of their sum D 2 . Since r 3 = 1, all two-rowed principal minors 
of as vanish; the rank of b being 1, the other two-rowed principal 
minors of a 4 differ by a factor and are therefore different from zero 
and of the same sign of Z> 2 . We conclude that the locus of the 
equation Q(x, y, z} = consists of two distinct parallel planes, 
whose equations are Qi = dh V2(# u 2 #11044); these planes will be 
real or imaginary according as Z) 2 is negative or positive. 

* If all the elements in the principal diagonal of a 3 were zero, it would 
follow since in this case an = a 22 = 3 3 = 0, that the elements au, a 2 a, and 
an also vanish, so that the rank r 3 of a 3 would be zero; all the non-vanishing 
elements of the principal diagonal have the sign of the invariant TV 



TWO PLANES 209 

Finally, the same discussion shows that if r 4 = r 3 = 1, then the 
equation Q(x, y, z) = reduces to the form Qf = 0, so that its 
locus consists of two coincident planes (compare Corollary 3 of 
Theorem 14, Section 85, page 180). 

THEOREM 8. If the rank of the matrix a 4 is and the rank of the ma- 
trix a 3 is 1, the locus of the equation (X>, y, a) = consists of a pair of 
parallel planes; these planes are real or imaginary according as the 
sum D> of the two-rowed principal minors of the matrix a 4 is negative 
or positive. If the ranks of the matrices a. and a. arf froth i f thr lonin 
of the equation Q(x, y, z) = consists of the plane Q,(x f y r ) = 0, 
counted doubly. 

Remark. Since the hypotheses of Theorems 7 and 8 exhaust all 
the possibilities as to the ranks of the matrices a 4 and a 3 , subject 
to the condition of Theorem 6 that r 4 must be less than 3; and 
since the conclusions of these two theorems include all the possible 
relative positions of two planes, it follows that the converse of 
each of these theorems also holds; that is, if a quadric surface con- 
sists of two real intersecting planes, two imaginary intersecting 
planes, two real parallel planes, two imaginary parallel planes, or 
two coincident planes, the ranks of the matrices a 4 and a ;{ are 2 
and 2(T 2 < 0), 2 and 2(T 2 > 0), 2 and 1(D 2 < 0), 2 and 1 
(D 2 > 0), 1 and 1 respectively. 

We state some further consequences of our discussion. 

COROLLARY 1. If the rank of the matrix a 3 is 2, the locus of the equa- 
tion q(x, y, z) = consists of a pair of intersecting planes, whose line 
of intersection passes through the origin; if the rank of this matrix 
is 1, the locus is a pair of coincident planes through the origin. 

COROLLARY 2. A function Q(x , y, s) of the second degree is factorable 
into two linear functions of*, y, and s with real or complex coefficients 
if and only if the rank of its discriminant matrix is less than 3; it 
is the square of a linear function of *, y, and z with real or complex 
coefficients if and only if the rank of its discriminant matrix is 1. 

COROLLARY 3. A homogeneous function q(x, y, z} of the second de- 
gree in x 9 y, and z is factorable into two linear homogeneous functions 
of x 9 y, z with real or complex coefficients if and only if the rank of the 
matrix a 3 is less than 3; it is the square of a linear homogeneous 
function of *, y, and z with real or complex coefficients if and only if 
the rank of this matrix is 1. 

Corollaries 2 and 3 are obviously restatements in algebraic form 
of the results formulated in Theorems 6, 7, and 8. 



210 CLASSIFICATION OF QUADRIC SURFACES 

97. Translation of Axes to the Center of a Quadric Surface. If 

a quadric surface has a center, its equation is materially simplified 
when the axes are translated to the center as origin. For it should 
be obvious from the definition of a center that the surface is sym- 
metric with respect to such a point (compare Definition VIII of 
Chapter VII, Section 85, page 176 and the first footnote on page 
137) ; therefore, if a, fc, c are the coordinates of a point in the new 
reference frame, then Q( a, 6, c) must vanish whenever 
Q(a, 6, c) vanishes. Consequently Q(a, 6, c) Q( a, 6, c) 
= for all sets of numbers a, fc, c for which Q(a, fe, c) = 0. But 
Q(a, 6, c) Q( a, 6, c) = 2(a u a + a^b + a 34 c); if this linear 
function is to vanish for all sets of values for which the quadratic 
function Q(a, 6, c) vanishes, then a^ = #24 = #34 = 0. Conse- 
quently the equation of a quadric surface referred to a reference 
frame whose origin is a center of the surface does not have any first 
degree terms. 

We shall now reach this result in another way, which will disclose 
some further properties. The translation of axes to the point 
(a, b, c) as origin is accpmplished by means of the equations of 
transformation 

x = x' + , y = y f + b, z = z' + c. 

The equation of the surface Q(x, y, z) = with reference to the 
new system of coordinates is therefore 

Q'(x', y', z')=Q(x'+a, y'+b, z'+c)=q(x', y', *')+*'<2i(, 6, <0 
, b, c)+z'Q 3 (a, 6, c)+Q(a, 6, c) = 0, 



(compare Section 93, formula (1), page 199). 

But, if a, b, c are the coordinates of a center of the surface, 
Qi(a, 6, c) = Q 2 (a, 6, c) = Q 3 (a, b, c) =0 (see Theorem 13, Chapter 
VII, Section 85, page 177) ; in this case the equation of the surface 
reduces therefore to the form 

(*', 2/', *') + Qfo, ?>, c) = 

and this equation is free from terms of the first degree in x' t y' f 
and z'. 

We observe, moreover, (1) that the second degree terms in the 
new equation have the same coefficients as the corresponding 
terms of the original equation; and (2) that the constant term 
Q(a, 6, c) is equal to \[aQ\(a y 6, c) +bQz(a, 6, c)+cQ 3 (a, 6, c) 



ROTATION OF AXES 



211 



b> c)] = ^Q^(d t by c) = a\\a + a^b + 0340 + 044. Further- 
more the discriminant of the simplified equation is, in virtue of 
Theorem 1 (Section 93, page 199) equal to the discriminant A of 
the original equation; on the other hand it is equal to 

Oil #12 #13 V 
Q>12 #22 #23 vl 
#13 #23 #33 **,, ( y 

^ 

Therefore A = ^4 44 ~; if the surface has a unique center, and 
& 

in no other case, ^.44 4^ 0, so that the constant term in the reduced 
equation can then also be put in the form -. . We summarize these 

A 44 

results as follows. 

THEOREM 9. If the quadric surface Q has a center at the point 
(a, &, c), its equation in a reference frame whose axes are parallel to 
the original axes and whose origin is at the center, has the form 



<?'(*' y', *') 



', y', *') + a 44 ' = 0, where a 44 ' = Q(a, 6, c) = 



if (a, 6, c) is the only center of the surface, we have, moreover, 044' 



98. Rotation of Axes to the Principal Directions of a Quadric 
Surface. It was proved in Theorem 24 of Chapter VII (Section 
90, page 195) that for every quadric surface there exist three 
mutually perpendicular principal directions; under some condi- 
tions these directions can be determined in one and only one way; 
under other conditions they can be determined in more than one 
way. We will suppose now that for the quadric surface Q(x, y, z) 
= three mutually perpendicular principal directions are given by 
the three sets of direction cosines Xi, MI> v\\ \z, ^ 2 > v ^ anc ^ ^ 3 > ^ 3 > V3 > 
and it is our purpose to determine the equation Q'(x', y', z 1 ) = of 
the surface when it is referred to a reference frame whose origin 
coincides with the origin of the original frame, but whose axes are 
in these principal directions. 

According to Theorem 5 of Chapter V (see Section 63, page 121), 
the transformation is carried out by means of the substitution: 



X = 



y 



212 CLASSIFICATION OF QUADRIC SURFACES 

We have therefore 
Q'(x f , y', z') = 



' + X 2 7/' + X 3 z', fjnx' + My' + &&', v\x 9 + v 2 y r 
X 2 7/ + X 3 z') + 2 a 24 ( M tz' + ny' + ^z') 
v z z') + a 44 . 

Since q(x, ?/, z) is a homogeneous function of the second degree 
in x, y, and z, and since the expressions which have been substituted 
for these variables are linear and homogeneous in #', y', and z 1 , it 
should be clear that the function q(\ix' + , v\x' + , 
vix' + ) which constitutes the first term in the new equation 
is homogeneous and of the second degree in x', y', and z'. The 
terms of degree less than 2 in the new equation can be determined 
readily; if we write that part of the new equation in the form 
2 au'x' + 2 au'y' + 2 a 34 V + 44 ', we find 



(134 = OwXs + 024M3 + 034*3 = ^ *' ^ 



It remains now to determine the coefficients of the second degree 
terms in the new equation. For this purpose we expand 
q(\\x' + , nix' + ji>ix' + ) by Taylor's theorem (com- 
pare Sections 75 and 80). First we look upon Xi#' + X 2 i/', nix'+ &y' 
and v\x 9 + v 2 y' as the (temporarily) fixed values of the variables in 
the function q(x, y, z) and upon Xsz', vtft , and v$f as their incre- 
ments. We find then 

X 3 z', AUX' + wy f + &', v&' + v 2 y' + v&') = 
q(\ix' + X 2 ?/, mx' + my', v\x' + v 2 y') + \ 3 z' qi(\ix' + \ 2 y', 
ix' + My', v& r + v 2 y') + &' q 2 (\ix' + \ 2 y', mx f + toy', 

\ 2 y' f . . . 9 . . . ) 



To the first four terms on the right we apply again Taylor's 
theorem, remembering that % = 2 a# (see Section 80); we find 



ROTATION OF AXES 213 

then that 

q'(x', y', z') = q(\ix', mx', vix') + 



ix', vix f ) + 2 a n \ 2 y' + 2 a V2 ^y f + 2 
' fe(Xio;', /Litre', vix 1 ) + 2 a 2i \2y' + 2 a^^y' + 2 
ix', v&') + 2 a 3 iX 2 7/ / + 2 a^y' + 2 



We recall once more that q is a homogeneous function of the 
second degree and that q\, # 2 , and q$ are homogeneous functions of 
the first degree; also the property of homogeneous functions of 
which we spoke in the proof of Theorem 3, Chapter VI (see Section 
70, page 136). If we make use of these facts, we should be able 
to see that the second degree terms in Q'(x', y', z') reduce, under a 
^i>c,rai rot^tv of axes, to 

" ' / a) +x'y r [ 

y'z' [X 3 ^i 



If in this expression we make use of the formulas established in 
the Corollary of Theorem 18, Chapter VII (see Section 88, page 
190) for the direction cosines of the principal directions, this ex- 
pression reduces to 



Finally we put into operation the hypothesis that the new coor- 
dinate axes are mutually perpendicular and that therefore their 
direction cosines satisfy, two by two, the condition of Corollary 1 
of Theorem 13, Chapter III (see Section 36, page 64) ; our second 
degree terms then become /bix' 2 + k z y' 2 + & 3 z' 2 . We have there- 
fore obtained the result stated in the following theorem. 

THEOREM 10. If the discriminating numbers of a quadric surface 
Q are fci, & 2 , and fc 3 and if the frame of reference is rotated so that the 
new X-, Y-, and Z-aies have the directions of the principal directions 
determined by fci, k 29 and fc s respectively, then the equation of the 
surface with respect to the new frame is 

<?'(*', y', *') = fci* /2 4- fetf' 2 + k*z'* + q<(\ l9 Ml , *) 
s> vz)s' + a 4 4 = 0. 



214 CLASSIFICATION OF QUADRIC SURFACES 

Remark. The phrase "the principal directions determined by 
fci, fc 2 , and & 3 " used in the statement of this theorem is to be under- 
stood in the same sense as in Theorem 24, Chapter VII (see 
Section 90, page 195). 

99. Classification of Quadric Surfaces the Non-singular 
Cases. We are now in a position to analyze the general equation 
of the second degree in x, y, and z, that is, to determine the types 
of surfaces that can be represented by the equation Q(x, y y z) 0. 
The problem of making this determination is usually referred to 
as the " classification of quadric surfaces."* 

The analysis of the equation Q(x, y, z) = will be based on the 
ranks r\ and r 3 of the matrices a 4 and a 3 respectively; and we treat 
first those surfaces for which r 4 = 4, that is, the non-singular 
quadrics. In virtue of Corollary 1 of Theorem 14, Chapter VII 
(see Section 85, page 179), this condition carries with it that 
7*3 ^ 2; we have therefore to consider two cases, namely, n = 4, 
7*3 = 3; and r 4 = 4, r 3 = 2. 
CASE I. r 4 = 4, r 3 = 3. 

We know from Theorem 14, Chapter VII (Section 85, page 178) 
that the surface has a single proper center. If the axes are trans- 
lated to this center as an origin, the equation becomes (see Theorem 
9, Section 97, page 211) 

(1) q& 9 y' 9 *)+-=0. 

A 44 

The three roots of the discriminating equation 

(2) A: 3 - 7W + TJc -.444 = 

are all real and different from zero. Since the first degree terms are 
absent from equation (1), rotation of axes to principal directions 
will carry the equation over into 

fci*" 2 + W 2 + fcaz" 2 + -- = 0, 

^44 

* This problem concerns itself therefore primarily with the question of 
determining what kind of surface is represented by given numerical equations 
and not with that of locating the position of the surface with respect to a 
frame of reference, nor with finding the particular numerical data which 
serve to specify the surface as an individual of its type. The method of 
treatment of OUT principal question is of such nature however as to develop 
means for answering these further questions. 



THE NON-SINGULAR CASES, 215 

kiy kz, fc 3 being the roots of equation (2), (see Theorem 10, Section 
98, page 213). Since A 4= 0, this equation may be written in the 
form 

r "2 ?y 2 "2 

- + -=-T- = I. 




This equation belongs to the types of equations whose loci were 
discussed in Section 72. If we make use of the discussion of this 
section, we reach the following conclusion : 

(a) If ^ -r- , j -r~ , ; - A are all negative, the surface is an 
/CiA 44 KzAu KsAu 

ellipsoid. 

* (6) If two of these numbers are negative, the surface is an 
hyperboloid of one sheet. 

(c) If one of these numbers is negative, the surface is an hyper- 
boloid of two sheets. 

(rf) If none of these numbers is negative, the surface is an 
imaginary ellipsoid. 

Remark. By reference to Example 2, Section 68, page 133, wo 
see furthermore that, if the discriminating equation has a pair of 
equal roots, the quadric surface will be a surface of revolution, 
namely, an ellipsoid of revolution (real or imaginary), or a hyperbo- 
loid of revolution (of one sheet or of two sheets), according as the 
sign of the double root does or does not agree with that of the re- 
maining root; if and only if the discriminating equation has a 
triple root, the quadric will be a sphere (real or imaginary). 

We observe that the complete determination of the character of 
the surface depends in this case on the signs of A and ^4 44 , and on 
the signs of the roots of the cubic equation (2), whose coefficients 
are all invariant with respect to translation and rotation of axes. 
Since the roots of this cubic are all real (compare Theorem 20, 
Chapter VII, Section 89, page 192), Descartes' Rule of Signs 
enables us to tell exactly how many positive and how many nega- 
tive roots it has. If the signs of T\ and of ^4 44 are both changed, 
all the roots of the cubic change sign, and therefore the numbers 

? . preserve their signs for i = 1, 2, 3; hence we need consider 



216 CLASSIFICATION OF QUADRIC SURFACES 

only the sign of the product TiA^* We distinguish now the fol- 
lowing cases: 

(1) A > 0, T<t > 0. In accordance with the remark made 
above, the sequences of sign in the cubic which have to be con- 
sidered are the following: 

When ^44 > 0, 7\ > 0, the signs are H 1 ; 

and, when AU > 0, TI < 0, the signs are + + -\ . 

If the first of these occurs, the equation has three positive roots 
and the three " coefficients "T-T~ , * = 1, 2, 3 are positive; the sur- 
face is therefore an imaginary ellipsoid; if the second sequence 
occurs, there are two negative roots and one positive root and 
hence two negative and one positive coefficients, so that the sur- 
face is an hyperboloid of one sheet. 

(2) A < 0, T 2 > 0. The sequences of sign are the same as 
before, but since now A < 0, all the coefficients will have changed 
their signs. Therefore we shall have an ellipsoid if A^Ti > 0, 
and an hyperboloid of two sheets if A^Ti < 0. 

(3) A > 0, T 2 < 0. If Au > 0, the sequence of signs will 

be + H or H ,so that we have one positive and two 

negative roots and also one positive and two negative coefficients. 

If A 4 4 < 0, the sequences of signs are + H h or H h, so 

that there are one negative and two positive roots but, since A 44 
has changed sign, again one positive and two negative coefficients. 
In this case therefore the surface is always an hyperboloid of one 
sheet. 

(4) A < 0, T 2 < 0. We have the same distribution of roots 
as in (3), but, since A has the opposite sign, the coefficients will be 
opposite in sign; the surface is therefore an hyperboloid of two 
sheets. 

It remains to consider the cases in which either T\ or T 2 vanishes; 
that they can not vanish simultaneously was shown in the Corollary 
of Theorem 21, Chapter VII (see Section 89, page 194). If either 

* It should be clear that the signs of T\ and of Au can not be significant in 
determining the character of the locus of the equation Q = 0. For, if this 
equation is multiplied through by -1, TI and A 44 clearly change their signs, 
but the locus of the equation is obviously not affected. This remark does not 
apply to TI, A or 7\A 44 . 



THE NON-SINGULAR CASES 



217 



Ti or 5T 2 vanishes, the roots can not all have the same sign; for 
since TI = ki + k 2 + k 3 and T 2 = fcife + & 2 /c 3 + & 3 fci, the former of 
these expressions would then have the sign common to the roots 
and the latter would be positive. Moreover AU = fcifc 2 fc 3 ; hence, if 
A> and A& > 0, there must be one positive and two negative roots 
and also one positive and two negative coefficients, and if A > 
and AM < 0, there are one negative and two positive roots, but 
again one positive and two negative coefficients. In either case 
the surface is an hyperboloid of one sheet. 

But if A < 0, there will be one negative and two positive coeffi- 
cients, so that the surface is an hyperboloid of two sheets. 

We summarize the results in the following theorem. 

THEOREM 11. If the ranks of the matrices a 4 and a 3 are 4 and 3 re- 
spectively, the locus of the equation Q = will be determined by the 
following table: 





A> 


A <0 


T 2 > 0, ATi > 


Imaginary ellipsoid 


Ellipsoid 


T 2 > o, AuTt ^ o 

or 
T 2 ^0 


Hyperboloid of one 
sheet 


Hyperboloid of two 
sheets 



Remark. We observe that the character of the surface can be 
completely determined in this case as soon as the signs of the in- 
variants A, AM, T 2 and TI are known and that it is not necessary 
for this purpose to solve the discriminating equation. Compare 
also the Remark on page 215. 
CASE II. r 4 = 4, r 3 = 2. 

According to Theorem 14, Chapter VII, the surface does not have 
a center in this case. Since AM is equal to zero, T 2 must be differ- 
ent from zero, for otherwise we could conclude by means of the 
Corollary of Theorem 7, Chapter II (see Section 26, page 44) that 
r 3 < 2. Consequently, one and only one root of the equation (2) 
vanishes; let it be k\. Rotation of axes to principal directions 
will then reduce the equation Q = to the form 

(3) fey' 2 + fc 3 2' 2 + #4(Xi, MI, vi}x' + g 4 (X 2 , M2, v^y' + q*(\3, MS, ?sX 

+ 44 = 0. 



218 



CLASSIFICATION OF QUADRIC SURFACES 



The discriminant of this equation is 
000 a 14 








a 24 ' 





a 24 ' 



where 2 ai/, 2 a 24 ' and 2 
coefficients g 4 (Xi, MI, "0, ^ 



are used, as before, to designate the 
2 , M2, ^2) and q(\ 3 , /i 3 , i> 3 ) of x', ?/', and 2' 
respectively. But the discriminant of a quadric surface is in- 
variant with respect to rotation of axes (see Theorem 4, Section 
94, page 203) ; hence A = an /2 kjc^ and, since fe =j= and 



& 3 ={= 0, an' = dby T-r ^p 0. The further reduction of the 

equation is now made as follows; completing the square on the 
terms in y' and z', it becomes 
^2 



/ fi '\ 2 / r, '\ 

, / . . a 24 y . 7 / , , 34 \ 

H y + ir) + H 2 + F) 



We translate the axes now to the point 

/ 044 , a 2 4 /2 , dU 2 __ 

V 2 ai/ "*" 2 a 14 '/c 2 "*" 2 ^4^3 ' 
as origin by putting 



/ // ^44 , ^24 " , &.34 / // ^ 2 4 

X' T* , I I /ij' I/ 

*^ o / i^ ?) TI~ i^ o // > / i/ *~7 > 

^ (7] 4 u (Ij 4 A/ 2 M dj 4 n/3 rC 2 



2 =2 



This transformation carries the equation of the surface over into 
the form 



We reach therefore the conclusion, by means of the results of 
Section 72, that in this case the locus of the equation Q = is an 
elliptic paraboloid if fc 2 and fc 3 have the same sign, and an hyper- 
bolic paraboloid if they are opposite in sign. But k\ and /c 2 are 
the roots of the quadratic equation fc 2 Tik + T 2 = 0, and there- 
fore the first or the second of these cases will arise according as 



THE NON-SINGULAR CASES 



219 



T 2 is positive or negative. We shall replace this criterion by 
another one; but before doing so we observe that the surface will 
be a paraboloid of revolution if and only if fc 2 = & 3 . 
We will prove now the following theorem. 

THEOREM 12. If A = 0, then A is the square of a linear homoge- 
neous function of a u , a 24 , and a 34 . 

Proof. Since AM = 0, the development of A according to the 
elements of its last column leads to the equation 



#12 #22 #23 
#13 #23 #33 
#14 #24 #34 



+ #24 



#11 #12 #13 
#13 #23 #33 
#14 #24 #34 



#34 



#11 #12 #13 
#12 #22 #23 
tti4 O24 #34 



If we develop each of these three-rowed determinants according to 
the elements of their last row and use the notation ,y for the co- 
factors of the elements a# of the matrix a 3 , as introduced on page 
184 in the proof of Theorem 16, Chapter VII, we find that 

A = ~#14(#1411 + #2412 
+ #2423 

2 ai 4 024ai2 + 2 a u a 34 ai 3 + a 24 2 a 22 + 2 



#1412~#2422 ~ 



It follows from Corollary 2 of Theorem 19, Chapter VII (Section 
89, page 192), since w~ are supposing that AM = 0, that a,-^ 
= a# 2 , for i, j = 1, 2, 3; therefore a// = Va^y, so that we 
may write _ _ _ 

A = -(duVan a 24 V / o: 2 2 # 3 4V / a 33 ) 2 

which proves our theorem, since the negative sign outside the 
parentheses may be introduced under each of the radicals. 

From this theorem we derive an important corollary. For the 
discussion recalls, as might also be derived from Corollary 3 of 
Theorem 19, Chapter VII (Section 89, page 192), that those of the 
principal minors an, a 2 2, 0:33 which do not vanish have the same 
sign as T z (and not all of them can vanish if r s = 2). Hence, if 
Tz > 0, Van, V22, and Vo^ are real and A g 0; while if 7 7 2 < 0, 
these square roots are pure imaginaries or zero (not all zero) and 
A > 0. 



COROLLARY 1. 
site in sign. 



If Au = 0, T 2 4= and A =|= 0, then T, and A are oppo- 



220 CLASSIFICATION OF QUADRIC SURFACES 

The method used to prove Theorem 12 enables us also to prove 
the following important formula: 

012 



COROLLARY 2. The value of the determinant 



Oi2 CI22 2 
13 23 3 

a 6 c 



is 



equal to -(ana 2 4- a 22 6 2 + assc 2 + ^ a 2 3&c 4- 2 a 13 co + 3 a^ab). 

Returning now to the discussion which precedes Theorem 12, we 
can state the following theorem. 

THEOREM 13. If the rank of the matrices a 4 and a 3 are 4 and 2 re- 
spectively, the locus of the equation Q = is an elliptic paraboloid if 
A < 0, and an hyperbolic paraboloid if A > 0. 

100. Classification of Quadric Surfaces the Non-degenerate 
Singular Cases. If r 4 = 3, we can have r 3 = 3, 2 or 1. 
CASE III. n = 3, r 3 = 3. 

From Theorem 14, Chapter VII, we know that in this case the 
surface has a single vertex and from Corollary 3 of this theorem we 
know that it is a proper quadric cone. The reduction of the equa- 
tion in this case is made in exactly the same way as in Case I, 

except that we have now -r = 0, so that the final form of the 

-A 44 

equation is 

/biz" 2 + hy" 2 + k 3 z" 2 = 0. 

The cone is real if and only if the discriminating numbers do not 
all have the same sign; this will always be the case unless the coeffi- 
cients in the discriminating equation present either no variations 
or three variations of sign, that is, unless T 2 > and A^Ti > 0. 
In this case we have therefore the following result : 

THEOREM 14. If the ranks of the matrices a 4 and a 3 are both equal 
to 3, the locus of the equation Q = is an imaginary cone, if T 2 > 
and AnTi > 0; in all other cases the locus will be a real quadric cone. 

Remark. The surface will be a real circular cone if and only if 
the discriminating equation has a simple root of one sign and a 
double root of the opposite sign. 

From Theorem 14 we shall derive an important algebraic 
theorem. It is an immediate consequence of Theorem 14 that the 
equation 
q(x, y, z) = ana: 2 + 022^ + a&z 2 + 2 a^gz -r 2 a 3i zx + 2 a^xy = 



THE NON-DEGENERATE SINGULAR CASES 221 

represents a cone if the determinant A 44 = |ay-|, i, j == 1, 2, 3 is 
different from zero; this cone will be imaginary if 7 7 2 = an + a<& 
+ 0:33 > and A&TI = Au(an + 022 + #33) > 0, but in all other 
cases it is real. 

In the first case, the function q(x, y, z) is reducible to the form 
kix 2 + &2?/ 2 + fc 3 z 2 , in which ki, & 2 , and & 3 are different from zero and 
are of like sign; the function q(x y y, z) will be zero ifx = y z = Q 
and it will be of one sign for all other sets of real values of the 
variables, namely, of the sign of its coefficients, which will be the 
sign of AM since AM is equal to k\k^. In the second case the func- 
tion q(x, y t z) is also reducible to the form kix 2 + & 2 y 2 + & 3 2 2 , but 
now the coefficients in this form are not all of the same sign, and 
the function can therefore take negative, positive and zero values 
for different sets of real values of the variables. We introduce now 
the following definitions. 

DEFINITION II. A homogeneous function of degree 2 in 3 variables 
is called a quadratic ternary form.* 

DEFINITION III. A positive (negative) definite form is one which 
takes the value zero only when all the variables vanish and is positive 
(negative) for all other sets of real values of the variables; an indefi- 
nite form is one which can take positive, negative and zero values for 
real values of the variables. 

We can now state the following important algebraic theorem. 

THEOREM 15. The quadratic ternary form q(x 9 y, *) for which the 
determinant An does not vanish is definite if and only if an + "22 
4- ass > and Au(au + 022 -f "33) > 0; it is positive or negative definite 
according as A is positive or negative. 

CASE IV. r 4 = 3, r 3 = 2. 

It follows from Theorem 14, Chapter VII, that the surface has a 
line of centers. We could therefore begin by translating axes to 
one of the centers as origin; but the reduction of the equation is 
accomplished more rapidly if we follow the method used in Case 
II. Rotation of axes to principal directions leads again to equa- 
tion (3) of Section 99 (see page 217); but since now A = and 
since the discriminant is invariant under rotation, we conclude 
from the discussion made in Case II (page 218) that a^ = 0. 
The equation of the surface reduces therefore to the form 
(1) fe/ 2 + fc 3 z' 2 + 2 a 24 y + 2 a 34 Y + a 44 = 0. 

* A homogeneous polynomial of degree 3, 4, . . . , n is called a cubic, quartic, 
. . . , n-ic form; a form in 2 variables is called a binary form, one in 4, 5, 
. . . , n variables is called quaternary, quinary, . . . , w-ary. 



222 



CLASSIFICATION OF QUADRIC SURFACES 



Completing the square on the terms in y f and on the terms in z' 
and translating the origin to an arbitrary point on the line 

y f = -1 y z' = ~- leads to the equation 



where 

y" 



024, 

fcT' 



fez" 2 = 



JL and 



The discriminant of this last equation is 
000 
fc 2 









- 



It will clearly be of rank 2, unless a 44 " =t= 0. But, since n = 3 and 
the rank of the discriminant matrix is invariant under rotation and 
translation of axes (compare Corollary 3 of Theorem 1, Section 93, 
page 201 and Theorem 5, Section 94, page 205), we conclude that 
a 44 " rjz 0. The locus of the equation is therefore a cylindrical sur- 
face ; it will be an hyperbolic cylinder if & 2 and fc 3 are opposite in 
sign, a real elliptic cylinder if & 2 , fc 3 , and a 44 " are of like sign, an 
imaginary cylinder if & 2 and fc 3 arc of like sign, opposite to that of 
a 44 ". 

As in Case II, we see that fc 2 & 3 = T 2 , so that fc 2 and & 3 will have 
the same sign or opposite signs according as TI > or !T 2 < 0; 
in the former case, they will have the sign of T\ & 2 + & 3 . To 
determine whether or not, in case fc 2 and fe are of like sign, their 
sign is the same as that of a 44 ", we consider the sum Z) 3 of the 
three-rowed minors of the discriminant. Since equation (1) was 
obtained from the original equation Q = by rotation of axes, we 
know from the Corollary of Theorem 4 (see Section 94, page 204) 
that Z> 3 ' = Z) 3 . The discriminant of equation (1) is 

0000 

fc 2 ttu/ 

h au' 

Or / 
#24 O 34 O 4 4 

and we see that every three-rowed minor of the matrix b, associ- 
ated with this discriminant, vanishes. We conclude therefore, by 
making use of the theorem stated in Exercise 6, Section 95 (page 



THE NON-DEGENERATE SINGULAR CASES 223 

205), that JD 3 " = DJ = 3 . Now D 3 " = -a 44 "/c 2 /c 3 ; therefore 
au"k 2 k 3 = Z) 3 . This relation enables us to say that if T 2 
= k 2 k 3 > 0, 044" will be opposite in sign to Z) 3 . Since moreover 
the signs of k 2 and k s are the same as that of Ti, we conclude that 
fc 2 , k s , and 044" will have one sign if TiD 3 is negative, but the sign 
of 044" will be opposite to that of k 2 and A; 3 if T\D^ is positive. We 
have therefore reached the following conclusion. 

THEOREM 16. If the ranks of the matrices a 4 and a 3 are equal to 3 
and 2 respectively, the locus of the equation Q = will be a real elliptic 
cylinder if and only if T 2 > and T { D 3 < 0, an imaginary cylinder if 
and only if T 2 > and TiD 3 > 0, an hyperbolic cylinder if and only if 
T 2 <0. 

Remark 1. We observe that, as in Case II, T 2 must be different 
from zero in this case. 

Remark 2. The surface will be a circular cylinder if and only 
if k 2 = k s . 

CASE V. r 4 = 3, r 3 = 1. 

The surface has no center in this case. Both T 2 and AM are 
equal to zero, but TI is different from zero ; for otherwise it would 
follow that r 3 = by means of an argument which is entirely sim- 
ilar to the argument in earlier discussions and which is therefore 
left to the reader. The discriminating equation is now 

= 0. 



Its roots are ki = k 2 = and fc 3 = TI 4= 0. In accordance with 
Theorem 24, Chapter VII (Section 90, page 195), only one prin- 
cipal direction is completely determined, namely, X 3 , /* 3 , v 3 ; the 
other two principal directions are subject only to the condition of 
perpendicularity to this first direction, and to mutual perpendicu- 
larity. We are therefore free to impose one additional condition 
on Xi, MI, PI or on X 2 , M2, v 2 . 

It is easy to show that in this case X 3 : Ma : vz i : 2 : a#, 
i = 1, 2, 3. For, since fc 3 = T\ = an + a 22 + a 33 , we can deter- 
mine X 3 , ju 3 , and v s from any two of the three linear equations 

~ (022 + 33)X 3 + a^Ma + i3^3 = 0, di 2 X 3 (an + a 33 )/z 3 + 
= 0, ai 3 X 3 + 02 3 M3 ~ (an + 022)^3 = 0. 

From the first two of these equations we find 

X 3 : MS : ?3 = 012023 + 013(011 + 033) : 013012 + 023(022 + 
: (0n + 033) (022 + 033) - 0i2 2 . 



224 CLASSIFICATION OF QUADRIC SURFACES 

But r 3 = 1 ; hence a i3 = #i20 23 #is#22 = 0, so that #12023 = #13022. 
Also e*23 #is#i2 #n#23 = 0, so that #13012 ^ #11023. 

And 0:33 = #n#22 #i2 2 = 0. 



(Consequently we find that X 3 : MS ^ = Oia^i : #23? 7 i : #33^1. And 
since TI ^ and r 3 = 1, so that the rows of the matrix 83 are pro- 
portional, we reach the conclusion that X 3 : ^ : v$ = a,-i : #,2 : 0*3, 
i = 1, 2, 3. 

If we rotate axes to the principal directions determined in 
accordance with these methods, the equation Q = will be carried 
over to the form 

/C 3 2' 2 + 2 # 14 V + 2 flju V + 2 34 Y + #44 = 0, 

where, as before, 2 #,- 4 ' = # 4 (X,-, /x/, ?;), i = 1, 2, 3. The matrix 
a 4 x of this reduced equation is 

000 OH' 
000 # 24 ' 
& 3 #34' 

#14' #24' #34' #44 

In virtue of the hypothesis r 4 = 3 and of Theorem 5 (Section 94, 
page 205), the rank of this matrix must be 3; since the matrix is 
obviously singular, it must contain, on the basis of the Corollary 
of Theorem 6, Chapter II (Section 26, page 44), at least one non- 
vanishing three-rowed principal minor. It should be easy to see 
that the only three-rowed principal minors of this matrix which do 
not vanish identically are those formed from 1st, 3rd, and 4th rows 
and columns, and from the 2nd, 3rd, and 4th rows and columns; 
also that the values of these are A; 3 #u' 2 and & 3 # 24 /2 . It follows 
that at least one of the numbers #i/ and #2/ must be different from 
zero. And now we make use of the freedom of choice left in the 
determination of either Xi, jui, v\ or X 2 , M2, ^2 to effect a further simpli- 
fication of the equation. 

If we adjoin the condition # 4 (Xi, MI> ^i) = to the condition 
XiX 3 + MiMs + ^1^3 = 0, which is imposed by the condition of per- 
pendicularity of the principal directions, the direction Xi, juj, v\ is 
determined; and then X 2 , /* 2 , vi will also be determined as the di- 
rection perpendicular to the other two. In this manner we secure 
the result that #14' = and therefore also the fact that # 24 ' 4= 0. 
Since X 3 : MS : v* = #a : #12 : ##, i = 1, 2, 3, we can determine 



THE NON-DEGENERATE SINGULAR CASES 225 

Xi, MI, v\ from any one of the 3 systems of two equations eacli given 
by #14X1.+ a 2 4jui + a 34 j>i = together with one of the equations 
iXi + a^Mi + fl^i = 0. Hence Xi, /xi, and v\ are proportional to 
the two-rowed determinants formed from one of the three matrices 

II a * a '' 2 a& II, i = 1, 2, 3. This will always determine these di- 
ll a M a 2 4 a 34 II 

rection cosines, unless every two-rowed minor of the matrix b 
vanished; but in this case the rank of the matrix b would be 1 and 
this is incompatible with the condition r 4 = 3, in view of the ob- 
servation (3) made in the proof of Theorem 14, Chapter VII (see 
page 178). We conclude therefore that the principal directions 
can in this case be so determined that a u ' = and 024' 4 1 0. The 
equation of the surface thus takes the form 

fc 3 z' 2 + 2 a* V + 2 au'z' + a 44 = 0. 

If we complete the square on the terms in z, this equation finally 
reduces to 

/C 3 Z //2 = -2024V' 

/*> / 

i // in i i ^44 ^34 " // / i ^34 

where x" = x', y" = *+^,- j^r , *' = *+- 

The locus of this equation is a parabolic cylinder. We may there- 
fore state the following conclusion. 

THEOREM 17. If the ranks of the matrices a 4 and a 3 are 3 and 1 re- 
spectively, the locus of the equation Q = is a parabolic cylinder. 

Remark 1. It should be obvious that we might equally well 
have determined the direction cosines \2, ^2, ^2 in such a way that 
a 24 ' = and au 3r 0. In this case the final equation would be- 
come fez" 2 = 2 ai 4 V, whose locus is also a parabolic cylinder. 
Indeed this change merely amounts to an interchange of the X"- 
and F"-axes. 

Remark 2. It follows from Corollary 3 of Theorem 8 (Section 
96, page 209) that in the case just treated the function q(x, y y z) 
is the square of a linear homogeneous function of x, y, z with real 
or complex coefficients. This observation is frequently useful for 
recognizing whether or not the equation Q = represents a 
parabolic cylinder. 

Example. 

To analyze the equation 

4 x 2 + 7/ 2 + 4 z 2 - 4 xy - 4 yz + 8 zx + 2 x - 4 y + 3 z + 1 =0, 



226 



CLASSIFICATION OF QUADRIC SURFACES 



we set up the matrices 84 



4 -2 

-2 1 

4 -2 

1 -2 



4 1 

-2 -2 

4 I 

1 1 



and a 3 = 



4-24 
-2 1 -2 

4-24 



It is obvious that r$ 1; hence r 4 < 4; and since the three-rowed minor in 

1 -2 -2 



the lower right-hand corner of 4, namely, the determinant 



-2 

-2 



has the value ^/, r 3 = 3. 

In accordance with Theorem 17, we conclude therefore that the locus of the 
equation is a parabolic cylinder. This settles the question as to the type of 
surface represented by the equation. We proceed now to determine its po- 
sition with reference to the given system of coordinates, partly in order to 
exemplify and to verify the method used in the discussion of Case V, and 
partly in illustration of the remark made in the footnote on page 214. 

From the matrix a 3 we conclude furthermore that TI = 9 and we verify that 
T 2 = 0. The discriminating equation is therefore fc 3 9 k 2 0, so that we 
may take k\ k^ = and /c 3 = 9. To determine X 3 , pi 3 , and * 3 we have the 
equations 

5 Xs 2 us -f" 4 *s = 0, 2 Xs 8 jus 2 *3 = 0, 4 X 3 2 ^3 5 *3 = 0. 

From any two of these three equations we obtain X 3 : Ai 3 : * 3 = 2 : 1 : 2, a 
result which was predictable from the discussion in the first part of Case V. 
Since ki = k 2 = 0, the conditions for X!, m, v\ (and also those for X 2 , ju 2 , z/ 2 ) 
reduce to the single equation 2 Xi /ui -f- 2 v\ = 0, which expresses the con- 
dition of perpendicularity to the direction X 3 , /* 3 , *> 3 . To this condition we 
adjoin the condition q*(\i, AH, *>i) = 2 X t 4 m 4- 3 *i = 0. From these two 
conditions we find then that Xi : /*i : PI = 5 : 2 : 6. For X 2 , /u 2 , * 2 we have 
now the conditions 

2 X 2 - ^2 + 2 v-i = and 5 X 2 - 2 & - 6 *> 2 = 

which express the condition of perpendicularity to the two directions already 
determined; from them we find that X 2 : ^2 : v 2 = 10 : 22 : 1. 

The rotation of axes to principal directions is therefore based on the fol- 
lowing table (compare Section 63): 





X 


Y 


Z 


X' 


5 

V65 


-2 
V65 


-G 
V65 


}n 


10 


22 


1 




3V65 


3V65 


3\/65 


Z' 


I 


-t 


2 
3 



THE DEGENERATE CASES 227 

The equations of transformation are therefore 

= 5x f 10 y' 2 z' __ 2^ 22 yV _ ^ 6s' 

X V65 + 3V65 ~T' y ~ "" V65 3V65 3' * ~~" VS5 



, 0' , 2 z' 

3V65 T- 



Hence* 



The equation of the given surface with respect to the rotated axes may there- 
fore be successively transformed as follows: 



81V65/ 



27 



This is the equation of the parabolic cylinder with respect to a frame of refer- 
ence obtained from the original frame by first rotating the axes in accordance 
with the table indicated above and then translating the rotated axes to a new 
origin whose coordinates with respect to the rotated axes are x' 0, y' = 

O f- 

^ 81 VE^' z> ~ ~~ 27' ^ e P om * determined by these coordinates is the 
vertex of a directrix parabola on the cylinder. 

101. Classification of Quadric Surfaces the Degenerate 
Cases. There remain to be considered the cases r 4 = 2, r 3 = 2; 
r 4 = 2, r 3 = 1, and r 4 = r 3 = 1. These cases have already been 
discussed in Section 96 (see Theorem 7, page 208 and Theorem 8, 
page 209) and the results stated there completely settle the prob- 
lem of classification for this case. It will however be instructive 
to derive these results also by means of the methods of reduction 
which were used in Sections 99 and 100. 
CASE VI. r 4 = r 3 = 2. 

The discriminating equation has the form fc 3 Tifc 2 + T z k = 0, 
where T 2 4= 0. As in Case IV, rotation of axes to principal direc- 
tions leads the equation Q = over into the equation 

fey' 2 + fe' 2 + 2 otiV + 2 a 34 'z' + a* = 0. 
* Compare Remark 2 following Theorem 17, page 225. 



228 CLASSIFICATION OF QUADRIC SURFACES 

Completing the square on the terms in y r and z f and translating 
axes, we obtain the equation 

W 2 + fez" 2 = 044", 



whore a 44 " = -~ h ~r -- a ^> an( l the now or igi n ^ an Y point on 

KZ A'a 

,i i. / #24 / ^34 

the line y ' = -- r- , z = -- 7 . 
fc 2 fcs 

It should now be easy to show that, since r 4 = 2 and since the 
rank of the discriminant is invariant under rotation and transla- 
tion of axes, a 44 " = 0. The final equation is therefore 

W 2 + fez" 2 - 0; 

and this equation represents a pair of intersecting planes, real if 
T'2 = A; 2 A; 3 is negative, imaginary if T% is positive; this is the result 
stated in Theorem 7, page 208. 

CASE VII. 7-4 = 2, r 3 = 1. 

The discriminating equation has the same form as in Case V 
and rotation of axes to principal directions leads again to the 
equation 

/c 3 z' 2 + 2 a 14 V + 2 awV + 2 a 34 Y + a 44 = 0. 

The argument used in the discussion of Case V (see page 224) 
shows that, since now r 4 = 2, a^ = a 34 ' = 0. Completing the 
square and translating the axes reduces this equation to the form 

/c 3 z" 2 + 044" = 0, 

where rj 44 " = r/ 44 -- ^- and where the new origin is any point on the 

/fy 

plane z f = --. The sum /V' of the two-rowed principal minors 

A3 

of the discriminant of this last equation is clearly equal to fcsa 44 " 
and this is the only two-rowed minor of the discriminant which 
does not vanish identically. Since r 4 = 2, this can not vanish and 
therefore a 44 " 4= 0. Moreover, an argument similar to the one 
used in the discussion of Case IV (see page 222) shows that the 
sum Z) 2 " for the final reduced equation is the same as the sum D 2 
for the original equation Q = 0. (The details of this argument 
are left to the reader.) Therefore fc 3 a 44 " = D 2 , so that fc 3 and a 44 " 
will be of like or of unlike signs according as D 2 is positive or nega- 
tive. We conclude therefore that the locus of the equation Q = 



SUMMARY AND GEOMETRIC CHARACTERIZATION 229 

is, in this case, a pair of parallel planes which are real or imaginary, 
according as D 2 is negative or positive; this result is stated in the 
first part of Theorem 8 (see page 209). 
CASE VIII. n = 1, r 3 = 1. 

In this case rotation of axes to principal directions and transla- 
tion of axes, as in Case VII, leads to the final equation 

fc 3 z" 2 = 0, 

which represents a pair of coincident planes, in accord with the 
last part of Theorem 8. 

From this discussion we derive some further algebraic theorems, 
which complement Theorem 15 (see page 221). 

If the rank of the matrix as is 2, the equation q(x, y, z) = 
represents a pair of intersecting planes, which are real or imaginary 
according as T% is negative or positive. In the latter case the 
function q(x, y, z) is reducible to the form fcix 2 + & 2 i/ 2 , in which 
fci and fc 2 have like sign, namely, the sign of TI, which is equal to 
ki + & 2 ; the function is therefore a definite quadratic ternary 
form, positive definite or negative definite according as T\ > or 
< 0. If T 2 is negative, the function q(x, y, z) is reducible to 
kix 2 + & 2 2/ 2 , and &i and & 2 will be opposite in sign; in this case the 
function is an indefinite form. 

If the rank of the matrix a 3 is 1, the equation q(x, y, z) = 
represents a pair of coincident planes; the function q(x, y, z) is 
therefore reducible to the form TVr 2 , which is a definite form, posi- 
tive or negative, according as TI > or < 0. 

We have therefore the following extension of Theorem 15. 

THEOREM 18. The quadratic ternary form q(x, y, s) for which the 
rank of the matrix a 3 is 2 is definite if and only if T 2 > 0, positive defi- 
nite or negative definite, according as 7\ is positive or negative; if the 
rank of the matrix a 3 is 1, q(x, y, s) is a definite form, positive definite 
or negative definite according as 7\ is positive or negative. 

102. The Classification of Quadric Surfaces Summary and 
Geometric Characterization. The results which have been ob- 
tained in Sections 99, 100, 101, in as far as they relate to the 
classification of quadric surfaces, are summarized in the following 
table, which specifies the type of surface represented by the equa- 
tion Q(x f y,z) =0 in terms of the invariants of this equation. We 
indicate once more the meaning of each of the symbols used in the 
table. 



230 



CLASSIFICATION OF QUADRIC SURFACES 



A = K'UW = 1,2,3,4; 
^44 = \aij\,i,j = 1, 2, 3; 

4 4 

A = 



u 

= 2 



r 4 = rank of matrix of A; r 3 = rank of matrix of AM. 







, 3 




\ r * 


4 


Singular 


2 1 


rsN, 


Non-singular quadrics 


non-degenerate 


Degenerate quadrics 






quadrics 






A>0 


A<0 










Imagi- 
T 2 >0; nary 


Ellip, 


Imaginary Cone 






3 


A 44 Ti>0 ellip- 
soid 


soid 




Impossible 


Impos- 
sible 




T 2 >0; Hyper- 


Hyper- 










^44^ ^0 boloid 
or of one 


boloid 
of two 


Real Cone 








7^0 sheet 


sheets 
















Imagi- 










Hyperbolic 


Ellip- - 
tic 


T*D\ 


nary 
elliptic 




Imagi- 












cylin- 




nary 












der 


T 2 >0 


inter- 
















secting 




2 


paraboloid 


parabo- 
loid 


f;& 


Ellip- 
tic cyl- 
inder 




planes 


Impos- 
sible 








7 1 2 <0 


Hyper- 
bolic 
cylin- 
der 


n<o 


Inter- 
secting 
plane* 












Imagi- 




1 


Impossible 


Parabolic 
cylinder 


A>0 


nary 
parallel 
planes 


Coin- 
cident 
planes 




Paral- 








D 2 <0 


lel 












planes 





SUMMARY AND GEOMETRIC CHARACTERIZATION 231 



In Theorem 16, Chapter VII (see Section 87, page 185) we proved 
that the quadric surfaces for which r 4 = 4 and r 3 = 3 have a single 
proper asymptotic cone; and that the surfaces for which r 4 = 3 
and 7*3 = 2 have a pair of asymptotic planes. In either case the 
asymptotic quadric of the surface Q(x, y, z) = is given by the 
equation Q(x, y t z) Q(a, 0, 7) = 0, where <*, 0, 7 are the coor- 
dinates of a center of the surface. Since the equations of a quadric 
and of its asymptotic cone differ therefore only in the constant term, 
the invariants AM, jT 2 , and TI, which depend on the coefficients of 
the second degree terms only, are the same for the two surfaces. 
It is clear then from the above table that the asymptotic cone of 
the ellipsoid is imaginary, whereas that of the hyperboloids is real ; 
also that the asymptotic planes are real for the hyperbolic cylinder 
and imaginary for the elliptic cylinder. 

If these observations are combined with the results of Sections 
84 and 85 we obtain the complete geometric characterization of 
the real quadric surfaces indicated in the table on page 232. 

Examples. 

1. To determine the character of the surface represented by the equation 

5 x 2 + 5 y 2 + 8 z 2 + 8 yz + 8 zx - 2 xy + 12 x - 12 y + 6 = 
we set up the matrices 84 and as. We find 



a 4 = 



5 -1 
-1 5 

4 4 

6 -6 



4 
4 
8 




G 

-6 

6 



arid 3 = 



-1 
5 
4 



The determinants of these matrices are both found to vanish because in each 
of them the third row is equal to the sum of the first and second rows. The 
third order principal minor of a 4 which is formed from its last three rows and 
columns, and the two-rowed principal minor of a s in its upper left-hand 
corner are both found to be different from zero. We conclude therefore that 
r 4 = 3 and r 8 = 2, and that the surface is a cylinder. Its axis, that is, its line 
of centers, is determined by any two of the system of three linear equations 
whose augmented matrix furnishes the first three rows of a 4 ; we can take for it 
the equations 5 z y + 4 2 + 6 =0 and x + y +2 z = 0. 
Moreover, 



and 

5-14 
Z> 3 = -1 54 

4 48 
Hence, T 2 > and 



lsr _ 5-11,154641 
8-18, T, - _j s | + | 4 8 + 4 8 | 




5-16 




546 




54-6 


-(- 


-1 5 -6 


+ 


480 


-j- 


48 




6-66 




606 




-60 6 



= 72; 



= -432; 



< 0, so that the surface is an elliptic cylinder. 



232 



CLASSIFICATION OF QUADRIC SURFACES 





Centers 


Lines on 
surface 


Asymptotic 
cone 


Ellipsoid 


Single proper 
center 


No lines 


Imaginary, 
proper 


Hyperboloid of 
one sheet 


Single proper 
center 


Two lines through 
every point 


Real, proper 


Hyperboloid of 
two sheets 


Single proper 
center 


No lines 


Real, proper 


Hyperbolic 
paraboloid 


No center 


Two lines through 
every point 




Elliptic 
paraboloid 


No center 


No lines 




Cone 


Single vertex 






Elliptic cylinder 


Line of proper 
centers 


Two coincident 
lines through 
every point 


Imaginary, 
degenerate 


Hyperbolic 
cylinder 


Line of proper 
centers 


Two coincident 
lines through 
every point 


Real, degenerate 


Parabolic 
cylinder 


No center 


Two coincident 
lines through 
every point 




Intersecting 
planes 


Line of vertices 






Parallel planes 


Plane of proper 
centers 






Coincident 
planes 


Plane of vertices 







The discriminating equation is k 3 18 k z + 72 k = 0; the discriminating 
numbers are therefore 0, 6, 12. We put ki 0, k z = 6, k 3 = 12. We know 
from the general discussion that rotation to principal directions will reduce the 
equation to the form 6 y' 2 + 12 z' 2 + 2 a^'y' + 2 a 84 'z' +6=0. To verify 
this fact, we proceed to determine the principal directions. 

From ki = 0, we find 5 Xi m + 4 v\ = and Xi + 5 p\ + 4 ^ =0, 
so that Xi : MI : f ! = 1 : 1 : 1 ; from fc 2 = 6, we find X 2 ^2 + 4 >> 2 = 
and 2 X 2 + 2 /i2 + ^2 = 0, so that X 2 : /z 2 : ? 2 = 1 : 1 : 0; from k s = 12, we 
find -7 X 3 - MS + 4 V9 = and ~X 3 - 7 ^3 + 4 v z = 0, so that X 3 : v* : t>s = 
1:1:2. The equations for the rotation of axes to principal directions are 

f' 11' y' T r ii f y f '*' 9 *' 

thereforex=4-+4-+4=,2/ = 4--4- + -^, 2 = -4= + ^. 



SUMMARY AND GEOMETRIC CHARACTERIZATION 233 



If these expressions are substituted for x, y, and z in the original equation 
of the surface, it becomes 

6 2/' 2 + 12 z' 2 + 12 t/V2 +6=0; 

completing the square on the terms in y', we obtain as the final form of the 
equation ?/" 2 + 2 z" 2 = 1, where y" = y' + V2 and z" = z f . The locus of 
this equation is indeed an elliptic cylinder; its axis is parallel to the X "-axis, 
i.e., to the line y" = 0, z" = 0. But the equations of this line may also be 
written in the form y' -f ^/2 = 0, z' = 0, or, with respect to the original frame 
of reference, in the form x ?/ -f 2 = 0, x + y+2z=Q, which is equivalent 
to the form of the equations of the axis of the cylinder given at the end of the 
first paragraph. The directrix of the cylinder is the ellipse y" 2 -f 2 z" 2 = 1, 
x" = 0; the equations of this ellipse with respect to the original axes are 
5 z 2 + 5 y 2 + 8 z 2 + 8 yz + 8 zx - 2 xy + 12 x - 12 y + 6 = 0, x+y-z=Q. 
2. We proceed in a similar manner with the equation 

2 x 2 + 20 ?/ 2 + 18 2 2 - 12 T/Z + 12 xy + 22 x + 6 y - 2 z + 2 = 0. 



We find a 4 = 



2 6 11 

6 20 -6 3 

-6 18 -1 

11 3-1 2 



and 213 = 



6 

20 -6 

-6 18 



r 3 = 2, r 4 = 4; A = 33, 124 < 0. Therefore the surface represented by 
the given equation is an elliptic paraboloid. 

Furthermore, T\ 40 and T 2 364 and the discriminating equation is 
/c 3 - 40 k 2 + 364 k = 0; its roots are 0, 14, and 26. From the discussion of 

Case II (see Section 99, page 218), we know moreover that an' = 4 / ^ 

V * 2 

= \/91. The equation of the surface is therefore reducible to the form 14 y" 2 
+ 26 z" 2 = 2 V9I x". 

This completes the determination of the type of surface represented by 
the equation and also of the numerical data necessary to fix its individuality. 
If we wish to determine its position with respect to the original frame of 
reference, we have to find the principal directions and also the new origin to 
which the axes have been translated. 

From ki = 0, we find Xi : MI : "i = 9 : -3 : -1; from k 2 = 14, follows 
\2 : M2 "2 = 1 2 : 3, and from 3 = 26, we derive X 3 : MS ^3 = 1 : 4 : 3. 
If we base the rotation of axes to principal directions on the table 





X 


Y 


Z 


X' 


9 
V91 


-3 
V9l 


-1 
Vol 


Y' 


1 


2 

Vl4 


3 

vT3 


Vl4 


Z' 


1 
V26 


4 

V26 


-3 

V26 



234 CLASSIFICATION OF QUADRIC SURFACES 

the equation is transformed into 14 y'* + 26 2'* + 2 yUl x' + 2 VI? y' -f 2 V26 2' 

+2 = 0: translation of axes to the point x' = 0, y' = 7^* P' = :=. 

__ V14 V26 

leads to the final equation 14 ?/" 2 + 26 2" 2 = -2V91 z". If the direction 
cosines of the X'-axis are changed in sign, that is, if its direction is reversed, 
the right-hand side of the final equation of the surface would be 2 V^l x". 
The surface extends indefinitely on the negative side of the plane x" = 0, 
that is, in terms of the original system of coordinates, on that side of the plane 

Oaj S?/ 2 = which is determined by the direction cosines X = -7=, 

VQl 
g j j j 

fji = T=i, v = -7=.; it has the point x' =0, ?/' = -7=., 2' = =, that 
V91 VST Vl4 V26 

10 27 9 

is, the point x = ~qT>2/ =: ~~Q7 2:= ~~qT m common with this plane. 

103. Exercises. 

Determine the type of quadric surface which is represented by each of the 
following equations and set up the reduced form of these equations; dis- 
cuss their position in space in those cases in which the numerical work in- 
volved does not become too laborious: 

1. x 2 -f 4 y 2 + 9 2 2 + 4 xy -f 6 xz + 12 yz - x + 2 y + 5 z = 0. 

2. x 2 -I- z 2 -f 2 xy -f 2 a* - 2 7/z - 2 x + 4 ?/ - 4 = 0. 

3. x 2 4- 2/ 2 -I- 2 2 + 2/2 + zx + a^ -f 2 x + 2 T/ + 2 2 -f- 2 = 0. 

4. 4 z 2 -f t/ 2 + 2 2 - 4 xy + 6 z + 8 = 0. 

6. x* + y* + z* + yz -f zx + xy + x -f- y + z + 1 = 0. 

6. z 2 4- 2/ 2 4- 2 z 2 - 4 X2 + 2 a;?/ + 1 = 0. 

7. 2 2 - xy -h a; = 0. 

8. 5z 2 -y 2 +z 2 + 6zz-Mo;?/ + 2z4-22/ + 23 = 0. 

9. x 2 + y* + z* xy yz - y = 0. 

10. 5x 2 + 13 2/ 2 -1-23 * 2 -f 36r/2 -f 22 a? -f 16^?/ + 10s + 167/4-222+5 = 0. 

11. s 2 + 5 ?y 2 + 9 2 2 -f 4 x?/ + 6 yz - z - 3 x = 0. 

12. 4 2/ 2 + 4 2 2 + 4 2/2 - 2 re - 14 ?/ - 22 z + 33 - 0. 

13. 22/ 2 4-42z-f2 :r -42/ + 62-h5 = 0. 

14. x 2 + 2/ 2 + 2^ - 6 xy -f 2 zx - 6 7/2 - 6 x - 2 1/ - 6 z + 1 = 0. 
16. 3 x 2 -f 3 y 2 + 3 z 2 + 2 ?/2 + 2 2z + 2 x?/ -f 1 = 0. 

16. x 2 + z?/ -f 2/2 + zx - 3 z - 2 ?y - 2 - 3 = 0. 

17. x 2 + 3 2/2 - 2 2 -f 2 x = 0. 

18. (x - 2 y + 2) 2 -f- 4 x - 8 y + 4 2 + 3 = 0. 

19. 36z 2 H-47/ 2 + z 2 -4i/2- 12ac + 24xy + 4 + 16y -262 -3 = 0. 

20. z 2 -f 4 ?/ 2 + 2 2 - 4 1/2 + 2 20; - 4 xy - 2 * + 4 y - 2 2 - 3 = 0. 

21. 2x 2 - 22/2 -f 2zz-2;n/-z -2y + 82 - 2 0. 

22. Determine the condition which must be satisfied by the discriminating 
numbers of a quadric surface in order that it may be a surface of revolution; 
also the conditions which will insure that the surface is a sphere. 

23. Set up the conditions which the invariants of the surface Q(x t y, z) =0 
must satisfy in order that the surface may be a sphere. 



CHAPTER IX 

QUADRIC SURFACES, SPECIAL PROPERTIES 
AND METHODS 

In this chapter we shall discuss some properties and methods 
which are concerned with one or another of the classes of quadric 
surfaces, with which we have become acquainted. 

104. The Reguli on the Hyperboloid of One Sheet. We have 
seen in Sections 84 and 102 that the hyperboloid of one sheet and 
the hyperbolic paraboloid are the only real quadric surfaces 
through every point of which there pass two real lines which lie 
entirely on the surface. The general method developed in Section 
84, by means of which the existence of these lines was demonstrated, 
is not very convenient for actually determining the rulings through 
a particular point on a given surface; we shall therefore in this 
section take up a special method, for the hyperboloid of one sheet, 
for the solution of this problem. And, having shown in the pre- 
ceding chapter that every quadric surface can be represented, 
with suitable choice of the frame of reference, by an equation char- 
acteristic of the type to which the surface belongs, we shall hence- 
forth make use of these standard forms of the equations of the 
quadric surfaces. 

The standard form of the equation of the hyperboloid of one 
sheet is 



It may be written in the form 



It should be easy to see from this that the line which is determined 
by the pair of equations 

-O - 

235 



236 QUADRIC SURFACES 

in which p\ and p 2 are arbitrary real numbers which do not both 
vanish, lies entirely on the surface. Since the value of the ratio 
PI : pi may therefore be chosen arbitrarily, we have here a single 
infinitude of lines, or a one-parameter family of lines on the surface. 
This family of lines is of the type known in Projective Geometry as 
a regulus.* We shall use this term for convenience of reference 
without entering further into its definition; and we shall refer to 
the regulus whose lines are determined by equations (1) as the 
p-regulus of the surface. 

It is moreover clear that we can obtain a second regulus of the 
surface in the form 



we shall call this the q-regulus, it being again understood that q\ 
and #2 are arbitrary real numbers which do not both vanish. 

A particular line I of the p-regulus is known as soon as the ratio 
Pi : p 2 is given, and we shall designate this line by the symbol 
l(pi> 2); similarly the symbol m(qi 9 g 2 ) will be used to designate 
a line of the g-regulus. To determine the particular line of each 
regulus which passes through a given point A (a, 0, 7) on the sur- 
face, we substitute the coordinates of this point in one of the 
equations (1) to determine pi : pz, and also in one of the equations 
(2) to determine qi : q 2 . 

Example, 

To determine the lines through the point A (4, 2, 3) on the surface r- 

-f X ~ <7 = *' we wr ^ e ^ ne equations of the reguli; for the particular sur- 
face under consideration, these are: 



and 



Substituting the coordinates of A for ac, y, z in these equations gives p\ X 
= p* X 2, p2 X 2 = pi X 0; qi X = q 2 X 0, ^ X 2 = qi X 2. Therefore 
Pz and p\ is arbitrary and q\ = qi. 

* For a definition and treatment of the regulus, see, for example, Veblen 
and Young, Projective Geometry, Vol. 1, pages 298-304. 



REGULI ON THE HYPERBOLOID OF ONE SHEET 237 
Hence the required lines are given by the following pairs of equations: 

105. Reguli on the Hyperboloid of One Sheet, continued. A 
number of questions concerning the reguli on the hyperboloid of 
one sheet, which suggest themselves naturally, will now be 
considered: 

(1) Will the determination of p\ : p 2 and of 51 : g 2 always be 
possible in one and only one way? 

(2) Do we get two distinct lines through every point on the 
surface? 

(3) Do two lines of one regulus ever lie in the same plane? 

(4) Will every line of the p-regulus be coplanar with every 
line of the g-regulus? 

(5) What is the relative position of the plane determined by 
the two rulings which pass through a point A on the 
surface, and the tangent plane to the surface at PI 

These questions, except the first, fall within the scope of Chapter 
IV and can be answered by the methods developed there. We 
shall take up these questions in some detail, because they furnish 
an opportunity to illustrate the application of these methods. 

(1) Will the determination of p\ : p% and of q\ : q 2 always be pos- 
sible uniquely? From an equation of the form ap\ bp 2 the ratio 
pi : pz is always uniquely determinable, unless a = 6 = 0. Hence 
if the point A (a, /?, 7) lies on the hyperboloid of one sheet repre- 
sented by the equation ^ ~ + ~i = 1 an d tf r > 1 > 

a o c (to c 

- + - , 1 + - are not all zero, then at least one and at most two de- 
a b' c 

terminations of each of the ratios pi : p% and qi : q% are possible. 
But if these four expressions all vanished, it would follow that 
a = & = 7 = 0, which is not a point on the surface. And if the 
two equations (1) of Section 104 gave rise to two different values 

of T>I : p 2 > it would follow that 1 - : r : fc- + ?:l+~> 

^ ^ c a b a b c 

cf B* *y 2 

and hence that 2 ro^l -^so that the point A (a, 0, 7) 



238 



QUADRIC SURFACES 



could not be on the surface. The same conclusion holds for the 
equations (2). The question is therefore to be answered affirma- 
tively; consequently the method of the preceding section is 
always effective to determine the lines whose existence was 
proved in Section 84. 

(2) Will the two lines through A(a, )8, 7) determined by the method 
of Section 104 olways be distinct? 

If these lines are not distinct, then the four planes represented 
by the equations (1) and (2) of Section 104 for the particular 
values of pi : p 2 and qi : # 2 determined as in (1), must have a line 
in common. According to Theorem 23, Chapter IV (Section 54, 
page 101) this will happen if and only if the ranks of the coefficient 
matrix and the augmented matrix of their equations are both 
equal to 2. 

Pi _pi P2 



The coefficient matrix of these equations is 



Pt 
b 



1 2? ^ 

\ b 

its rank is not affected by elementary transformations (compare 
Definition XIV and the Corollary of Theorem 14, Chapter I, 
Section 10, pages 18, 19), so that we may multiply its 1st, 2nd, 
and 3rd columns by a, fe, and c respectively, and add the 1st 

Pi p 2 

column to the 2nd. This leads to the matrix * ^ l 

qi q 2 

02 2 q 2 q\ 

It is found that the values of the third order determinants ob- 
tained from this matrix by omitting the 4th, 3rd, 2nd, and 1st 
rows respectively are equal to 2 p 2 (piq 2 + P*qi), 2 pi(p\q 2 + p^qi), 
2q 2 (piq 2 + P2<?i) and 2qi(piq 2 + p 2 qi). Therefore, since, in 
accordance with (1), pi, p 2j q\, and q 2 never vanish simultaneously, 
the rank of the coefficient matrix can be 2 only if piq 2 + p 2 qi = 0. 
Let us now consider the matrix formed from the 1st, 3rd, and 
4th columns of the augmented matrix; after applying to it similar 



REGULI ON THE HYPERBOLOID OF ONE SHEET 239 



transformations as to the coefficient matrix we have to consider 
Pi 2 p 2 



the matrix 



P2 








pi 



the values of its third order determi- 



nants are found to be 
(pzqz piQi) and 2#i(p 2 g 2 Piqi)- Hence if the rank of the 
augmented matrix is also 2, we must have p 2 # 2 Piqi = as well 
as PI& + P2<7i = 0. If we look upon these equations as linear 
homogeneous equations in q z and q\, we find, by using Theorem 2, 
Chapter II (Section 22, page 38), that either q\ = g 2 = or 
Pi 2 + P2 2 = 0, that is pi = p 2 = 0. Neither of these cases can 
arise, in virtue of the discussion in (1). The answer to our second 
question is therefore also affirmative. 

(3) Do two lines of one regulus ever lie in one plane? 

Let us consider the lines I(p i9 p 2 ) and l'(pi, p 2 ') of the p-regulus. 
They are given by the two pairs of equations 



and 



where pi : p 2 = Pi Pa'. 

The value of the determinant of the augmented matrix of these 
equations is 

J)2 



abc 



Pi -Pi 

P2 P2 

ft' -ft' 

ft' ft' 



-Pi -Pi 
ft' -ft' 
-Pi' -Pi' 



Pi 

p 2 2 p 2 
Pi' 



P2 

-Pi -5 







The evaluation of this determinant is accomplished most con- 
veniently by means of the Laplace development of the 1st and 3rd 
columns (compare Theorem 15, Chapter I, Section 12, page 23); 

Pi ; 



we find then that its value is 



Pi 



X 



2p 2 ' -2 Pi' 



/ Pzpi) 2 t which is different from zero, in virtue of the 
hypothesis p\ : p 2 4= Pi' p*- It follows from this, on account of 
Theorem 22, Chapter IV, (Section 54, page 101), that the two lines 



240 QUADRIC SURFACES 

of the p-regulus can not have a point in common. Can they be 
parallel? 

There should be no difficulty in showing that the direction cosines 
of the lines l(pi, p 2 ) and V(pi, 7)2') are given by the proportions 

X : /i : v = a(pi 2 - p 2 2 ) : b(pi 2 + p 2 2 ) : 2 cpip 2 , 
and 

X' : M ' : i/ = a(p x ' 2 - p 2 ' 2 ) : &(/>i' 2 + p 2 ' 2 ) : 2 cpi V- 

Since pi, p 2 , p/ and p 2 ' are all real, both pi 2 + P2 2 and p/ 2 + p 2 ' 2 
must be positive. Therefore, if the lines are to be parallel, there 
must exist a positive factor of proportionality p 2 , such that 

Pi 2 -p 2 2 = p 2 (pi' 2 -p 2 ' 2 ), Pi 2 +p 2 2 = p 2 (pi' 2 +p2 /2 ) and 



From the first two of these equations we conclude, by addition 
and subtraction, that pr = p 2 p/ 2 and p 2 2 = p 2 p 2 ' 2 , and hence 
that pi = zb ppi and p 2 = dbpp/. If opposite signs were used in 
these two equations, it would follow that pip 2 = P 2 p/p 2 '> which 
is in conflict with the third of the above equations. Therefore 
the lines can have the same direction cosines only if pi : p 2 = 
Pi - pit fchat i s > ^ they coincide. 

Our discussion has therefore brought us to the conclusion that 
no two lines of the p-regulus can lie in the same plane. 

The reader is urged to carry through the same argument for 
the #-regulus. When he has done this, he will have completed 
the proof that the answer to the third question is in the negative. 

(4) Will every line of the p-regulus be coplanar with every line 
of the q-regulus? 

The discussion of this question will be based on Theorem 25, 
Chapter IV (Section 54, page 104). If Z(pi, p 2 ) is an arbitrary 
line of the p-regulus and m(gi, (? 2 ) an arbitrary line of the g-regulus, 
these lines will be coplanar unless the determinant 

X Xi a. i 

M Ml 0-01 
v 1/1 7 71 

is different from zero; here X, /*, v and Xi, /n, vi are the direction 
cosines of the lines I and m respectively, and (a, 0, 7) and (], 
0i, 71) are arbitrary points on these lines. We have already 
seen that X : /* : v = a(pi 2 p 2 2 ) : 6(pi 2 + p 2 2 ) : 2 cpip 2 (see (3) 



REGULI ON THE HYPERBOLOID OF ONE SHEET 241 

above) ; in similar manner we find that Xi : /*i : v\ = a(qi 2 
<?2 2 ) : Ktfi 2 + <? 2 2 ) : 2 c^ 2 . It remains therefore to investigate 
whether or not the determinant 

(pi 2 ~ p 2 2 ) a(?i 2 - ? 2 2 ) a- o 



has a value which is different from zero, when for a, ft 7 and i, ft, 71 
are substituted sets of numbers which satisfy the pairs of equations 



<Ml + )> ft/- + r ) = q\( 1 -- ) respectively. It turns out that 

this determinant vanishes for every such choice of a, ft 7 and 
i, ft, 7iJ the details of the proof of this statement are given in 
Appendix, VII (page 299). Hence we conclude that every line of 
the p-regulus is coplanar with every line of the g-regulus. 

(5) What is the relative position of the plane determined by the two 
rulings which pass through a point on the surface and the tangent 
plane to the surface at this point? 

The equation of the tangent plane to the surface at a point 
A (a, j8, 7) on the surface is 

OiX Vt 72 



(compare Theorem 4, Chapter VII, Section 81, page 162). This 
plane and the line of the p-regulus through A have at least the 
point A in common; the line will therefore lie in the plane or meet 
it in a single point, according as 

- P2 2 ) X - b(pi 2 + ?2 2 ) X 2 + 2 cp lP t X 



is equal to or different from zero (compare Theorem 21, Chapter 
IV, Section 52, page 99). This expression is equal to pi 2 f - rj 

7>2 2 ( ~ + v ) + 2 pip 2 - ; and this reduces, by virtue of the equa- 
\ct o / c 

tions of the lines of the p-regulus, to p\pji 1 - J p*pi( 1 + -J 



242 QUADRIC SURFACES 

+ 2pip 2 - = 0. Consequently, the line of the p-regulus lies in 
c 

the tangent plane; the reader should have no difficulty in proving 
that the line of the g-regulus which passes through a given point 
on the surface also lies in the tangent plane to the surface at that 
point. These conclusions could also have been reached by a 
geometric discussion, namely, by observing that any line which 
connects A with a point on the plane determined by the lines 
/(pi, pz) and m(qi, #2) meets the surface in two points which co- 
incide at A, and is therefore tangent to the surface at A (compare 
Definitions V and VI, Chapter VI, Section 77, pages 154, 155). 
We summarize the results obtained in this section in a theorem. 

THEOREM 1. A hyperboloid of one sheet contains two one-parameter 
families of lines. Through every point of the surface passes one and 
only one line of each family. No two lines of the same family are 
coplanar; every line of one family is coplanar with every line of 
the other family. The plane determined by the two lines which pass 
through an arbitrary point on the surface is coincident with the 
tangent plane to the surface at this point. 

106. The Reguli on the Hyperbolic Paraboloid. The standard 
form to which the equation of an hyperbolic paraboloid can be 
reduced is 

T 2 7/ 2 

__ y _ 9 ~~ 

5 T-T & nz. 
a 2 6 2 

If we write this equation in the form 

_AA + f) = 2n .s, 

a b/ \a b/ ' 

it becomes clear that the surface contains the two one-parameter 
families of lines represented by the following pairs of equations 



Pl 
bers, not both zero; 

g \a ~~ b) = 2 nq *' ^ 2 (a + 1) = qiZf q 
bers, not both zero. 

With reference to these reguli on the hyperbolic paraboloid, the 
same questions arise as were discussed for the hyperboloid of one 



STRAIGHT LINES ON SINGULAR QUADRICS 243 

sheet. The discussion of these questions is left to the reader 
(see Section 108). 

107. the Straight Lines on the Singular, Non-degenerate 
Quadrics. It was proved in Theorem 12, Chapter VII (Section 
84, page 175), that through every point on a non-degenerate singu- 
lar quadric which is not a vertex of the surface, there pass two 
coincident lines which lie entirely on the surface. This class of 
surfaces includes the proper cone and the cylinders. 

If the equation of the proper cone 



is written in the form 



a c a c 

we recognize that the one-parameter families of lines, represented 
by the pairs of linear equations 

(x z\ y 

---)= P2 -f, 

j 

and 



z 



in which pi, p^ qi, and q<* are real numbers and neither pi and p 2 , 
nor qi nor (j 2 vanishes simultaneously, lie entirely on the surface. 
It should be clear however that the lines l(p\> p 2 ) and m(qi 9 # 2 ) 
are identical when qi = pi and g 2 = -p* Consequently, these 
two families of lines are identical; the proper cone contains there- 
fore two coincident reguli. 

A similar argument shows that the elliptic cylinder, the hyper- 
bolic cylinder, and the parabolic cylinder, also each contain two 
coincident reguli. 

108. Exercises. 

1. Determine the equations of the straight lines on the surface -r ~ 

4 y 

- 2z, which pass (a) through the point A(10, 9, 8); (b) through the point 
B(-10, 9, 8); (c) through the point C(10, -9, 8); (rf) through the point 
ZK-10, -9,8). 

x 2 t/ 2 

2. Determine the equations of the rulings of the surface -j- 4- fj- z 2 * 1 



244 QUADRIC SURFACES 

which pass (a) through the point A (4, 3, 2); (6) thro'.gh the point 
(-4, 3, 2); (c) through the point C(4, -3, -2); (d) through the point 
D(-4, -3, -2). 

3. Determine the equations of the straight lines on the cone x 2 y 2 -f z a 
= 0, which pass (a) through the point A (3, 5, 4); (6) through the point 
B(-3, 5, 4); (c) through the point C(3, 5, 4); (d) through the point 
/>(-3,5, -4). 

4. Show that the elliptic cylinder, the hyperbolic cylinder and the parabolic 
cylinder each contain two coincident one-parameter families of lines. 

6. Prove that any two lines which belong to one regulus of the hyperbolic 
paraboloid are skew. 

6. Prove that every line of one regulus of the hyperbolic paraboloid is 
coplanar with every line of the other regulus of that surface. 

7. Prove that the plane tangent to the hyperbolic paraboloid at an arbitrary 
point contains the two rulings of the surface which pass through that point. 

8. Determine the cosine of the angles made by the two lines which pass 
through an arbitrary point of the hyperboloid of one sheet and determine 
the condition under which these two lines are perpendicular. 

9. Discuss the corresponding question for the hyperbolic paraboloid. 

10. Prove that to every point A (a, 0, 7) on the hyperboloid of one sheet 
there corresponds a point A' on the surface such that the line of the p-regulus 
through A is parallel to the line of the ^-regulus through A', and vice versa. 

11. Determine the locus of all points on the hyperboloid of one sheet for 
which the angle between the rulings of the surface which pass through them 
is constant. 

12. Show that it is possible to set up a correspondence between each of the 
reguli of the hyperboloid of one sheet on the one hand, and the regulus of its 
asymptotic cone on the other, such that corresponding lines are parallel. 

109. Circles on Quadric Surfaces, the General Method. In 

Section 66 we discussed the problem of determining the curve of 
intersection of a plane and a surface. In accordance with Corol- 
lary 2 of Theorem 8, Chapter V (Section 66, page 128), a plane 
section of a quadric surface is a curve of degree not higher than the 
second. It is of interest to inquire whether plane sections of 
quadric surfaces can be circles. In answer to this question we 
shall prove in the first place the following theorem. 

THEOREM 2. The sections of a quadric surface made by two parallel 
planes are either both circles or else neither is a circle. 

Proof. The method developed in Section 66 for determining 
the character of a plane section of a surface consisted in rotating 
the axes in such a way as to make one of the new coordinate planes 
parallel to the plane of the section. The rotation required by this 



CIRCLES ON QUADRIC SURFACES 245 

method is the same for the sections of a surface by each of two 
parallel planes \x + py + vz p\ = and \x + ny + vz p 2 = 0. 
Let us suppose that the rotation of axes is made in such manner 
as to make X, /*, v the direction cosines of the new Z-axis; then 
the equations of the two given planes will be, with reference to the 
new axes, z' = p\ and z f = p 2 . Consequently the plane equations 
of the two sections will be obtained when, in the new equation of 
the quadric surface, z' is replaced by pi and p 2 . If the transformed 
equation of the quadric surface is Q(x, y, z) = 0, then the plane 
equations of the sections will be 

2 a i2 xy + a 22 ?/ 2 + 2 (ai 3 pi + a^x + 2 
33 pi 2 + 2 a 34 pi + 44 = 0, 
2 a l2 xy + a 22 y 2 + 2 (a 13 p 2 + a u )z + 2 
2 a 34 p 2 + a 44 = 0. 



Since the condition that the plane locus of either of these equa- 
tions shall represent a circle is that a u = #22 and a i2 = 0, the 
theorem has been established. 

Remark. To determine the circular sections of a quadric sur- 
face, it suffices to consider the planes through some fixed point. 

We could now proceed for the further discussion of our problem, 
to determine the plane equation of the section of the quadric by 
an arbitrary plane through the origin and then impose the con- 
ditions which insure that the locus of this equation is a circle. 
But we have had opportunity to observe before that the most direct 
rhethod is not always the most convenient and that the end we are 
seeking to accomplish is frequently reached in a more elegant and 
more instructive way by a more sophisticated procedure. This 
will be our program in the present case. 

We begin by recalling from Elementary Plane Geometry that, 
if lines Zi, Z 2 , . . . are drawn through a point P in the plane of a 
circle, meeting the circle in pairs of points AI, BI, A 2j J9 2 , . . . , 
then the products PAi PBi, PA 2 PB 2 , . . . are all equal to each 
other. It is true conversely, that if A\B\ 9 A 2 B 2 , ... are chords 
of a conic section which pass through a fixed point P and the 
products PAi PBi, PA 2 PB 2) ... are all equal for any fixed 
position of P in the plane of the conic section, then this conic is 
a circle.* 

* A proof of this converse theorem will be found in Appendix, VIII, page 300. 



246 QUADRIC SURFACES 

Suppose now that the plane through the origin, whose equation is 
(1) ax + by + cz = 

cuts the quadric surface Q in a circle and that the lines Zi, Z 2 , h . . . 
through the origin which lie in this plane cut the surface in the 
pairs of points A\,B\] A^ 5 2 ; . . . . If the direction cosines of 
an arbitrary one of these lines is designated by X, M, v, these di- 
rection cosines satisfy the condition aX + fyu + cv = (compare 
Theorem 21, Chapter IV, Section 52, page 99). And the dis- 
tances from to the points in which the line meets the quadric 
are given, in magnitude and direction, by the roots of the equation 



q(\, n, v)s* + 2 (<z 14 X + a 24 ju + a z ^)s + a 44 = 0, 

(compare Theorem 1, Chapter VII, Section 80, page 160 and re- 
member that in this case a = /3 = 7 = 0); and the product of 

these roots is equal to ^ 44 r It follows that if the section is a 
ff(X, /*, ") 

circle, then /x 44 r must have the same value for all those ad- 

9(*, M, v) 

missible values of X, /x, v for which aX + hp + cv = 0. This 
means that there must exist a number k, which is independent of 
X, M, and v, such that q(\ n, v) = fc, whenever aX + ?>M + cv = 0, 
or again that the quadratic equation <z(X, /z, v) k == is satisfied 
whenever the linear equation a\ + bfj, + cv = is satisfied. 
From this we conclude, first that the quadratic function q(\, p. y v)k 
must be factorable in two linear functions of X, /i, v with real or 
complex coefficients (compare the argument made in the proof 
of Corollary 2 of Theorem 12, Chapter VII, Section 84, page 175); 
and secondly, that aX + bp + cv must be one of the factors. 
Furthermore, the function <ft(X, ju, v) ss j(X, ju, v) k = g(X, /x, v) 
fc(X 2 + M 2 + J' 2 ) is factorable, according to Corollary 3 of 
Theorem 8, Chapter VIII (Section 96, page 209) if and only if 
the rank of its discriminant matrix is less than 3, that is, if and 
only if 



CL\\ k dm #18 

diz #22 "k flsa 
#13 023 flaa k 



0. 



CIRCLES ON QUADRIC SURFACES 247 

But this equation is the discriminating equation A(k) = of the 
surface <3 (compare Sections 88 and 89). If, as before, we desig- 
nate its roots by &i, & 2 , and & 3 , we conclude that if the plane 
ax + by + cz = cuts the quadric Q in a circle, then ax + by + cz 
must be a factor of one of the quadratic functions q(x, y, z) 
MX* + 2/ 2 + z 2 ), i= 1,2 or 3. 

Conversely, if ax + by + cz is a factor of g(z, y, z) 
&*(# 2 + y 2 + z 2 )> for i = 1, 2, or 3, then aX + bp + cv is a factor 
of q(\ /i, v) kt(\ 2 + M 2 + " 2 ), that is, of q(\ p, v) fc,-; conse- 
quently g(X, /i, v) = & 4 * for all admissible values of X, M, v for which 
aX + bfj, + cv = 0. If we take now an arbitrary point P(a, P, 7) 
in the plane ax + by + cz = 0, the lines in the plane through P 
will cut the quadric Q in pairs of points A, /? whose distances 
from P are the roots of the equation Los 2 + 2 Li + L 2 = 0, 
where L = q(\, M, *>) and L 2 = Q(a, ft 7) (see Theorem 1, Chap- 
ter VII, Section 80, page 160); hence the product PA PB == 

Q i?' ^ y ) = Q(a> / >7) for all admissible values of X, M , v for 
g(X, M, v) fc,- 

which a\ + by -\- cv = 0. But this expresses the fact that the 
product of these distances is constant for all lines through 
P, no matter what point P is chosen, and therefore (see footnote 
on page 245) that the plane cuts the quadric in a circle. We have 
therefore established the following theorem. 

THEOREM 3. The necessary and sufficient condition that the plane 
through the origin represented by the equation ax 4- by -f cz = 
shall be a plane of circular section of the quadric surface Q(x, y, s) =0 
is that ax -f- by + cz must be a factor of the homogeneous quadratic 
function q(x, y, *) - ki(x 2 -f y 2 + * 2 ) for i = 1, 2 or 3, where fei, k 29 and 
fcs are the discriminating numbers of the surface. 

The discriminating numbers are all real (see Theorem 20, 
Chapter VII, Section 89, page 192), but they need not all be 
distinct. Moreover, even though fci, fe, and fc 3 are real, it is not 
certain whether the linear factors of q(x, y, z) k(x 2 + y* + z 2 ) 
are real for i = 1, 2, 3, that is, whether the planes of circular 
section are real. In view of Theorems 2 and 3 we can state there- 
fore that through every point in space there pass six planes which 
cut an arbitrary quadric surface in circles; of these planes some 
may be coincident and some may be imaginary. We proceed 
now to a further study of the different possibilities. 



248 



QUADRIC SURFACES 



We suppose first that the equation A(k) = has no multiple 
roots; then the rank of the matrix 



*(*,-) = 



a 23 



is equal to 2, for t = 1, 2, 3 (see Theorem 19, Chapter VII, Section 
89, page 190) and hence the invariant T z (ki) of the quadric surface 
represented by the equation q(x, y t z) ki(x 2 + y 2 + z 2 ) =0 is 
different from zero. It follows therefore from the table in Section 
102 (page 230) that this surface consists of two distinct real planes 
or two imaginary planes according as T 2 (ki) < or T 2 (ki) > 0. 
But 



| 



- AV) (ass 
+ (an 
2 T 7 ^ + 



Cl\2 #22 "-" ' 

- #23 2 + (a n - ki) (a 33 ki) - #i 3 2 



a i2 2 



3 k? = - 



K 



Now A'(ki) represents the slope of the curve y = A(k) at the 
points where it crosses the K-axis; since we are supposing that 

the equation A(k) =0 has no 
multiple roots, and since the co- 
efficient of fc 3 in A(k) is 1, the 
curve y = A(k) has the general 
character indicated in Fig. 33. 
From this it should be evident 
that of the three numbers A'(ki), 
i = 1, 2, 3, two are negative, 
namely, those which correspond 
to the least and to the greatest 
of the numbers k\, k 2 , /c 3 , whereas 
one is positive, namely, the one 
which corresponds to the middle 
discriminating number; conse- 

* Notice, that in A(k) the coefficient of fc 3 is 1. This interesting formula 
can be obtained directly from the proof of Theorem 19, Chapter VII (see 
Section 89, page 190); the alternate proof given in the text does not make use 
of the formula for the derivative of a determinant, but is not well suited for 
extension to derivatives of higher order. 




FIG. 33 



CIRCLES ON QUADR1C SURFACES 



249 



qucntly, of the numbers T 2 (/Ci)> two are positive and one is 
negative, and if the notation be so chosen that ki < k 2 < k 3 , 
T 2 (ki) > 0, T 2 (k 2 ) < and T 2 (fa) < 0. Therefore the equation 
q(x, y, z) k 2 (x 2 + # 2 + 2 2 ) =0 represents a pair of real planes 
of circular section through the origin, but the equations q(x, y, z) 
- ki(x 2 + y 2 + z 2 ) = and q(x, y, z) - fa(x 2 + y 2 + z' 2 ) = rep- 
resent pairs of imaginary planes. 

If the equation A (k) = has a pair of double roots, let us say 

fci = fc 2 , then the rank of the matrix a 3 (fci) is 1 (see Theorem 19, 

Chapter VII) and therefore the equation q(x, y, z) ki(x 2 +y 2 +z 2 ) 

= represents a pair of coincident planes (see Corollary 3 of 

Y 




K 



K 




FIG. 34a 



FIG. 346 



Theorem 8, Section 96, page 209). From the discussion in the 
preceding paragraph we conclude that in this case the graph of the 
function A(k) has the character indicated in Figs. 34a and 34b, 
from which it should be clear that A' (fa) < 0, therefore that 
T 2 (fa) > and hence that the planes represented by the equation 
q(x, y } z) fa(x 2 + y 2 + z 2 ) =0 are imaginary. In this case 
there are therefore four coincident planes through the origin which 
cut the surface in a circle; each of these planes is represented 
by the equation [q(x, y, z) - ki(x* + y 2 + z 2 )]* = 0. 

Finally, if fci is a triple root of the equation A(k) = 0, the rank 
of the matrix a 8 (fci) is (see Theorem 19, Chapter VII) and there- 
fore the function q(x, y, z) ki(x 2 + y 2 + z 2 ) vanishes identically, 
so that every function of the form ax + by + cz is a factor of it. 
In this case every plane through the origin is a plane of circular 



250 QUADRIC SURFACES 

section. In view of the remark on page 215 this constitutes a 
proof of the well-known fact that the section of a sphere by an 
arbitrary plane is a circle. 

We should recall moreover that if the discriminating equation 
has a pair of equal roots which are not zero, the quadric is a surface 
of revolution, while if it has a pair of zero roots, the quadric is a 
parabolic cylinder or a pair of parallel or coincident planes. 
In view of these facts we can state the following conclusion : 

THEOREM 4. The quadrics which are surfaces of revolution but not 
spheres, the parabolic cylinder, and the pair of parallel or coincident 
planes possess through every point of space A(<* 9 /?, 7) four coincident 
planes of circular section; in these cases there exists a double root, fci, of 
the discriminating equation, the function q(x 9 y, ) - k^x 2 -f y 2 -f * 2 ) 
is a perfect square and the planes of circular section through A are 
given by the equation [g(x - , y - p, s ?)] 4 =0, where [g(x 9 y, *)] 2 = 
q(x 9 y, z) - fci(* 2 -f-y 2 + 2 ). All other quadric surfaces possess 
through every point of space A(a 9 (3, y) two distinct planes of circu- 
lar section; no two of the discriminating numbers k l9 7c 2 , k 3 are 
equal to each other and, if ki < k 2 < fc 3 , the function q(x 9 y, z) 
kz(x* + y 2 + * 2 ) is factorable into two linear factors with real coeffi- 
cients; if we write q(x 9 y, *) - fc 2 (* 2 + y 2 4- * 2 ) = gi(x 9 y, *) g 2 (x 9 y, *), 
the planes of circular section through A are given by the equations 

gi(* - , y , z - 7) =0 and g*(x - a, y - 0, s - 7) = 0. 

Remark 1. The circles of these circular sections may be ordi- 
nary circles with a finite center and finite radius, or they may 
be "degenerate circles" (compare Appendix, VIII, page 301). It 
should be clear that the circular sections of degenerate quadrics 
are always degenerate circles. And it should be clear that this 
will also be the case when the middle root or the double root of 
the discriminating equation is equal to zero (compare also Sec- 
tions 110 and 111). 

Remark 2. If the notation for the discriminating numbers is 
so chosen that in all cases ki g k 2 g & 3 , the planes of circular 
section through the origin are always given by the equation 
q(x, y, z) - fc 2 (x 2 + y 2 + z 2 ) = 0. 

Example. 

The locus of the equation 



is an hyperboloid of one sheet; for A = 



300-3 

0-1 3 2 

3-1-1 

-3 2 -1 -2 



= 99, 



CIRCLES ON QUADRIC SURFACES 251 

300 

0-1 3 = -24, T 2 = -14, so that A > and T 2 < 0. More- 

03-1 

over we find that T\ = 1. The discriminating equation is k 3 k 2 14 k + 24 
= 0; its roots are 2, 3, and 4. Therefore in the notation of Remark 2 
above, /b 2 = 2, and the planes through the origin which cut the surface in 
circles are given by the equation 3 x 2 y 2 z 2 -f 6 yz 2 (x 2 -{- y 2 + z 2 ) =0 f 
that is, by x 2 - 3 (y - z) 2 = or by x - \/3 y -f V3 z = and x -f \/3 ?/ - 
-s/3 2 = 0. The planes of circular section through an arbitrary point A (a, 0, 7) 
are given by the equations (x a) V^(?/ 0) -h \/3(z 7) = and 
(x a) -f- V3(y 0) V3(z 7) =0. 



110. Circles on Quadric Surfaces, continued. To determine 
the planes of circular section for a particular quadric surface, whose 
equation is given in numerical form, we can proceed by the general 
method developed in the preceding section. The work becomes 
very simple if the equation of the surface has first been reduced 
to the standard forms of Sections 100-102. 

Examples. 

X* V* Z 2 

1. If in the equation of the ellipsoid i -h p + -5 = 1> a < b < c, the 

middle discriminating number of the surface is r~ 2 . Therefore the planes 
of circular section through the origin are given by the equation 



that is, by 

In virtue of the relative magnitude of a, b, and c, which has been presupposed, 
the coefficients of z 2 and z 2 in this equation are both positive. Hence the 
planes of circular section through the origin are represented by the equations 

cxVb* - a 2 - azVc 2 - 6 2 = and czV& 2 a 2 + azVc 2 - 6 2 = 0; 

and the planes of circular section through the arbitrary point A (a, 0, 7) are 
given by the equations 

cVb* - a?(x - a) = aVc 2 - b*(z - 7) and cVb 2 - a 2 (x - a) = 
-aVc 2 - b*(z - 7). 

2. For the parabolic cylinder y 2 - 4 pz, the discriminating numbers are 
0, 0, 1; therefore fc 2 = and the planes of circular section are given by the 
equation y* = 0. This equation represents the ZZ-plane counted fourfold. 
The planes of circular section through the point A (a, 0, 7) are represented by 
the equation (y 0) 4 = 0; therefore they are the planes y = ft counted 
fourfold. The intersection of a plane of this family consists of the generating 



252 QUADRIC SURFACES 

s 2 
line y = ft, z = ~- and of the infinitely distant line of the plane y = 0; the 

circular section is therefore a degenerate circle (compare Appendix, page 301). 
111. Exercises. 

1. Determine the planes of circular section through the point A ( 2, 3, 1) 
for each of the following surfaces: 



2. Prove that the circular sections of a hyperbolic paraboloid are always 
degenerate. 

3. Prove that the two families of planes of circular section of a central 
quadric are not affected when the surface is translated. 

4. Determine the planes of circular section through the point A (3, 4, 1) 
for each of the following surfaces: 

(a) x 2 + 3 y 2 - z 2 = 0, (6) 4 x 2 + 9 if = 1, (c) 4 x 2 - 9 y* = 1. 

6. Prove that the circular sections of the hyperbolic cylinder are always 
degenerate. 

6. Determine the angle between the two planes through the origin which 

2 ? ,2 2 2 

cut the ellipsoid -5 -f r? + = 1 in circles; and set up the condition under 
a* o* c* 

which these planes will be perpendicular. 

7. Solve the similar problem for the hyperboloid of one sheet and also for 
the hyperboloid of two sheets. 

8. Determine the planes of circular section through the point A (2, 1, 1) 
for each of the following surfaces: 

(a) 4 x 2 + 6 ?/ 2 + 4 * 2 = 1, (6) x* - 2 y* + z 2 = 1, 
(c) 2 z 2 - r/ 2 - z 2 = 1, (d) 4 x 2 + 4 ?/ 2 = 5 z, 

(e) 6 x 2 - if 4- 6 z 2 = 0. 

9. Prove that for an ellipsoid of revolution the planes of circular section 
are perpendicular to the axis of revolution of the surface; prove the same prop- 
erty for the hyperboloids of revolution of one and of two sheets. 

10. Derive the equations of the planes of circular section through an arbi- 
trary point for the hyperboloid of one sheet, and also for the hyperboloid of 
two sheets with respect to a system of axes which are parallel to the principal 
directions of the surfaces. 

11. Solve the corresponding problem for the proper cone and for the elliptic 
cylinder. 

12. Determine the condition under which the two planes of circular section 
of the elliptic cylinder which pass through a fixed point are perpendicular to 
each other. 



TANGENT PLANES PARALLEL TO A GIVEN PLANE 253 

112. Tangent Planes Parallel to a Given Plane. The Umbilics 
of a Quadric Surface. It may happen that, even though the 
planes of circular section of a quadric are real, yet the sections 
themselves fail to be real because the plane does not meet the 
surface in real points; a limiting case arises when a plane of cir- 
cular section is tangent to the surface. In that case, if we are 
dealing with a family of real planes of circular sections, which are 
non-degenerate, the circle of section reduces to a point; such a 
point on a surface is called an umbilical point, or an umbilic. 

DEFINITION I. An umbilic of a quadric surface is a point on the 
surface at which the tangent plane is parallel to a plane through the 
origin which cuts the surface in a non-degenerate circle. 

Remark. It follows from this definition and from Sections 110 
and 111 that umbilical points can exist at most on the central 
quadrics, the cone, the elliptic paraboloid and the elliptic cylinder. 

Since these quadrics have at most two sets of parallel planes of 
circular section, the existence of umbilical points depends upon 
the existence of points on the surface at which the tangent plane is 
parallel to the planes of these sets. On account of the intrinsic 
interest of the question we shall preface the further discussion of 
umbilics by a treatment of the general question of determining 
points on a quadric surface at which the tangent plane is parallel 
to a given plane; and we shall discuss this problem for all real 
non-degenerate quadric surfaces. 

Let ax + by + cz = be an arbitrary plane through the origin 
(a, 6, and c not all zero). Does there exist a point P(a, 0, 7) on the 
surface Q such that the plane tangent to the surface at P is parallel 
to the given plane? Since the tangent plane to the surface at P 
is represented by the equation 

(x - )Qi(, ft 7) + (V ~ PXMa, ft 7) + (z - 7)Q 3 (, ft 7) = 0, 

the conditions of the problem require that there shall exist a non- 
zero factor of proportionality 2 p, such that the coordinates of P 
satisfy the equations 

Qi(, ft 7) = 2 pa, Q 2 (a, ft 7) = 2 P 6, Q 3 (a, ft 7) = 2 pc; 

and moreover these coordinates must satisfy the condition Q(a, ft 7) 
= 0. The latter equation may be written in the form 

oQi(a, ft 7) + j8Q 2 (a, ft 7) + 7Qs(, ft T) + Qi(<*> ft T) = 0, 



254 



QUADRIC SURFACES 



which, by use of the first three equations, may be replaced by the 
equation 

Q 4 (, ft 7) + 2 p(aa + bft + 07) = 0; 

and this equation has the advantage of being linear in a, ft and 7. 
We find therefore that a, ft and 7 must satisfy the following four 
linear equations: 

ana + o 12 + ai 3 7 + OH - pa = 0, 

Oi 2 a + 022/3 + a 2 37 + 024 P& = 0, 
013CK + 023/3 + 0337 + 034 PC = 0, 

(o 14 + po)a + (024 + pb)P + (o 34 + pc)7 + O 44 = 0. 



(1) 



If these equations are looked upon as forming a system of linear 
equations in a, ft 7, it follows from Corollary 2 of Theorem 24, 
Chapter IV (Section 54, page 102) that they possess no solution, 
unless the rank of the augmented matrix is less than 4. We are 
led therefore to the following equation for p: 



R(p) 



On 



OH + pO 



Oi 2 

2 2 

2 3 

24 + pb 



Oi 3 OH ~ PO 

2 3 2 4 pb 

33 3 4 PC 

034 + PC 044 



= 0. 



We write this determinant as the sum of two determinants, 
using OH, o 2 4, 034, o 44 as the elements of the 4th column in one and 
pa, p&, pc, as the elements of the 4th column in the other; 
each of these determinants is again written as the sum of two de- 
terminants by making a similar distribution of the elements of the 
4th row. The equation will then take the following form : 



OH Oi 2 


Oi3 


OH 




On 


Oi2 


Oi3 


O 


Oi2 O22 


2 3 


O24 




Oi2 


O22 


23 


& 








P 










Ois O 2 3 


033 


3 4 




Oi3 


23 


033 


C 


b 


C 







OH 


024 


034 







On 


Oi2 


Oi3 O 










P 2 


Ol2 
Oi3 


022 
O23 


023 b 
033 C 


= 0. 










O 


b 


c 











Now it should be an easy matter to see that the two middle terms 
are equal numerically and opposite in sign; moreover the coeffi- 



TANGENT PLANES PARALLEL TO A GIVEN PLANE 255 

cient of p 2 is of the same form as the determinant which we denoted 
by the symbol A*(Q) in Section 84 (see page 172), and will by 
analogous notation be designated by Aa(a, b, c). The equation 
for p can then be written in the simple form 

(2) A = p 2 X A 3 (a, 6, c). 

To each root of this equation there corresponds a single value for 
each of the variables a, /?, y, to be determined from the equations 
(1), provided the rank of the coefficient matrix of these equations 
is 3. The discussion of the different possibilities, and also of the 
cases in which the roots of the equation (2) are real and distinct, 
real and equal, or complex is made most conveniently after the 
equation of the quadric surface has been reduced to the standard 
forms, discussed in Chapter VIII. The translation and rotation 
of axes which are involved in this reduction will of course affect 
the equation ax + by + cz = of the plane. It is therefore of 
importance to establish first the following theorem. 

THEOREM 5. The value of the determinant A 3 (a, 6, c) and the rank 
of its matrix are invariants of the configuration consisting of the sur- 
face Q and the plane ax + by +cs =0 with respect to translation and 
rotation of axes. 

Proof. The most general transformation of coordinates which 
can be made by rotation of axes is carried out by means of the 
equations 



x = XiZi + \ 2 yi + XaZi, y = 

+ vtfi 

(compare Theorem 5, Chapter V, Section 63, page 121). If these 
expressions are substituted for x, y, and z in the equation ax + by 
+ cz = 0, it is carried over into the equation o!x\ + Vy\ + G'ZI = 0, 
where 

a' = a\i + bp,i + cvij V = aX 2 + 6ju 2 + cv^ c' = a\ 3 -f 6/x 3 



We observe now that this transformation of the coefficients of 
x, y, z in the equation of the plane is exactly the same as the trans- 
formation of the coefficients a i4 , a 24 , a 34 of the linear terms in Q 
under rotation of axes (compare page 212) ; consequently the de- 
terminant -A 3 (a, b, c) is transformed by rotation of axes exactly 



256 



QUADRIC SURFACES 



like the discriminant of the quadric surface q(x, y,z) + 2 ax + 2 by 
+ 2 cz = 0. It follows therefore from Theorems 4 and 5, Chapter 
VIII (Section 94, pages 203 and 205) that the value of the de- 
terminant A 3 (a, 6, c) and the rank of its matrix are invariant with 
respect to rotation of axes. That this invariance also holds with 
respect to translation of axes becomes evident if we recall that the 
coefficients of the second degree terms in Q are not changed by 
translation of axes (compare proof of Theorem 1, Chapter VIII, 
Section 93, page 199) and if we observe that the transformation 

x = x' + p, y = y' + q, z = z' + r 

carries the equation ax + by + cz = over into the function 
ax f + by' + cz' + ap + bq + cr = 0, so that the coefficients a, 6, 
and c are also invariant under translation of axes. 

In the further discussion of our problem we shall now be able to 
use the standard forms of the quadric surfaces. 
CASE I. r 4 = 4, n = 3. 

(a) Ellipsoid. The standard form of the equation is 



We find that A = - 



and A 3 (a, 6, c) = 



-(33+33+^5)- The c q uation ( 2 ) 



1 








a 




p2 












1 
















b 


^ 






1 














c 








r 2 






a 


& 


c 







)ecomes therefore 



P 2 (a 2 p 2 + & 2 # 2 + c 2 r 2 ) = 1. It has two real roots for every set'of 
values of a, 6, and c and, since r 3 = 3, a single set of values of 
a, /?, 7 is given by equations (1) for each root of (2). Therefore, to 
every plane there correspond two points on the ellipsoid at which 
the tangent planes are parallel to the given plane. 

(6) Hyperboloid of One Sheet. From the standard form of the 
equation, namely, 



TANGENT PLANES PARALLEL TO A GIVEN PLANE 257 

1 a 2 b 2 c 2 

we find A = , and A 3 (a, 6, c) = -^ + -r-r -- =-= . In this 
p2g2 r 2 ' \ j j / qt r z r 2 p 2 p 2 q 2 

case A*> 0, although A 3 (a, 6, c) is positive, zero or negative ac- 
cording as a 2 p 2 + b 2 q 2 c 2 r 2 is positive, zero or negative. Since 
we have again r 3 = 3, we conclude that if a, 6, and c are so chosen 
that a 2 p 2 + b 2 q 2 > cV 2 , there are two points on the surface at 
which the tangent plane is parallel to the plane ax + by + cz 0; 
if a 2 p 2 + b 2 q 2 < c 2 r 2 , there are no such points on the surface, and 
if a 2 p 2 + b 2 q 2 c 2 r 2 , there is no finite point on the surface which 
has this property. 

/j2 

~ 2 



/ /j 

(c) Hyperboloid of Two Sheets. Using the standard form -5 ~ 2 



z* 1 a 2 b 2 

2 = 1, we find A = -- r-y- and A 3 (a, 6, c) = -- =-3 + -7-5 
r- 2 p 2 q 2 r 2 q 2 r 2 r 2 p 2 

c 2 
H 2"! 2* Now A < 0, so that equation (2) furnishes real values of 

p only in case b 2 q 2 + c 2 r 2 < a 2 p 2 . In this case there will be two 
real points on the surface of the desired kind; in no other case will 
points of this kind exist at finite distance. 
The conclusions for this case are therefore as follows: 

THEOREM 6. On the ellipsoid there are, for every plane In space, 
two finite points at which the tangent plane is parallel to the given 
plane; on the hyperboloid of one or two sheets two such points exist 
for certain planes but none for others. 

CASE II. r 4 = 4, r 3 = 2. 
(a) Elliptic Paraboloid. The standard form of the equation is 



n 2 c 2 

We find A = ^ and A^(a y 6, c) = ^ . Equation (2) be- 
comes therefore n 2 = p 2 c 2 . If c = 0, there is no finite value of p 
which satisfies this equation and hence no finite point on the sur- 
face which satisfies the conditions of our problem. If c ^ 0, we 

71. YL TL 

have p = - ; let pi = - and p 2 = . If we substitute pi or p2 

c c c 

in the equations (1) for this case, we obtain a system of linear equa- 
tions whose augmented matrix is 



258 



QUADRIC SURFACES 



-, o o -< 



3 - P1 6 












-2n 




or 



1 








-P2a 


p 2 










1 






o 







P2& 




f 


















P2 


P2& 


2n 






The rank of each of these matrices is clearly less than 4. The 
three-rowed principal minors formed from the 1st, 2nd, and 4th 

<a 2 6 2 \ /a 2 6 2 \ 

-r + -r ) and p2 2 ( -r + - I 
g* p 2 / \q 2 p 2 / 

respectively; and since pi and p2 are both different from zero, we 
conclude that these two matrices are both of rank 3. 

The rank of the coefficient matrix for the first of these systems 
of equations is manifestly less than 3; for the second system the 
coefficient matrix contains the non-vanishing three-rowed minor 







o 4 



whose value is 



2n 
Z>V 



P2& -2 n 



In accordance with Corollary 1 of Theorem 



24, Chapter IV (Section 54, page 102) we conclude that for p = pi, 
the system (1) has no solutions, while f or p = P2 it possesses one 

a 



solution. For p 



y3 
= 0, 



P2, the equations (1) are 
= 0, pz(aa + 6)3) 2 717 = 0. From these we obtain, since 

P2 = -- , the following solution of our problem: 
c 



a = 



anp* 



y = 



Thus, while the equation (2) has two real solutions in this case, 
only one of them gives rise to a point (a, 0, 7) on the surface at 
which the tangent plane is parallel to an arbitrarily given plane 
through the origin, except when this plane has the equation 
ax + by = 0. 



TANGENT PLANES PARALLEL TO A GIVEN PLANE 259 

(6) Hyperbolic Paraboloid. From the standard form of the 
equation 

T 2 7 ,2 

x JL = 2 nz 
p% q2 

n 2 c 2 

we obtain A = -y-^and At(a, 6, c) = 2 . The equation (2) re- 
duces, as in (a) to the form n 2 = p 2 c 2 . For c = 0, there is no finite 

solution of the problem; for c ={= 0, we find as before, pi = - and 

c 

P2 = . For these two values of p the augmented matrices of the 
c 

systems of equations (1) become 



1 








-Pia 




1 








-P20 




1 










1 














p\b 


and 










pzb 




9 2 










<T 2 















-2n 
















plC 


i pib 










P2<* 


P2& 


-2n 






It is seen that, as in (a), both these matrices are of rank 3; also 
that the rank of the coefficient matrix for the first system of equa- 
tions is 2 and the rank of the coefficient matrix for the second sys- 
tem of equations is 3. The conclusion is therefore the same as 

for the elliptic paraboloid; we obtain the point a = 



_ bnq* __ n(a 2 p 2 
3 -~> y - 



- 7>2,,2 



THEOREM 7. On the elliptic paraboloid and on the hyperbolic parab- 
oloid, there is for every plane in space, except for planes parallel to 
the axis of the surface, one point at which the tangent plane is parallel 
to the given plane. 

CASE III. n r 3 = 3. 

The on}y real quadric in this case is the proper cone, whose 
standard equation may be put in the form 



2CO 



QUADR1C SURFACES 



We find A = and 



A 3 (a, b, c) = 



1 








a 





_! 





6 




a 



b 


_!_ 

r 2 
c 


C 





__ p 2 a 2 + & V - c 2 r 2 



If -A 3 (a, b, c) = ^p 0, the only solution of equation (2) is p 0, 
and equations (1) reduce to the equations for the vertex (compare 
Theorem 13, Chapter VII, Section 85, page 177). Since at the 
vertex Qi = $2 = Qa = 0, the tangent plane at this point does not 
exist and our problem has therefore no solution in this case. On 
the other hand, if A 3 (a, 6, c) = 0, i.e., if p 2 a 2 + 6 2 g 2 - c 2 r 2 = 0, 



every value of p satisfies equation (2). 



the system of equations (1) is 



-4 o 



The augmented matrix of 




-pa 



i -p6 





pa p& 



pc 



-pc 




; its deter- 



minant is equal to p 2 As(a, 6, c) and vanishes. It should be clear 
that the rank of this matrix is 3 and that the rank of the coefficient 
matrix of the system of equations (1) is also 3. Hence, for every 
value of p there is one point on the surface which satisfies the con- 
ditions of our problem. By solving the system (1), we find 



a 



pap 2 , 



|3 = pbq 2 , 7 = per 2 . 



As p varies, these equations are the parametric equations of a line 
and it is readily seen that this line lies entirely on the cone. For, 

since a 2 p 2 + b 2 # 2 - c 2 r 2 = 0, it follows that ^ + ^ - \ = 

P 2 (a 2 p 2 + & 2 # 2 "~" c * r ^ ~ 0> independently of the value of p. And 
the tangent plane to the cone at any point on this line is repre- 



sented by the equation ^ H ^ -- ^ = 0, i.e. by the equation 



TANGENT PLANES PARALLEL TO A GIVEN PLANE 261 

ax + by + cz = 0. We have therefore reached the following 
conclusion : 

THEOREM 8. If and only if a, 6, and c are so chosen that a 2 p 2 + 6 V 
c 2 r 2 = 0, the plane ax -f by + cs = will be tangent to the cone 

*_ -|_2_ _ ?L = o; this plane touches the surface along the line on the 

cone whose parametric equations are x = pap 2 , y = p6q 2 , z = per 2 , 
p being the parameter. For any other choice of a, b, and c there will 
be no point on the cone at which the tangent plane is parallel to the 
plane ax + by -f- cs = 0. 

CASE IV. n = 3, r 3 = 2. 

(a) Elliptic Cylinder. The standard equation is \ + ~ = 1 ; 

P 5 

c 2 
A = and A 3 (a, 6, c) = ^-: 2 . If c = 0, the only solution of 

equation (2) is p = 0; the system (1) is inconsistent for this value 
of p and we have therefore no solution of the problem. If c 0, 
the system of equations (1) reduces to 

". = pa $- = pb p(aa + bft) = 1 
p2 ^2 P y 

Values of p can always be determined for which this system is 
consistent; with these values of p, we find for a. and ft the follow- 
ing results: 



These equations determine a line parallel to the Z-axis, which lies 
entirely on the cylinder; and the plane tangent to the cylinder at 
any point on this line is represented by the equation ax + by 
= Va*p* + 6y. 

x 2 ?/ 2 
(b) Hyperbolic Cylinder. From the standard equation ~ 

c 2 
= 1, we find A = and Az(a, 6, c) = -j- 2 . Conditions are similar 

to those in (a). If c ^ O y there is no tangent plane parallel to the 
plane ax + by + cz = 0. If c = 0, the values of p and of a, /3, y 
are to be determined from the equations (1) which reduce in this 

case to 

a = pap 2 , ft = -pbq\ p(aa + b/3) = 1. 



262 QUADRIC SURFACES 

Elimination of a and ft leads to the equation p 2 (a 2 p 2 6 2 # 2 ) = 1 ; 
hence if and only if a 2 /? 2 ~ &V > 0, does there exist a real point 
which satisfies the conditions of the problem. In this case, we find 

ap 2 bq 2 

a "~ > ~~ 



The line determined by these equations lies entirely on the surface > 
and the equation of the plane tangent to the cylinder at any point 
of this line is ax + by = db Va 2 p 2 6 2 j 2 . 
CASE V. r 4 = 3, r 3 = 1. 

In this case the locus of the equation Q - is a parabolic cylin- 
der; the equation of this surface may be reduced to the standard 
form y 2 4 px = 0. We have 

a 



A = and -As (a, b, c) = 



0106 
c 
a b c 



= 0. 



Hence equation (2) imposes no restriction on the choice of p. The 
system of equations (1) takes the form 

-2 p - pa = 0, ft - pb = 0, -pc = 0, (-2 p + pa)a 
+ pb/3 + PCT = 0. 

If c 4= 0, the third equation requires that p == 0, which leads to a 
contradiction with the first equation, since p ^ 0; in this case 
there is therefore no solution. For a similar reason, we must have 

2 v 
a ^p ; and in this case we find p = , and hence ft = pb = 

2 vb vb^ 

, a = ^~. These equations represent a line which lies en- 
tirely on the cylinder; and the tangent plane to the surface at any 
point on this line is given by the equation yft 2 p(x + a) = 0, 

vb 2 

that is, by the equation ax + by + = 0. There is no finite 

a 

point on the cylinder at which the tangent plane is parallel to the 
plane ax -f by = 0, if a = 0; that is, there is no tangent plane 
parallel to the plane y = 0. 

We summarize now the conclusions reached in Cases IV and V 
in the following theorem. 



THE UMBILICS OF A QTTADRIC SURFACE 263 

THEOREM 9. The tangent planes to the elliptic cylinder, the hyper- 
bolic cylinder and the parabolic cylinder are all parallel to the gener- 
ating line of the cylinder; they have contact with the surface along an 

entire generator. The elliptic cylinder -- + ^ = 1 has two tangent 

planes parallel to an arbitrary plane ax -f by = through its axis, 
namely, the planes ax + by = V a 2 p 2 + 6 2 q 2 . The hyperbolic cylin- 
der 2 - ^7, = 1 has two tangent planes parallel to an arbitrary plane 

ax + by = through its axis, namely, the planes ax + by 
= Va~p 2 - 6V, unless a 2 p 2 - &V :g 0, in which case no such plane 
exists. The parabolic cylinder y 2 = 4px has one tangent plane parallel 
to the arbitrary plane ax -j- by = through the Z-axis, namely, the 

plane ax + by H- = 0, if a ^ 0; if a = no such plane exists. 

113. The Umbilics of a Quadric Surface, continued. The 
determination of the umbilics on the central quadrics, the cone, 
the elliptic paraboloid and the elliptic cylinder can now be effected, 
on the basis of Definition I (Section 112, page 253), by combining 
the results of Theorems 6, 7, 8, and 9 with those of Theorem 4. We 
shall take the equations of these surfaces in the standard forms 
used in the preceding section. 

(a) Ellipsoid. 

If p < q < r, the equations of the planes of circular section, 

through the origin, of the ellipsoid -2 + ^2 + -^! =0 are (see 

Section 110): 

rVq* - p 2 x pVf 2 - q 2 z = 0. 

Hence, in the notation of Section 112, we have a = rVq* 
b = 0, c = pVr 2 q 2 , and p = =b 



g cr 

, the double sign of p being independent of that of c. 

prVr' 2 p 2 

The first three equations of the system (1) of Section 112 are for 

a 3 *v 

this case z = pa, 2 = pb, -^ = pc. We find therefore 




the double signs of a and 7 being independent of each other. 
conclusion is given by the following theorem. 



264 QUADRIC SURFACES 



THEOREM 10. The ellipsoid ~+^ + ^j = I, in which p < q < r, has 




four umbilics. Their coordinates are ( = __ , . 

\ Vr 2 - p 2 Vr' 2 - p 2 

they He on the ellipse in which the surface is cut by the XZ-plane. 

(b) Hyperboloid of One Sheet. If the equation be taken in the 

X 2 ?/ 2 2 2 

standard form -^ + ^r -- -: 0. p < q. the central planes of cir- 
p2 qt r 2 r i 

eular section are (compare Exercise 10, Section 111, page 252): 
rVq 2 - p' 2 x =FpVq' 2 + r 2 2 = 0. 



Hence a = rVq 2 p 2 J b = 0, c = ^FpVq' 2 + r 2 ; and a 2 p 2 

+ 6 2 g 2 - c 2 r 2 - p 2 r >2 (f - p 2 ) - r*-p*(q 2 + r 2 ) = - p 2 f 2 (p 2 + r 2 ) 

< 0. Therefore, the hyperboloid of one sheet has no umbilics. 

(c) Hyperboloid of Two Sheets. We proceed as in the preceding 

/v2 7/^ 2J 

case, for the equation -7, ~ -- r > = l,q < r. The central planes 

p2 g2 r 2 ^ * x 

of circular section are represented by the equations 

(j Vr' 2 + ;> 2 .r =h p\/r 2 - ^/ 2 ?y = 0. 



We have a = ^V^ 2 + /> 2 , b = =t /;Vr 2 - <? 2 , c = 0; a 2 p 2 - 6 2 ^ 2 
-c 2 r 2 = pV(r 2 + p 2 ) - 2 p 2 (r 2 - g 2 ) - p 2 g 2 (p 2 + g 2 )> 0. There- 
fore there are umbilical points on this surface. Their actual de- 
termination proceeds as in the case of the ellipsoid; the reader 
should have no difficulty in completing the proof of the following 
theorem : 

THKORKM 11. On the hyperboloid of one sheet there are no umbilics. 
On the hyperboloid of two sheets, which is not a surface of revolution, 

X 2 V 2 -2 

there are four umbilics; if the equation of the surface is -, - , -- 

pi q 2 r 2 

( D \/r 2 4- n 2 n\/r' 2 n z \ 
p . ^" , L ~ "~ , o \ the 
Vp 2 -f q 2 Vp 2 + q* ) 

two double signs being independent of each other. 

/$ ^,2 ^2 

(d) Cone. If the equation be written in the form -7, + ~ -- - 

p z q z r~ 

= 0, p < q, the planes of circular section through the origin are: 
r Vq* - p' 2 x pVg 2 + r 2 z = 0. 



THE UMBILICS OF A QUADRIC SURFACE 265 



Here a = rVq* - p*, b = 0, c = dbpVV + r 2 ; and a 2 p 2 + 6 2 g 2 
- cV 2 *= pV 2 (g 2 - p 2 ) - p 2 rV + r 2 ) = - pV 2 (p 2 + r 2 ) 4= 0. 
Therefore the cone has no umbilics. 

x 2 ?y 2 
(e) Elliptic Paraboloid. From the standard equation + -% 

= 2nz,p < q, we obtain for the planes of circular section through 
the origin, the equations Vq 2 p 2 x pz = 0. Hence 



b = 0, c = 

Tliere are therefore two umbilics on this surface; from the formulas, 
given in Section 112, Case II, (a), page 258, we find that their co- 
ordinates are 



= 0, 7 = 



(/) Elliptic Cylinder. We take the equation in the form H - 

= 1, /; < g, and we find that the circular sections through the ori- 
gin lie in the planes represented by the equations v q 2 p 2 x pz 
= 0. Since these planes are not parallel to the generators of the 
cylinder, it follows from Theorem 9 that there are no umbilics on 
this surface. 

The results found in cases (rf), (e), and (/) lead to the following 
theorem. 

THEOREM 12. There are no umbilics on the cone, nor on the elliptic 
cylinder. There are two umbilics on the elliptic paraboloid which is 
not a surface of revolution; if the equation of this surface be reduced 

to the form ^ 2 -h ~ = 2 ns 9 p < q 9 the coordinates of the umbilics are 



(g) Surfaces of Revolution. We have seen in Theorem 4 that for 
a surface of revolution there is one plane of circular section through 
every point of space; these surfaces can therefore have at most 
two umbilics. The single central plane of circular section may in 
these cases be obtained from the results already found, 'by setting 
two of the discriminating numbers equal to each other. So, for 
example, if in the ellipsoid, treated under (a), we put p = q y the 
four umbilics reduce to two, namely to the points (0, 0, =hr), that 



266 QUADRIC SURFACES 

is, to the points in which the axis of revolution meets the surface. 
The reader should have no difficulty in obtaining the corresponding 
result for the hyperboloid of revolution of two sheets and for the 
paraboloid of revolution. On a sphere every point is an umbilic. 

THEOREM 13. On the ellipsoid of revolution, the hyperboloid of 
revolution of two sheets and on the paraboloid of revolution, the urn- 
bllics are the points in which the surface is met by the axis of revolu- 
tion; on the other quadrics of revolution there are no umbilics. 

We summarize the results of this section as follows: 

Umbilics exist on the ellipsoid, the hyperboloid of two sheets and 
the elliptic paraboloid. If these surfaces are not surfaces of rev- 
olution, the number of umbilics on them are 4, 4, and 2 respectively; 
if they are surfaces of revolution, the umbilics fall two by (wo into 
the points where the surface is met by the axis of revolution. 

114. Exercises. 

1. Determine the planes parallel to the plane z 27/4-2=0 which are 
tangent to the following surfaces: 

(a) J 2 - 6 ^ - 3 z 2 = 1, (6) 2 x* + 4 7/ 2 + 5 2 2 = 1, 

(c) x 2 4- 4 ?y 2 = 2 z, (d) 4 x* - ?/ 2 = 2 z. 

2. Determine the umbilics on each of the following surfaces: 



(c) 4- = 2*, (d) x* 4- y* + 4 z 2 = 12. 

3. Prove that the planes through the point P(a, ft 7) which are parallel to 

# 2 ?/ 2 2 

the planes tangent to the cone ~ -\- ~ = are tangent to a cone whose 

vertex is at P. 

4. If u, v, w are called the " coordinates " of the plane ux -\-vy-\-wz + 1=0, 
set up the equation which the coordinates of a plane must satisfy in order that 

the plane be tangent to the cone ( ^~ + (V ""/^ - ( -^~ = 0. 

X z 1/ z 2 2 

6. Determine the umbilics on the surface -- 2^2 I = 1* P < r - 

6. Prove that the planes of circular section of the quadric surface Q(x, y, z) 
= are also planes of circular section of all the quadric surfaces whose equa- 
tions have the form Q(x, y, z) + k(x* + y 2 + z*) = 0. 

7. Prove that the four umbilics of an ellipsoid are the vertices of a rectangle; 
and determine the condition under which they will be the vertices of a square. 

8. Prove that the umbilics of the ellipsoids with the same center and the 



EXERCISES 267 

same principal directions, whose semi-axes are kp, kq, and kr, in which p, q, and 
r are fixed while k is variable, lie on four lines through the common center; 
and determine the direction cosines of these lines. 

9. Determine the condition under which the tangent planes to an elliptic 
paraboloid at the umbilical points are perpendicular to each other. 

10. Prove that the planes of circular section of an hyperboloid of one sheet 
are also planes of circular section of its asymptotic cone. 

11. Prove that the tangent planes to an hyperboloid of two sheets at its 
umbilical points are planes of circular section of its asymptotic cone. 

12. Prove that the planes of circular section of a proper cone cut a tangent 
plane of the cone in lines which make equal angles with the generator of the 
cone along which the tangent plane touches it. 



CHAPTER X 
PROPERTIES OF CENTRAL QUADRIC SURFACES 

115. Conjugate Diameters and Conjugate Diametral Planes 
of Central Quadrics. Enveloping Cylinder. 

DEFINITION I. A diameter of a central quadric surface is 21 chord 
which passes through the center of the surface. 

The common form of the standard equations of the central 
quadrics is 

(1) mix 2 + W2?/ 2 + m s z 2 = 1. 



Here m\, 11^ and m^ are the quotients of the discriminating num- 

bers of the surface by j ; the center of the surface is at the 

^144 

origin and the principal directions arc the directions of the co- 

ordinate axes. 

Let us now consider an arbitrary diameter d of the surface; 

we shall designate its direction cosines by Xi, MI, v\- The diametral 

plane corresponding to this di- 
rection (see Definition X, Chap- 
ter VII and Theorem 17, Section 
88, page 186) in the surface (1) 
is given by the equation 




(2) 

If a second diameter 
rection cosines X 2 , M, 
this plane, 

(3) 



= 0. 

with di- 
^ lies in 



= 0, 

(compare Theorem 21, Chapter 
Fio. 35 IV, Section 52, page 99); and 

the equation of the diametral 
plane corresponding to the direction of c? 2 has the equation 

(4) WiX 2 + miMJ + m$v& = 0. 
268 



CONJUGATE DIAMETERS OF CENTRAL QUADRICS 269 



Equation (3) can now be interpreted as stating that the line 
di lies in the plane (4). We have therefore proved the following 
theorem (see Fig. 35). 

THEOREM 1. If one diameter of a central quadric surface lies in the 
diametral plane determined by the direction of another diameter, 
then the second diameter lies in the diametral plane determined by 
the direction of the first. 

COROLLARY. The plane determined by two diameters is the diametral 
plane which corresponds to the direction of the line of intersection of 
the diametral planes of the first two diameters. 

Proof. The diameter (/ 3 in which the planes (2) and (4) inter- 
sect has direction cosines X 3 , MS, vs determined by the proportion 



(5) X 3 : 



Mi v\ 

M2 V'2. 



MS MI 



v\ 



, 

A'2 



MI MO 



A! MI 

X-2 M2 



The diametral plane of this line* is represented by the equation 

Xi , 



M'2 



M2 



= 0. 



But this equation is equivalent to the equation 



x 



y z 

XL Mi "i 

X2 M2 ^2 

which is indeed the equation of the plane determined by the lines 
di and d%. 

The corollary is also proved by the fact, that since rf 3 lies in the 
diametral planes of d\ and d 2 , the diametral plane of rf 3 contains 
the diameters d\ and d%. 

Of the three diameters d\, d 2 , and e/ 3 any two determine the 
diametral plane of the third; and of the three planes (d h d 2 ), 
(di y d 3 ), and (d 3 , d\) any one is the diametral plane of the line of 
intersection of the other two. 

The correspondence between diameters and diametral planes is 
a reciprocal one-to-one correspondence : not only is there one and 
only one diametral plane of any line through the center, but there 
is also one and only one diameter of which any given plane through 

* Whenever it can be done without danger of confusion, we shall use the 
phrase "diametral plane of a line V 1 in place of the more exact but less con- 
venient expression "diametral plane corresponding to the direction of the 
line L" 



270 PROPERTIES OF CENTRAL QUADRIC SURFACES 

the center is the diametral plane. For if ax + by + cz = is 
an arbitrary plane through the origin, there is one and only one 
set of values X, /x, v such that mi\ : w 2 M : ni$> = a :b : c, hence one 
diameter d whose diametral plane mi\x + ^2M2/ + m & z co " 
incides with the given plane. 

We introduce now the following definition. 

DEFINITION II. A set of three diameters of a central quadric, such 
that the diametral plane of any one of them is the plane determined 
by the other two, is called a set of conjugate diameters; and a set of 
three diametral planes, such that any one of them is the diametral 
plane of the line of intersection of the other two, is called a set of con- 
jugate diametral planes. 

We shall find it convenient to refer to such sets as a conjugate 
set of diameters and a conjugate set of diametral planes. 
In terms of this definition we have the following theorem. 

THEOREM 2. For every diameter (diametral plane) of a central quad- 
ric there exist an infinite number of conjugate sets; for every pair of 
diameters (diametral planes), of which one lies in the diametral plane 
of the other (passes through the diameter of the other) there exists one 
conjugate set. For every diameter together with a diametral plane 
passing through it, there exists one set of conjugate diameters and 
one set of diametral planes, such that the diameters of the first set are 
the lines of intersection of the planes of the second set. 

Remark 1. The axes of symmetry of the ellipsoid furnish an 
example of a conjugate set of diameters. 

Remark 2. Any two of a conjugate set of diameters are " con- 
jugate diameters" of the conic section in which their plane cuts 
the quadric, where the words in quotation marks are to be under- 
stood in the sense in which they are used in Plane Analytical 
Geometry. 

The coordinates (a, 0, 7) of the point P in which the diameter d 
of direction cosines X, /i, v meets the surface, are proportional to 
X, ^, v] that is, a = Xs, = ps, 7 = vs. The tangent plane to the 
surface at this point may therefore be represented by the equation 
wiXz + m 2 M2/ + m&z = 1. Comparison with equation (2) shows 
that this plane is parallel to the diametral plane of d. 

And if P(a, 0, 7) is a point of the conic in which the diametral 
plane D of the diameter d cuts the surface, then m\a\ + w 2 ftu + 
= 0; and the tangent plane to the surface at this point is 



CONJUGATE DIAMETERS OF THE ELLIPSOID 271 

represented by the equation m\ax + m$y + m^yz 1. Con- 
sequently this tangent plane is parallel to the line d (compare 
Theorem 21, Chapter IV, Section 52, page 99); hence there exists 
also a tangent line through P parallel to d. If P moves along the 
curve of intersection of the surface with the plane J5, these tangent 
lines which are parallel to d generate a cylinder. 

DEFINITION III. The enveloping cylinder of a quadrfc surface cor- 
responding to a line d is the cylinder generated by the tangent lines to 
the surface which are parallel to d. 

We have therefore obtained moreover the following result. 

THEOREM 3. The tangent planes to a quadric surface at the points 
where it is met by a diameter d are parallel to the diametral plane of 
d. The tangent planes at the points where it is met by the diametral 
plane of d are parallel to d; the tangent lines drawn through these 
points and parallel to d form an enveloping cylinder of the surface. 

116. Exercises. 

1. Show that the enveloping cylinder of an ellipsoid is an elliptic cylinder 
for every direction of the generator. Obtain its equation. 

2. Show that the enveloping cylinder of an hyperboloid of two sheets is an 
hyperbolic cylinder for every direction of the generator; obtain its equation. 

3. Show that the enveloping cylinder of an hyperboloid of one sheet is an 
elliptic cylinder for some directions of the generator and an hyperbolic cylinder 
for other directions of the generator; obtain the equation of the enveloping 
cylinder. 

4. Determine the directions of the generator for which the enveloping cylin- 
der of an hyperboloid of one sheet will be an elliptic cylinder, and also the 
directions for which it will be an hyperbolic cylinder. 

X 2 ?/ 2 3 2 

6. Determine the diameter of the surface T + 77 ~~ Tp ^ * which is con- 
jugate to the diameters whose direction cosines are proportional to 2 : 1 : 2 
and 1 : 9:4 respectively; determine also the diametral planes of these 
diameters. 

?/ 2 z 2 

6. Solve the corresponding problem for the surface x 2 -f- 4- -h TT = 1 and 

the directions 1:2: 3 and 2:3: 1. 

7. Determine the enveloping cylinders of the ellipsoid in the preceding 
problem for the directions of each of the three conjugate diameters. 

8. Determine for the surface of Exercise 6 the diametral plane which is 
conjugate to the planes x -f- 2 y -\- 2 z = and 4 x -{- y z = 0. 

117. Conjugate Diameters of the Ellipsoid. We shall denote 
the length of the chord determined by the diameter d of a quadric 



272 PROPERTIES OF CENTRAL QUADRIC SURFACES 

surface by 2 5, the points where d meets the surface by E(a, 0, y) 
and #'(', jft', 7'); and the direction cosines of d by X, ju, v. If 
several diameters are under consideration at the same time, we 
shall distinguish between the numbers related to them by the use 
of subscripts. 
The parametric equations of d may be taken in the form 

(1) x = Xs, y = us, z = vs. 

The number 5 is the numerical value of the roots of the equation 
(2) (mjX 2 + msM 2 + m& z )s* = 1, 



provided these roots are real; and the coordinates of E and E' 
are obtained by substituting these roots for s in equations (1). 
It should be clear that the roots of equation (2) never coincide, 
that for the ellipsoid they are always real and finite, while for 
the hyperboloids of 1 or 2 sheets, they will be real and finite, 
infinite, or imaginary according as raiX 2 + m^f + m^v 2 is positive, 
zero, or negative. 

We will show now that the ellipsoid is the only quadric for which 
there exist sets of conjugate diameters, for each of which the roots 
of the equation (2) are real and finite. That such sets do exist for 
the ellipsoid follows from Theorem 2 (Section 115, page 270) in 
conjunction with the observation in the preceding paragraph. 
Suppose now that we have three diameters d\, c? 2 , and d 3 such that 

(3) miXi 2 + w 2 Mi 2 + ^3*v > 0, ?ttiX 2 2 + m 2 ju 2 2 + m^ > 0, 

WiX 3 2 + ?2M3 2 + 7ft 3 *> 3 2 > 0, and 

(4) WiX 2 X3 + M 2 M2/*3 + m^va = 0, WiX 3 Xi + w 2 M3Mi + ^aWi = 0, 

0. 



From the first two of equations (4) we derive equations (5) of 
Section 115; and from these we conclude that there exists a non- 
zero constant fc, such that 



Mi 



MI 



If we multiply the inequalities (3i) and (3 2 ) and subtract the 
square of equation (4 3 ) from the result, we obtain, by an easy re- 
arrangement of terms: 

0. 



CONJUGATE DIAMETERS OF THE ELLIPSOID 273 
The equations just preceding enable us to replace this inequality by 



and from this we conclude that 

2 my 2 



and finally on account of (3 3 ), that mim 2 m 3 > 0. Hence, either 
nil > 0, M 2 > and ra 3 > 0, in which case our objective has been 
reached, or else one of these numbers is positive, the other two 
negative. Let us suppose m\ > and m 2 < 0, m^ < 0. We 
derive then from (3i) and (3 2 ), the inequalities 

miXi 2 > m 2 /zi 2 m 8 j>i 2 > and WiX 2 2 > m 2/ u 2 2 m-pf > 0; 
multiplication of these inequalities leads to 

mi 2 Xi 2 X 2 2 > m 2 2 MiW + rn^sCMi V + M2 V) + infrfyf. 
If from the two sides of this inequality we subtract the equation 
?Wi 2 Xi 2 X 2 2 = ?n 2 ViW + 2 



obtained from (4 3 ) we reach the inequality 
> 



from which would follow that m 2 w 3 < and therefore that m 2 and 
7M 3 are opposite in sign; but this contradicts the supposition which 
we started from. We conclude therefore that mi > 0, w 2 > 
and m 3 > 0; and we have the following theorem. 

THEOREM 4. The ellipsoid is the only central quadric for which 
there exist sets of conjugate diameters each of which meets the surface 
in real Unite points. 

In developing further properties of the conjugate sets of diam- 
eters of the ellipsoid, we shall use equation (1), Section 115, with 

the understanding that mi = ^ , m z = ^ and m 3 = -^ 

If di, d*, and d 3 are conjugate diameters of the ellipsoid, we derive 
from the equations 

+ waM,- 2 + m&i*)6t* = 1 and m^X/ + mww + 
= 0, i,j = 1,2,3, ij, 



274 PROPERTIES OF CENTRAL QUADRIC SURFACES 

the interpretation that 

Xi6i v Wi, jui5i v mo, 

and 

are the direction eosines of three mutually perpendicular lines. 
If we apply to them the results obtained in Theorem 6, Chapter V, 
the Remark 3 following this theorem, and Theorem 7, Chapter V 
(see Section 65, pages 123 and 124), we find the following further 
relations : 

(5) 
(6) 



= 0, 

= 0, <5rViXi + <5 2 2 j> 2 X2 + <5 3 2 *> 3 X 3 = 0; 

3 2 ) = 1, 7/t 2 (iW+2W+3W) 

3 3 V) = \ ; 



(7) 



Mi 



From these formulas we derive the following interesting results. 

THEOREM 5. The sum of the squares of the semi-diameters of a set 
of conjugate diameters of an ellipsoid is the same for all conjugate 
sets of diameters. 

Proof. The semi-diameters of the conjugate set are 61, 5 2 , and 
5 3 . If we divide the three equations in (6) by Wi, w 2 , and m 3 
respectively and add the results, we find: 

5i 2 (Xi 2 + Mi 2 + "i 2 ) + <5 2 2 (X 2 2 + M 2 2 + 2 2 ) + 5 3 2 (X 3 2 + ^ + ^ 3 2 ) 



that is 



THEOREM 6. The volume of the parallelepiped of which the three 
semi-diameters of a conjugate set are coterminous edges is the same 
for all conjugate sets. 

Proof. To determine the required volume we have at our dis- 
posal the formula in Corollary 2 of Theorem 3, Chapter V (Section 



CONJUGATE DIAMETERS OF THE ELLIPSOID 



275 



62, page 118). In the present case, the symbols used in this 
formula have the following values: 

a = Si, b = S 2 , c = 5 3 , cos 7 = XiX 2 + Miju2 + 
XsXj. + MsMi + ^i; COS a = X 2 X 3 + M2Ms + ^2^3. 

If we make use now of formula (7) above and observe that 



cos ft = 



Xi Mi ' 
X 2 M2 ' 
Xs Ms * 

(compare part (3) of the Remark following Theorem 10, Chap- 
ter I, Section 14, page 27), we obtain for the volume of the par- 
allelopiped the expression 



= pqr. 



THEOREM 7. If the ellipsoid ~ + ~ -f ~ = 1 Is referred to a reference 
frame O-X'Y'Z', whose axes are the lines of a conjugate set of diam- 
eters, its equation is ~ -f ~ + f-r = 1, where 5i, 5 2 , and 5 3 are the semi- 

Oi" O2 W3 

diameters of the conjugate set. 

Proof. Let P(a y (3, 7) be an arbitrary point on the ellipsoid, 
as referred to the frame 0-XYZ, so that mia 2 + m 2 /3 2 + rn^y 2 = 1 ; 
the coordinates a 7 , /3 X , y' of P with respect to O-X'Y'Z' are the 
lengths of the segments P y > z 'Pj Pz'x'P, and P x yP of the lines 
through P parallel to OX', OF', and OZ' respectively (compare 
Fig. 14, page 113). The equations of the line through P parallel 
to OX' are 

The equation of the plane Y'OZ' is m{KiX + m^\y + m&iZ = 0; 
and the distance PP y v is the value of s determined by the equation 
which results when the expressions for x, y, and z just preceding 
are substituted in this equation, that is, by the equation 

s = 0. 



276 PROPERTIES OF CENTRAL QUADRIC SURFACES 
Therefore 



In similar manner we find 



Consequently we find 



03 



If the squares of the trinomials on the right side of this equation 
are expanded, we obtain a homogeneous polynomial of the second 
degree in <*, /3, and 7. The coefficient of a 2 is found to be 
?Hi 2 (Xi 2 Si 2 + X 2 2 5 2 2 + X 3 2 <5 3 2 ) = mi, by virtue of formula (61); in 
similar manner we find that the coefficients of /3 2 and 7 2 are w 2 
and m s respectively. For the coefficient of /3y we find 2 morris 
(Mivi5i 2 + jU2^ 2 5 2 2 + M3^5,3 2 ) = 0, on account of formula (5 2 ); and 
it should be a simple matter to verify that the coefficients of ya 
and aft are likewise zero. Hence we find 



Since a', 0', and 7' are the coordinates with respect to 0-X'Y'Z f 
of an arbitrary point on the ellipsoid, our theorem is proved. 

118. Exercises. 

1. Prove that a set of conjugate diameters of the hyperboloid of one sheet 
mix 2 + m> 2 y 2 -f w 3 2 2 = 1 (mi > 0, w 2 > 0, m 3 < 0) is also a conjugate set 
for the hyperboloid of two sheets mix 2 -\- m> 2 y 2 + msz 2 = 1. 

2. Prove that of a set of conjugate diameters of the two hyperboloids of the 
preceding exercise, two and only two meet one of these surfaces in real, finite 
points, whereas the third diameter of the set meets the other surface in real, 
finite points; and that this set of conjugate diameters for the two surfaces 
determines three finite chords, two chords of one of the surfaces, and one of the 
other. 

3. Prove that if 2 81, 2 5 2 , and 2 5 3 denote the lengths of the diameters of a 
conjugate set for the two hyperboloids of Exercise 1, then 5i 2 + 5 2 2 + 5s 2 = 

-L + -L+-L. 

nil wi* 7^3 



LINEAR FAMILIES OF QUADRICS 277 

v 2 z z 

4. Determine the diameter of the surfaces z 2 ^- + = 1, which is 

4 y 

conjugate to the two diameters whose direction cosines are proportional to 
1:6:9 and to 1 : 2 : 4 respectively. 

6. Prove that three tangent planes of an ellipsoid which are parallel to 
the three diametral planes of a conjugate set meet in a point. Determine the 
locus generated by this point when all conjugate sets of diametral planes are 
considered. 

6. Determine the length of the diameter conjugate to two diameters which 
lie in a plane of circular section through the center of an ellipsoid. 

119. Linear Families of Quadrics. When two quadric surfaces 
are given, as for instance by the equations Q(x, y y z) = and 
Q'(x, y, z) = 0, the points common to these surfaces determine 
a space locus with one degree of freedom, that is, a space curve. 
The study of space curves constitutes an important part of the 
field of Differential Geometry. Without knowing anything fur- 
ther about the character of the curve of intersection of two quadrics, 
we can affirm that, no matter what polynomials are represented 
by the symbols A(x, y, z) and B(x, y, z), the locus of the equation 
A(x, y, z) Q(x, y, z) + B(x, y, z) Q'(x, y, z) = 

will be a surface which passes through this curve; for this equation 
is surely satisfied by the coordinates of any point which belongs 
to the locus of Q = and to that of Q' = 0. And if A (x, y, z) 
and B(x, y, z) reduce to constants which are not both zero, the 
equation represents a quadric through this curve. 
Thus the equation 

(1) k&(x, y, z) + k 2 Q'(x, y, z) = 

represents, when the ratio ki : k< 2 varies, a family of quadric sur- 
faces, all of which pass through the points common to the surfaces 
Q = and Q f = 0. It is called a linear one-parameter family of 
quadrics, also a pencil of quadrics (compare Section 49). For 
fcj = o, fc 2 = 1, we obtain the surface Q'; for ki = 1, fc 2 = 0, 

the surface Q. Upon division by ki and putting ~ = X, the 

KI 

equation (1) takes the form 

(2) Q(x,y,z)-\Q'(x,y,z) = 0; 

and this equation is equivalent to (1), except that it does not in- 
clude the surface Q' for any finite value of X. 



278 PROPERTIES OF CENTRAL QUADRIC SURFACES 

The value of the parameter X may be so selected as to make the 
surface represented by (2) satisfy a single condition, for example, 
that of passing through one prescribed point, which does not 
lie on the surface Q f . This particular problem has a unique solu- 
tion. For if (a, ft 7) is the prescribed point, X must satisfy the 
condition 

Q(, ft 7) - AQ'(, ft 7) = 0, 
so that 

x = Q(a, ft 7) t 

The condition that a surface of the pencil be a surface of revolu- 
tion leads to an equation of higher degree in X; for it requires that 
X be determined so that the equation 



= 



shall have a double root (compare the Remark on page 215). 
We shall not pursue this problem further. 

Of special interest is the question as to the singular quadrics 
contained in the pencil of quadrics (2). A surface of this pencil 
will be singular if and only if the rank of the matrix 

Xon' 012 Xa^' 013 Xais' OH 

2 ~ ^ a 12' 022 ~ Xa 22 ' 023 X023' O24 



On Xon' k tti2 Xoi 2 ' a J3 

012 Xai 2 ' 022 Xa 22 ' k a 23 Xa 23 ' 

a 2 s Xa 2 s' 033 \(IM k 



/\ \ _ 

' 023 Xaas' o 33 Xo 33 ' a 34 



O24 XO 2 4' O 3 4 XO34' 044 ~ XO4 

is less than 4 (compare Definition V, Chapter VII, Section 82, 
page 166); it will be a non-degenerate singular quadric, that is, a 
proper cone or a cylinder, if and only if the rank of this matrix is 3. 
In either case a necessary condition is that the determinant 
A(X) of this matrix shall vanish. But the equation A(X) = 
is of the fourth degree in X; there will therefore be at most four 
singular quadrics in the pencil. A further discussion of the charac- 
teristics of these surfaces, of the questions whether they are de- 
generate surfaces, whether they are cylinders or cones leads into 
a more extensive treatment than we can undertake here. The 
interested reader is referred to Snyder and Sisam, Analytic Geom- 



LINEAR FAMILIES OF QUADRICS 279 

etry of Space, Chapter XI, or to Bocher, Introduction to Higher 
Algebr^, Chapter XIII, for a consideration of these problems. 

To determine the ratios of the ten coefficients which appear in 
the general equation Q(x, y, z) = of a quadric surface, nine 
points must be prescribed. Substitution of the coordinates of 
these nine points in the general equation furnishes nine linear 
homogeneous equations for an, a 22 , a., a 44 , a 12 , a 23 , a 3 4, i4, ai 3 , a 2 4 
from which the ratios of these coefficients can in general be de- 
termined. In order that this be possible, the 9 points (a,-, ft, 7,-), 
i = 1, 2, . . . , 9 must have such coordinates that the rank of the 
matrix, formed by the 9 rows 

(3) a,- 2 , ft 2 , 7i 2 , 2ft T ,-, 2 7 W, 2a;ft, 2 a*, 2ft, 2 7,', 1 

(i = 1, 2, . . . , 9) of 10 elements each, is 9. Whenever a set of 
points (ai, ft, 7,-) is so constituted that the matrix, in which for 
each of the points there is a row of 10 elements as indicated in (3), 
has a rank equal to the number of points in the set, we shall say 
that these points are in " general position." We can therefore 
say that there passes a single quadric surface through a set of 
nine points in " general position." 

If only eight points are given, we have eight equations; if the 
eight points are in general position, these equations will enable us 
to determine 8 of the coefficients a y - as linear homogeneous func- 
tions of the other two, by Cramer's rule; if these two be called k\ 
and fc 2 , the solution will be of the form Oy- = o/fci + a,/'/^, where 
a,/ and a y -" are known. The equation of the quadric through the 
given 8 points will then be Q(x t y, z) = 0, where Q(x, y, z) = 
k\Q'(x, y y z) + fc 2 Q"(z, y, z). The surface Q = will therefore 
belong to a pencil of quadrics determined by the quadrics Q' = 
and Q" = 0, provided the latter two surfaces are distinct. It 
will be worth our while to inquire further whether this will indeed 
be the case. If we suppose that the value of the determinant of 
8th order, which is formed by the first 8 elements in (3) for 
i = 1, 2, . . . , 8, is different from zero (let us denote this deter- 
minant by J5g), then we may take fci = a 3 4, fc 2 = a 44 ; and it follows 
from Cramer's rule that an is equal to a fraction whose denominator 
is equal to D 8 ; and the numerator is obtained from the denominator 
by writing the column 2 a^y* + Q<u(i =1,2,. . . , 8) in place of 
the column a,- 2 , i = 1, 2, . . . , 8. Hence if we denote by D 8 ' 



280 PROPERTIES OF CENTRAL QUADRIC SURFACES 

and Z) 8 " the determinants obtained from D 8 by replacing its first 
column by 2 7;, i = 1, 2, . . . , 8 and by a column of ones re- 

Ds D 8 " 

spectively, we find that an = -rr- X a 3 4 + -ry- X a 4 4 = kiQu + A; 2 an , 

D r D tf 

where an' = ~- and an" = -~- . Similarly we find 22 = &ia 22 ' 

+ fc 2 a 22 / ', where a 22 ' and a 22 " are obtained from Z> 8 by replacing 
the second column by the columns which were used in the first 
column for the formation of TV and AJ" respectively, and dividing 
by D 8 ; and so on for all the coefficients a# except a 34 and a 44 . 
Finally, if we write a 34 ' = 1, a 34 " = 0, and a 44 ' = 0, a^" = 1, we 
have also a 34 = k\a^ f + A; 2 a 34 ", and 44 = fcia 44 ' + & 2 a 44 ". Hence 
we can write Q(x, y, z) = kiQ'(x, y, z) + k 2 Q"(x, y, z), where the 
coefficients of Q 1 are not proportional to those of Q". It should 
be clear that the same conclusion will be reached if we start from 
the supposition that another 8th order determinant of the matrix 
is different from zero. Therefore any quadric Q which passes 
through the 8 given points belongs to the pencil of quadrics de- 
termined by the quadrics Q' and Q" and passes through the curve 
of intersection of these surfaces. 

THEOREM 8. All quadrics which have In common eight points In 
general position have a curve In common which passes through these 
eight points. 

A system of three quadrics Q(x, y, z) = 0, Q'(x, y, z) = and 
Q"(x, y,z) = which do not belong to one pencil, that is, such 
that the only values of fci, fe, and fc, for which the relation 
kQ(x 9 y, z) = kiQ'(x, y, z) + k 2 Q"(x, ?/, z) holds identically are ki = 
&2 = k 0, determines a linear two-parameter family of quadrics 
kQ + k,Q f + k 2 Q" = 0, or Q + XQ' + M Q" = 0. Such a family 
is called a bundle of quadrics. All surfaces of the bundle pass 
through the points common to the three initial surfaces. These 
equations will in general have 8 points in common, since their 
equations are of the second degree in x, y, z. But these 8 points 
are not in general position, since otherwise the three surfaces Q, 
Q' and Q"' would, in accordance with Theorem 8, belong to one 
pencil. 

Suppose now that there be given 7 points in general position; 
then seven of the coefficients a# can be determined in terms of the 



FOCAL CURVES AND DIRECTRIX CYLINDERS 281 

remaining three; if these be called fci, fc 2 , and fc 3 , we shall find 
dij = k]fLi/ + k^dij" + kzdij" ', and therefore Q(x y y, z) = k\Q f 
(x, y, z) + k<2,Q"(X) y, z) + kzQ'"(x, y, z). The determination of 
the functions Q'> Q", and Q'" proceeds by a method analogous 
to the one used in the proof of Theorem 8. On the supposition 
that the determinant of 7th order whose rows are formed by the 
first seven elements of (3) for i = 1, 2, . . . , 7 does not vanish, 
we may take ki = a 2 4, fe = a 3 4, fc* = d^. Therefore if a# = k\(ii/ + 
kidij" + k^dij" , we have d^ = 1, #24" = 0, #24"' = 0; d%t 0, 
a 34 " = 1, a34 '" = 0; a 4 / = 0, a 44 " = 0, a 44 '" = 1. This shows 
that the three surfaces Q' = 0, Q" = 0, and Q!" = do not 
belong to the same pencil. Therefore any surface through the 7 
given points belongs to the bundle of quadrics determined by these 
three surfaces and will therefore pass through the eight points 
common to them. 

THEOREM 9. All quadrics which have In common 7 points In general 
position also have an eighth point in common. 

120. Focal Curves and Directrix Cylinders of Central Quadrics. 

In Plane Analytical Geometry, the focus and directrix of a conic 
section are defined as a point and a line such that the ratio of the 
distances of any point on the curve from the focus and from the 
directrix is the same as for any other point on the curve. A 
corresponding definition of focus and directrix for the quadric 
surfaces seems to have been given first by Chasles (in 1835) or 
perhaps by Salmon (in 1862). 

DEFINITION IV. A foeus and a directrix of a quadric surface are a 
point and a line such that the ratio of the square of the distance of 
any point P on the surface from the focus to the product of its dis- 
tances from any two planes through the directrix is constant.* 

Although the concepts defined in this manner are entirely defi- 
nite, it is still an open question whether quadric surfaces possess 
foci and directrices. We shall not deal with this question for the 
general quadric but only for the central quadrics, which are not 
surfaces of revolution. 

* The planes through the directrix need not be real. The line of inter- 
section of two planes may be real even though the planes themselves are not 
real; for example, the planes x -f iy z = and x iy z = both contain 
the line x = t, y = 0, z = t. 



282 PROPERTIES OF CENTRAL QUADR1C SURFACES 
If F(a, /3, 7) is a focus of the central quadric 
(1) mix 2 + w 2 7/ 2 + nitf 2 = 1 



(mi, w 2 , and m 3 being distinct and all different from zero), there 
must exist two linear functions l(x, y, z) and /'(#, y> z), and a 
constant c such that 



(x - a) 2 + (y - 0)* + (z - 7) 2 = c X Z(a?, y, 0) X Z'fo 

for every set of values of x, y, z which satisfy equation (1). 

Consequently the functions mix' 2 + m 2 y 2 + m& 2 1 and (x a) 2 
+ (y ~~ $Y + (^ T) 2 cZZ' must differ by a constant factor X; 
that is, there must exist a factor X such that 

niiX 2 + m 2 ?/ 2 + m&* - 1 - \[(x - a) 2 + (y - 0) 2 + (2 - 7)'] 
= \cl(x, y, z)l'(x, y, z). 

Therefore the function on the left-hand side of this identity must 
be equal to the product of two factors which are linear in x, y, and 
z and the directrix which corresponds to the focus F(a, /3, 7) will 
then be the line of intersection of the planes represented by the 
equations which result when these linear factors are equated to 
zero. According to Theorem 6, Chapter VIII (Section 96, page 
206) the necessary and sufficient condition for the possibility of 
thus factoring the function is that the rank of the discriminant 
matrix of the quadric surface 

mix* + 2 */ 2 + m*z* - 1 - \[(x - a) 2 + (y - 0) 2 + (z - 7) 2 ] = 

is less than 3. The necessary and sufficient conditions which 
will make certain that the point F(a, 0, 7) is a focus of the quadric 
Q are therefore that the coordinates a, 0, 7 cause the determinant 

i X 00 Xa 

W2-X X0 

w 3 -X X7 



and all its three-rowed minors to vanish. But because the matrix 
of this determinant is symmetric, the second condition may be 
replaced, in view of Theorem 6, Chapter II (Section 26, page 43), 
by the requirement that the principal three-rowed minors vanish. 
If the principal three-rowed minor in the upper left-hand corner 



FOCAL CURVES AND DIRECTRIX CYLINDERS 283 



is to vanish, we must have (mi X) (m 2 X) (m 3 X) = 0, 
that is, X = mi, or X = m 2 , or X = m 3 . 

Let us consider the case X = mi. The principal minor formed 
from the 1st, 2nd, and 4th columns and rows then becomes 



m 2 mi m\& = mi 2 a' 2 (m 2 mi). 



mifl -[l+m 1 (a 2 +^ 2 +7 2 )] 

Since mi ^p and m 2 ^ m } , the vanishing of this determinant 
requires that = 0; and when X = mi and a 0, the fourth- 
order determinant and the three-rowed principal minor formed 
from the 1st, 3rd, and 4th columns and rows obviously vanish 
as well. There remains therefore the condition that the three- 
rowed principal minor in the lower right-hand corner shall vanish, 
that is, that 

m 2 mi nii/3 

m 3 mi 

-[i- 



0. 



If we develop this determinant, we are led to the equation 

(m 3 Wi)0 2 + WiW 3 (w 2 m\)y 2 + (ra 2 Wi) (m 3 Wi) = 0; we 
write this equation in the form 



(2) 



J_ 

mi 






+ 1 = 0. 



This equation, in conjunction with the equation a = 0, determines 
for varying /3 and 7 a real or imaginary conic section in the YZ- 
plane; it is called a focal conic of the quadric surface. 
If, conversely, (a, 0, 7) is a point on this conic section, 

mix* + m 2 y* + m 3 z 2 - 1 - mjx 2 + (y - 0) 2 + (z - y) 2 ] 

= (ra 2 mi)?/ 2 + (ra 3 mi)2J 2 + 2 mifty + 2 mi72 mi/3 2 






[(m 2 - mi)y + mi/3] 2 



1 



m 3 mi 



[(m 3 



and this function is factorable into two factors linear in y and z 
(the variable x is absent), with coefficients which are real or imagi- 



284 PROPERTIES OF CENTRAL QUADRIC SURFACES 

nary according as m 2 mi and m 3 mi are of opposite signs or 
of like signs. In either case, the equations which are obtained by 
equating these factors to zero represent planes which pass through 
the line determined by the equations 

(3) (m 2 m\)y + miff and (m 3 nii)z + m\y = 0. 



This line and the point (a, 0, 7) on the conic section determined 
above arc therefore a directrix and a focus of the quadric surface 
(1). As the point (a, 0, 7) describes the conic (2), the line (3), 
which is parallel to the X-axis describes a cylindrical surface, 
whose equation is obtained when and 7 are eliminated from 
equations (2) and (3). This cylindrical surface is called a direc- 
trix cylinder of the quadric. We conclude therefore that from 
the value X = mi, we obtain the focal conic 

(4) - + -^-- + 1 = 0, x = 0, 



and the directrix cylinder 

77" 

(5) m 2 (/ni w 2 ) 

i /LI ifi/i 

The reader should have no difficulty in showing that from X = w 2 
and X = m 3 , we obtain the focal conies represented respectively 
by the pairs of equations 

(6) -^- + -~- +1-0, y = 0, 



and 

x 2 ?/ 2 

(7) j-+ i : j- + 1=0, z = 0; 

ma mi m 3 m 2 

and the corresponding directrix cylinders given respectively by 
the equations 

z 2 x 2 

(8) m 3 (r?i 2 m 3 ) + mi(m 2 mi) = 1, 



FOCAL CON1CS AND DIRECTRIX CYLINDERS 285 

and 

(9) Wi(w 3 wO h w 2 (w 3 w 2 ) = 1. 

The results of this discussion are summarized in the theorem 
which follows. 

THEOREM 10. For every central quadric surface which is not a sur- 
face of revolution there exist three real or imaginary focal conies and 
three corresponding directrix cylinders; with every point on a focal 
conic there is associated a generating line of a directrix cylinder, so 
that point and line are focus and directrix of the quadric, as defined in 
Definition IV. 

121. Focal Conies and Directrix Cylinders, continued. We turn 
now to a further consideration of the focal conies and directrix cyl- 
inders for each type of central quadric, particularly with a view to 
determining the conditions under which they are real. 
CASE I. Ellipsoid. 

If we take the equation in the standard form 

71 + ^2 + ^2 = 1, P < q < r, 

u 1 1 j 1 , 1 1 o 2 

we have mi = , w 2 = -r, and w 3 = ; and = p- q 2 

p 2 q 2 r 2 m\ w 2 

< 0, = p 2 r 2 < 0. Hence equations (4) and (5) of 

the preceding section give the real focal ellipse 



V \Jy o i) O 9 -*-> 

q 2 p 2 r 2 p 2 
and the corresponding real elliptic directrix cylinder 

Equations (6) and (8) lead to the real focal hyperbola 

z 2 x 2 

and the associated real hyperbolic directrix cylinder 



286 PROPERTIES OF CENTRAL QUADRIC SURFACES 

Finally, equations (7) and (9) lead to an imaginary focal ellipse 
and an imaginary elliptic directrix cylinder. We reach therefore 
the following conclusion: 

THEOREM 11. An ellipsoid possesses a real focal ellipse in the prin- 
cipal plane determined by the two longer semi-axes, and a real focal 
hyperbola in the principal plane determined by the two extreme semi- 
axes. The associated directrix cylinders are real, elliptic and hyper- 
bolic respectively, their generators being perpendicular to the planes 
of the corresponding focal curves. 

CASE II. Hyperboloid of One Sheet. 
With the standard equation in the form 

~2 9.2 ~2 

-2 + ^ -~2 = 1 P<9> 
p2 q2 r 2 r *7 

we find 

mi = - 2 , m 2 = -, m 3 = -- 2 - 

From equations (4) and (5) of Section 120, we obtain the real 
focal hyperbola 

7/ 2 ? 2 

r _ n y z = i 

x " u ' q* - p 2 p 2 + r 2 ' 
and the real hyperbolic directrix cylinder 



Equations (7) and (9) give the real focal ellipse 

fy C\ _J J 1 

& U, ; n ^T 01 O * 

p 2 + r 2 q 2 + r 2 
and the real elliptic directrix cylinder 



The loci determined by equations (6) and (8) are imaginary in this 
case. The following theorem states the results. 

THEOREM 12. An hyperboloid of one sheet possesses a real focal 
ellipse in the principal plane which cuts the surface in an ellipse, and 
a real focal hyperbola in the principal plane determined by the con- 
Jugate axis and the longer of the two transverse semi-axes. The as- 
sociated directrix cylinders are real, elliptic and hyperbolic respectively , 
with generators perpendicular to the planes of the corresponding focal 
curves. 



FOCAL CONICS AND DIRECTRIX CYLINDERS 287 

CASE III. Hyperboloid of Two Sheets. 
The standard form of the equation 

~2 7 /2 ~2 

x v _ z = i, q<r 

p2 q2 r 2 

gives 

1 1 1 

mi = 5, m 2 - - -, w 3 = - ^- 

Equations (6) and (8) yield the real focal hyperbola 

__^ ___ z^___ 
2/ u > p* + q2 r 2 - q 2 ' 

and the real hyperbolic directrix cylinder 



From equations (7) and (9) we obtain the real focal ellipse 

/v2 /i2 

^ ^J O i > I n i> J- 

p 2 + T I r 2 q* 
and the real elliptic directrix cylinder 



The loci determined by equations (4) and (5) are imaginary in 
this case. The conclusions are therefore as stated in the next 
theorem. 

THEOREM 13. An hyperboloid of two sheets possesses a real focal 
ellipse In the principal plane determined by the transverse axis and 
the shorter of the two conjugate semi-axes, and a real focal hyperbola 
in the principal plane determined by the transverse axis and the 
longer conjugate semi-axis; the associated directrix cylinders are real, 
elliptic and hyperbolic respectively, with generators perpendicular 
to the planes of the corresponding focal curves. 

122. Exercises. 

1. Determine the focal curves and the directrix cylinders of each of the 
following surfaces: 

, , x 2 2/ a z 2 ... x 2 ?y 2 z 2 ' 

(a) -9+7+6-1. (6) - + ---=1, 

M ?! - nl _ l 2 - i 

(c; 6 9 4 



288 PROPERTIES OF CENTRAL QUADRIC SURFACES 

2. Prove that an ellipsoid of revolution has a real focal circle in the prin- 
cipal plane which cuts the surface in a circle and a real circular directrix 
cylinder. 

3. Prove that an hyperboloid of revolution of one sheet has a real focal 
circle and a real circular directrix cylinder. 

4. Prove that the focal ellipse of an ellipsoid is similar to the ellipse in which 
the surface is cut by the plane of the focal ellipse if and only if the surface is 
an ellipsoid of revolution. 

6. Prove that the focal ellipse of an hyperboloid of one sheet is similar to 
the ellipse in which the surface is cut by the plane of the focal ellipse if and 
only if the hyperboloid is a surface of revolution. 

6. Prove that the principal plane determined by the two shorter semi-axes 
of an ellipsoid cuts the focal ellipse and the associated directrix cylinder in a 
point and a line respectively which are the focus and the directrix of the ellipse 
in which the surface is cut by this plane. 

7. Prove that the semi-axes of the ellipse in which an ellipsoid is cut by the 
plane of the focal ellipse are mean proportionals between the corresponding 
semi-axes of the focal ellipse and of the directrix curve of the associated elliptic 
directrix cylinder; also that the semi-axes of the ellipse in which an ellipsoid 
is cut by the plane of the focal hyperbola are mean proportionals between the 
corresponding semi-axes of the focal hyperbola and of the associated hyperbolic 
directrix cylinder. 

8. Prove theorems analogous to those of the preceding exercise for the 
hyperboloid of one sheet and for the hyperboloid of two sheets. 

9. Determine the distance from the origin of the points in which the focal 
hyperbola of an ellipsoid is met by the planes of central circular section. 

10. Determine the conditions under which the focal hyperbola of a central 
quadric is a rectangular hyperbola. 

11. Prove that the foci of the focal curves of an ellipsoid coincide with 
the foci of the conic sections in which the surface is cut by the planes of these 
focal curves. 

12. Prove the corresponding theorem for the hyperboloids of one and two 
sheets. 

123. Confocal Quadric Surfaces. Elliptic Coordinates. It fol- 
lows from formulas (4), (6), and (7) of Section 120 (see page 284) 
that two central quadrics 

= 1 and w/z 2 + m^y 2 + m 3 '2 2 = 1 

will have their focal curves in common if and only if = 

nii ntj 

/ / > for i> 1 = 1> 2, 3. 

m^ m/ 

This will certainly be the case therefore for all surfaces repre- 



CONFOCAL QUADRIC SURFACES 289 



sented by the equation 

~2 ,,,2 



in which X is a real parameter. For all these surfaces the focal 
ellipse is given by the equations 

+ 



;L ~~^ rt 

p 2 r 2 p 2 ' 

and the focal hyperbola by the equations 

z 2 x 2 

y = 0, - - r -- - - = 1. 
r 2 q 2 q 2 p 2 

The family of surfaces represented by equation (1) is called a 
confocal family of quadric surfaces. 

If X < p 2 , all the denominators in equation (1) are positive; 
the surface is therefore an ellipsoid. If p 2 < X < g 2 , the first 
denominator is negative, the other two are positive, so that the 
surface is an hyperboloid of one sheet, of which the X-axis is the 
conjugate axis, If q 2 < X < r 2 , the first two denominators are 
negative and the third one is positive; hence the surface is an 
hyperboloid of two sheets, of which the Z-axis is the transverse 
axis. Finally if X > r 2 , the surface is an imaginary ellipsoid. 

For the critical values X = p 2 , X = g 2 , and X = r 2 , the equation 
(1) has no meaning. If we multiply both sides of equation (1) 
by p 2 X, we obtain the equation 



which is equivalent to (1) except when X = p 2 . For this value of 
X, equation (2) reduces to x 2 = 0, whose locus is the FZ-plane 
counted doubly. If equation (1) is multiplied through by q 2 X 
and by r 2 X, we obtain equations, which for X = q 2 and X = r 2 
reduce respectively to the equations y 2 = and z 2 = 0. We 
complete now the definition of the confocal family of quadrics 
given by equation (1) by the statement that to the values X = p 2 , 
X = g 2 , and X = r 2 shall correspond the FZ-plane, thq ZX-plane, 
and the .XT-plane respectively, each counted doubly. The 
character of the surfaces in the confocal family is indicated dia- 
grammatically in Figure 36. 



290 PROPERTIES OF CENTRAL QUADRIC SURFACES 

We shall now try to determine in what manner the surfaces of 
the family change as X tends towards the critical values, passing 
through values which remain steadily on one side of a critical 
value. To indicate that X tends toward p 2 through values which 
are greater than p 2 , we shall write X > p 2 + 0; to indicate that 
X tends toward p 2 through values which are less than p 2 , we shall 
write X p 2 0. Similar meanings are to be attributed to the 
notations X - q 2 + 0, X -> q 2 - 0, X - r 2 + and X - r 2 - 0. 

,Q p* 0* r^ a-axis 

KlHpsoid Hyperboloid Hyperboloid Imaginary KUipsoid 

I of 1 sheet I of 2 sheets I 



Pu 
N 



FIG. 30 



As X-p 2 0, the surface is steadily an ellipsoid; its semi- 
axis along the Z-axis tends to zero. Since X < p 2 , the factor 
p 2 X in the second term of equation (2) is positive; it should 

y 2 z 2 

therefore be clear from this equation that -~r + - - - 1< 0. 

q 2 X r 2 X 

But points in the FZ-plane for which this inequality holds lie 

y 2 z 2 

on the inside of the ellipse 9 x + -5 - - = 1.* Hence as 

^ q 2 X r 2 X 

X > p 2 0, the surface tends toward those points of the YZ- 
plane which lie on the inside of the focal ellipse 



i 
T~ 



~ , 

q 2 p 2 r 

* To be convinced of this fact, it is only necessary to observe that at the 

v 2 z 2 

origin, the function -~ r + -5 - - 1 reduces to 1 and that since the 

C[ A T ~~~ A 

function is a continuous function of X, y, and z for all values of X which differ 
from q 2 and r 2 , it cannot change from negative values to positive values without 
becoming zero. Since this can take place only on the ellipse, points for which 
the function is negative lie on the same side of the curve as the origin, that is, 
on the inside of the ellipse; and points for which it is positive lie on the out- 
side of the ellipse. 



CONFOCAL QUADRIC SURFACES 291 

As X p 2 + 0, the surface is steadily an hyperboloid of one sheet, 
whose semi-axis along the X-axis tends toward zero. But now 

7/2 pj" 

p 2 X is negative, and therefore -=-^ - + -5 - - 1 > 0. 

q X T A 

Therefore as X > p 2 + 0, the surface tends toward those points 
of the FZ-plane which lie outside the focal ellipse. 

As X > q 2 0, the surface is always an hyperboloid of one sheet, 
whose semi-axis along the F-axis tends to zero. For those values 
of X, q 2 X is positive; it follows then from an equation analogous 

x 1 z 2 

to (2) that 2 _ + 2 _ \ 1 is negative. An argument 

similar to the one made in the footnote on the preceding page 
shows that points for which this inequality holds lie on the same 

x 2 z 2 

side of the hyperbola -r - - + -= - - = 1 as the origin. If we 
p 2 X r 2 X 

call this the inside of the hyperbola, we conclude that as X g 2 0, 
the surface tends toward the points of the ZJST-plane which lie 
on the inside of the focal hyperbola 

n * X * i 

ffl II _ _ I 

y v, t> n 4> i. 
r 2 q 2 q 2 p 2 

And the same reasoning shows that as X > q 2 + 0, the surface 
tends toward the points of the ZX-plane which lie outside the 
focal hyperbola. 

Finally, as X r 2 0, the surface is an hyperboloid of two 
sheets, whose semi-axis along the Z-axis (that is, the transverse axis) 

x 2 y 2 

tends to zero. Since r 2 X > 0, it follows that - - + -~ r 

*p A q X 

1 < 0; but now X > p 2 and X > q 2 and therefore this inequality 
is satisfied by all points in the -XT-plane. Consequently as 
X r 2 0, the surface tends toward the entire .XT-plane. 
We summarize the discussion by a theorem. 






THEOREM 14. The equation * , + ~~ - + ~ - = 1, p < q < r, 

p* A q* A r* A 

in which X Is a real parameter, represents a confocal family of quadrlc 
surfaces. As X Increases from negative Infinity to p 2 , t.he locus of 
the equation Is an ellipsoid which tends toward the inside of the 
focal ellipse of the family; as X increases from p 2 to q\ the locus is 
an hyperboloid of one sheet tending from the outside of the focal 
ellipse to the inside of the focal hyperbola of the system; as X increases 



292 PROPERTIES OF CENTRAL QUADRIC SURFACES 

from q 2 to r 2 , the locus is an hyperboloid of two sheets, which tends 
from the outside of the focal hyperbola to the entire XY-plane. For 
x = p 2, x = q% an( | x = r % the locus is respectively the YZ-plane, the 
ZX-plane, and the XY-plane, each counted doubly. 

We shall now prove two properties of confocal families of 
quadrics. 

THEOREM 15. Through every point In space that does not lie on one 
of the coordinate planes, there pass three surfaces of every confocal 
family of quadrics, namely, an ellipsoid, an hyperboloid of one sheet 
and an hyperboloid of two sheets. 

Proof. Let P(a, 0, 7) be an arbitrary point of space that does 
not lie on any coordinate plane; then a, 0, and 7 are all different 
from zero. If P is to lie on a surface of the confocal family repre- 
sented by equation (1), the parameter X must be so determined that 

a 2 6 2 v 2 

__rL_4_ p i T _ i = n- 

p 2 - \ ^ q 2 - \ ^ r 2 - X 
that is, X must be a root of the equation 

F(\) = (X - p 2 )(X - ? 2 )(X - r 2 ) + (X - <? 2 )(X - r 2 ) a 2 + (X - r 2 ) 
(X - pW + (X - p 2 ) (X - <? 2 )7 2 = 0. 

This is a cubic equation in which the coefficient of X 3 is +1; 
consequently for large positive values of X, F(\) > and for large 
negative values of X, F(\) < 0. Moreover 

= (q 2 -r 2 ) (q 2 -p 2 )P 2 <0; 



The graph of the function F(\) will therefore have the general 
character indicated in Fig. 37.* And from it we conclude that 
the equation F(\) = has three real roots, Xi, X 2 , and X 3 . Hence 
there are three surfaces of the confocal family which pass through 
the given point P(a, 0, 7). But we observe also from Fig. 37 
that Xi < p 2 , p 2 < X 2 < <? 2 , and (f < X 3 < r 2 ; therefore, in virtue 
of Theorem 14, one of these surfaces is an ellipsoid, one an hy- 
perboloid of one sheet, and one an hyperboloid of two sheets. 

* We are here assuming that the polynomial F(X) is a continuous function 
of X, as in the argument in the footnote on page 290 we assumed that a rational 
function is continuous except for a finite number of values of the independent 
variable. A satisfactory proof of these facts is found in treatises on the 
Theory of Functions of a Real Variable. 



CONFOCAL QUADRIC SURFACES 



293 



If a = 0, the root Xi becomes equal to p 2 , so that in place of the 
ellipsoid we have the FZ-plane counted doubly; similarly, if 
ft = 0, the hyperboloid of one sheet is replaced by the double 
ZX-plane, and if 7 = 0, the hyperboloid of two sheets is replaced 
by the double XY-plane. 

Fa) 




FIG. 37 

COROLLARY. If P(, p, 7) lies on one or more of the coordinate planes, 
it is still true that three surfaces of every confocal family pass through 
P; but one or more of the central quadrics of the family are then re- 
placed by the coordinate planes in which P lies. 

THEOREM 16. The three quadrics of a confocal family which pass 
through an arbitrary point P(, p, 7) in space are mutually orthogonal 
at P. 

Proof. Suppose first that P(a, 0, 7) does not lie on any co- 
ordinate plane. Then the three quadrics of the confocal family 
(1) which pass through P have the equations 

i = 1,2,3, 



P' 2 ~ X, 

where Xi, X 2 , and X 3 are the roots of the equation F(\) = 0, dis- 
cussed above. The equations of the tangent planes to these 
surfaces at the point P are 

ax . fly 



(3) 



yz 



= 1, i = 1,2,3. 



- X f - ' q* - X,- ' r 2 - X,- 
Since P lies on each of the three surfaces, we have moreover 



X,- q~ X f - 



r 2 _ 



= 1, i = 1, 2, 3. 



294 PROPERTIES OF CENTRAL QUADRIC SURFACES 

If we subtract any two of the last three equations from each other, 
we find 



___ 1 

- \i g - Xj 



- X,- p' - X, 

' U- 1,2, 3; , 



A simple reduction transforms these three equations to the fol- 
lowing form: 



But since X,- 4= X/, we conclude from this last equation that 

cP_ , ^ - T; __ n 

(p 2 X,-) (p 2 X;) (q~ X,-) (g 2 X,-) (r 2 X,-) (r 2 X,-) 

And this equation expresses the fact that any two of the tangent 
planes represented by equations (3) are perpendicular to each 
other (compare the Corollary of Theorem 9, Chapter IV, Section 
46, page 82). 

If P lies in a coordinate plane, one of the quadrics of the family 
which pass through it is that plane itself. Let us suppose that 
a = 0; then the surface in question is the FZ-plane counted 
doubly. And let the equation 

A. 2 72 2 

I y I ^ ^ 

p~ \z q 2 \z T 2 ~~ Xa 

be one of the non-degenerate quadrics passing through P; the 
tangent plane to this surface at the point (0, 0, 7) is represented 
by the equation 

@y a. ** z i 

This plane is parallel to the X-axis and therefore perpendicular to 
the double FZ-plane. If P lies on a coordinate axis, two of the 
quadrics degenerate into double coordinate planes; and these are 
surely perpendicular. Our theorem has therefore been proved. 

* * * 

From Theorems 15 and 16, it follows that the quadric surfaces 
of a confocal family cover the whole of space with a network of 



ELLIPTIC SPACE COORDINATES 295 

mutually perpendicular surfaces. To each of these surfaces a 
number is attached, namely, the value of the parameter X to which 
it corresponds; and for every point P in space there are three such 
numbers. These numbers are called the elliptic space coordinates 
of the point P. Our discussion has therefore shown that every 

x 2 ?/ 2 2 2 
ellipsoid -5 + H ^ ^ niay be made the basis of a system of 

elliptic space coordinates. We have obtained a frame of reference 
which generalizes in a remarkable way the Cartesian frames of 
reference with which we began our study of Solid Analytical 
Geometry in Chapter III. 

And this return to our starting point provides a suitable stop- 
ping point, ending in the key in which we began. 

In our journey through this book we have examined a few ques- 
tions in some detail and we have had a glimpse of many things 
which lay outside our path. It is the author's hope that the reader 
may have learned to appreciate the beauty and the power of the 
theory of determinants and matrices, and that he may experience 
the desire not only to continue the study of the subject to which this 
book is primarily devoted, but also to enter some of the fields, 
such as Projective Geometry and the Theory of Functions of a 
Real Variable, to which we have had occasion to allude now and 
then in the course of our work. 



APPENDIX 



(Compare Section 84, page 171) 
To prove: If X Is eliminated from the equations 

Li = 1 [\Qi(, ft 7) + M<M, ft 7) + *<?(i ft 7)] = 
and Lo = ttnX 2 -f a 22 M 2 -f ass** 2 -H 2 azzjuip -f 2 a 3 n/X 4- 2 a^A/* = Of 

the resulting quadratic equation In /* and v is 



On 012 



<?. 
<?* 





Oil 



023 



<?1 

<?3 





an 



<?3 



Proof. To simplify the writing we shall treat this problem in a slightly 
modified form. The given equations are clearly equivalent to the non-homo- 
geneous equations ax + by + c = and p\\x z + /?22?/ 2 + Pss H~ 2 p 2 s^ -f- 



= obtained by writing x and y in place of - and - respectively, 

and using general coefficients. We assume now that a =t= and solve the 
linear equation for x in terms of y; substitution of the result in the second 
degree equation leads to the following quadratic in y: 

pii(fy/4-c) 2 -2 puay(by+c)+a 2 pMy 2 -2 p 13 a(by+c)+2 7> 23 a 2 

upon reduction this equation becomes 

(pub* 2 p u ab + z>2 2 a 2 )?/ 2 -f 2 (p n bc - p i2 ac - p n ab -f 
2 p^ac -f Pasa 2 = 0. 

Direct expansion of the third order determinants shows that the coefficients 
differ in sign only from the respective determinants: 



Pii 



Pit 

?>22 

b 



Pn 
Pis 
a 



Pi2 a 

P'2S C 

b 



and 



Pn 
Pi* 
a 



Pl3 
P33 
C 



If we now return to the homogeneous form of the given equations and to the 
coefficients as given, we have completed the proof. 

II 

(Compare Section 84, page 174) 
To prove: If A*(Q) = and ^ 22 (^) = A 23 (Q) = A*(Q) = 0, then every 



third order, minor of the determinant 



a u 



a 22 023 
a 2 s ass 



vanishes; 



A,(Q) designates the value of this determinant and Aij(Q) the value 
of the cofactor of the element aij. 

296 



APPENDIX 297 

Proof. We shall use the notation Qi to designate the cofactors of the ele- 
ments Qi in the determinant. Since A$(Q) = 0, it follows from the Corollary 
of Theorem 5, Chapter II (Section 26, page 43) that A u (Q)An(Q) - A 12 Z (Q) 
= and Aii(Q)Aw(Q) A\^(Q} = 0, so that the hypothesis leads at once to 
the result that Aw(Q) = A\*(Q) 0. In the same way we find that ^4 22^44 
Q 2 2 = 0_ and ^33^44 Q 3 2 = 0, so that also Q 2 -^ Q* = 0. Moreover 
QiQi + QzQz -f QaQa = At(Q) 0; and therefore, since we are working on 
the hypothesis that Qi =J= (compare page 171, opening paragraph of Case I), 
it follows that Qi = 0. Finally we observe that, in virtue of ThftormrMS, 
Chapter I_(Section 7, page 13), QiAn + QzAiz + QaAu - and auQ { -f- 
aizQz + auQz 4- QiAu = 0; and from these equations we conclude that 
An = Au = 0. This completes the proof of our statement. 

Ill 

(Compare Section 87, page 185) 
To prove: q(a ls , 23, #33) = 0, If A u = 0. 

Proof. Here aij are the cofactors of the elements 0$ in the third order 
determinant ^.44 and q(x, y, z) is the homogeneous function of the second 
degree introduced on page 159. By the use of Euler's theorem on homogeneous 
functions (see footnote on page 161), we find 

2 <K13, 23, 33) = Ofl3tfl(<*13i 23, Ot^) + a^falS, 23, 3s) + 33<?s(13, 23, 3s) ,* 

and 

, <x 2 3, ass) = 2 (anaw -f ai 2 a 23 + flisass) = 0, 

, 23, ^33) = 2 (ai 2 13 + 22a 23 + 2333) = 0, 
<7 3 (13, 23, ^33) = 2 (ai313 + 02323 + 3333) = ^44 = 0, 

by Theorems 13 and 12, Chapter I (Section 7, page 13). 

IV 

(Compare Section 94, page 205) 
To prove: If r 4 = 1, then r 4 ' <2. 

Proof. Here r 4 is the rank of the discriminant matrix of the quadric surface 
Q and r 4 ' is the rank of the discriminant matrix of the equation Q'(x', y', z') 
obtained by rotation of axes from the equation Q(x, y, z) = 0. 

If r 4 = 1, A, D 3 , and D 2 vanish and therefore, by Theorem 4, Chapter VIII and 
its Corollary (Section 94, pages 203 and 204), A r = /V = /V = 0. It follows that 
every three-rowed minor of A' vanishes and that the sum of the principal two- 
rowed minors also vanishes. Since the three-rowed principal minors are them- 
selves symmetric, we can apply to each of them Theorem 7, Chapter II (Section 
26, page 44) ; hence the two-rowed principal minors of any one of the four three- 
rowed principal minors are of like signs, and since any two of these three-rowed 
principal minors have a two-rowed principal minor in common, all the prin- 
cipal two-rowed minors have the same sign. It follows then from the fact 
that ZV = that every two-rowed principal minor of A ; vanishes. Now we 



298 



APPENDIX 



apply Theorem 6, Chapter II (Section 26, page 43) to each of the three-rowed 
principal minors; and we conclude that every two-rowed minor of A' which is 
also a minor of a three-rowed principal minor must vanish. It remains to 
consider the two-rowed minors of A' which do not occur in any three-rowed 
principal minor; the only ones of this kind are the minors 



014 



flu 

34' 



and 



an' 
043' 



which do not have any element of the principal diagonal of A'. To show that 
these minors vanish also, we consider the three-rowed minors: 



a,/ 
024' 
034' 



and 



a 23 



a 44 



These determinants vanish and all their two-rowed minors vanish, except 
possibly the minors with which we are concerned; and of these, two occur as 
minors in each of the two three-rowed minors. If we write down the de- 
velopments of these determinants according to their last rows, we can conclude 
that the first of these two-rowed minors also vanishes. In a similar way, 
consideration of the pairs of three-rowed minors /I 2 i', A\i, and An', A 2 i' 
shows that the remaining two-rowed minors vanish. This completes the 
proof of the proposition. 



(Compare Section 96, page 206) 
To prove: The determinant 



2a<u 

abi -f- a\b 
aci -f- a\c 
CM /i + a ^/ 


abi + a^ 
2 661 
bci + 6iC 
6r/! -f- bid 


aci + ic 
6ci -f bic 
2cc, 

C/i -f Cid 


atfi 4- aid 
brfi 4- bid 
cdi + <?irf 



and Its three-rowed principal minors vanish. 

Proof. Theorem 8, Chapter I (Section 5, page 9) enables us to write this 
determinant as the sum of 2 4 fourth order determinants whose elements are 
the product of one of the numbers a, 6, c, or d by one of the numbers a\ y 61, Ci, 
or d\. A somewhat careful inspection shows that in every one of these 2 4 
determinants at least two columns are proportional; for, after common factors 
have been removed from the elements of the columns, these columns must 
consist either of the numbers a, 6, c, d, or else of the numbers i, 61, Ci, d\. 
Consequently, the value of the given determinant is zero. And every one of 
the three-ro\yed principal minors can be written as the sum of 2 s three-rowed 
determinants, in each of which there are at least two proportional columns. 

The reader should have no difficulty in carrying out the details of this 
proof; he is urged to write down explicitly a number of the simpler deter- 
minants into which those under consideration are broken up. 



APPENDIX 299 

VI 

(Compare Section 96, page 208) 

To prove: If r 4 = 2, not all the principal two-rowed minors of the 
matrix a 4 can vanish and those which do not vanish are of one sign. 

Proof. The reader should have no difficulty in proving this statement on 
the basis of Appendix IV. 

VII 

(Compare Section 105, page 241) 

a(pi 2 p 2 2 ) a Ofi 2 rj 2 2 ) a 



To prove: The determinant 



4 p 2 2 ) fe(qi 2 4- 



= 0, 



if for a, 0, 7 there are substituted the coordinates of an arbitrary point 
on the line 



and for i, 0i, yi the coordinates of an arbitrary point on the line 



Proof. If we substitute a, (3, y for x, ?/, z in the equations of the first line, 
we obtain a pair of linear equations, which may be solved for and by 
Cramer's rule; for the coefficient determinant of these equations with respect 

to - and - is equal to p L 2 -f 7*2* =N 0. We find 
a c 



and = 2 



= (Pi 2 - P2 2 ) + 2 p lpa 4- ( Pl -f 

= [ 2 
In a similar manner we obtain from the equations of the second line: 

~ = [(<7i 2 - <?2 2 ) + 2 r M2 ] -s- fe' 4- 92 s ), 
and -^ = I 27172-^ + q^ r/ 2 2 4- ((/r -f q.>~). 

If we subtract the corresponding equations of those two sots, we can deter- 
mine a ai and 7 71. We substitute those values of a a\ and 7 71 
in the determinant D and make the obvious simplifications; thus we obtain 
the following result: 



D = 



(Pi 2 4 P2 2 ) (?i 2 4- <72 2 ) 

Pi 2 - P2 2 ?1 2 - ?2 2 (qi 2 4- 72 2 ) (Pi 2 - 7>2 2 )0 - (Pi 2 4 P2 2 ) (7l 

0i 4 2 6pip 2 (<7i 2 4 72 2 ) - 2 fytf 2 (pi 2 4 
Pi 2 4- p 2 2 7i 2 4- ?2 2 (pi 2 + p 2 2 ) (7i 2 4 ?2 2 ) (0 - 00 

P1P2 -qiq* 



300 APPENDIX 

To the third column of this determinant we add the product of the first 
column by (qf + q< 2 *)0 and the product of the second column by (pi* -f- 
then we add the second row to the first. Thus we find: 

4abc 



(Pi 2 4- 7>2 2 ) (<7l 2 4- ?2 2 ) 

Pi* </i 2 Pip2(gi 2 4- ?2 2 ) - <M2(/?i 2 4- P2 2 ) 



X 



4- /> 2 2 ) (<7i 2 4- 7 

- PiV) (/>i V - 7>*V) - /'iWfar 4- tt 2 ) 2 + <7i V(Pi 2 + P2 2 ) 2 ] 
= 0. 

VIII 

(Compare Section, 109, page 245) 

To prove: If li/?i, Ij/^ , . . . are chords of a conic section which pass 
through a fixed point P and if the products PA l PD 19 PA 2 PB 2 , . . . are 
all equal, no matter what point P is taken in the plane of the conic 
section, then this conic section is a circle. 

Proof. We take a plane Cartesian frame of reference in the plane of the 
conic section. Let the equation of the conic with respect to this reference 
frame be 

C(x, y) = anx 2 + 2 a n xy + a^y 2 -\- 2 a^x + 2 a^y H- 33 = 0, 

and let P(a, ff) be an arbitrary point of the plane. We write the equations 
of an arbitrary line through P in the parametric form as follows: 

x = a + Is, y = + ms; 

here s is the parameter which designates the length of the segment of the line 
from P to the variable point (x, /y); I = cos 0, m = sin 0, where is the in- 
clination of the line. We find then that the distances from P to the points 
A and B in which the line meets the conic arc the roots of the equation 

s*c(l, m) + s[C,(, p)l + C 2 (, ft)m] 4- C(, 0) = 0, 

where c(#, y) = a n x' 2 4- 2 a^?/ 4- a-wy 2 * and (?!, C 2 are the partial derivatives 
of C(x, y) with respect to :c and y respectively (compare Sections 76 and 80). 

Therefore PA PB = "' ! ; and we have to show that if c(Z, m) is inde- 
c(L) ni) 

pendent of I and m, then the locus of C(x, y) = must be a circle. 
For = 0, we have I = 1, m = and c(Z, m) = an; 
for = 90, we have I = 0, m = 1 and c(7, m) = a 2 2 ; 

for $ = 45, we have I = ^?, m = ^ and c(/, m) = ^ + a 12 4- ~ 

& 

Therefore, if c(l, m) is independent of the direction of the line, we must have 
an = a 2 2, and a\ 2 = 0. The equation of the conic reduces then to the form 



APPENDIX 301 

Oii( 2 -h y z ) + 2 aisx H- 2 a^y + 44 = 0. And if an = 0, the locus of this 
equation is indeed a circle. If an = 0, but a^ and a 2 a do not both vanish, the 
locus is a straight line; and if an an a^ = 0, the equation has no finite 
locus. From the point of view of Protective Geometry, the locus consists, 
in these two cases of a finite line together with a line at infinite distance, and 
of a line at infinite distance counted doubly. And these pairs are also 
recognized as circles; we shall refer to them as degenerate circles. 



INDEX 

(The numbers refer to pages.) 



Adjoint of a determinant, 28 

, minor of the, 29 
Adjoint of a vanishing determinant, 

42 

Admissible values, 56 
Algebraic complement, 17 
a.m., 36 

Anchor ring, 135 
Angle between line and plane, 79 

between two lines, 63, 64 

between two planes, 82, 198 
Angles, direction, 55 
Asymptote of a quadric surface, 160 
Asymptotes of an hyperbola, 141 
Asymptotic cone, 183, 185 

, equation of the, 184 

direction, 160 
Augmented matrix, 36 
Axes, coordinate, 49 

of ellipse, 140 

of hyperbola, 141 

-, rotation of, 118, 211, 213 
, translation of, 114, 210 
Axis, conjugate, 140, 141, 145 

of parabola, 141 

of revolution, 131 

, transverse, 140, 141, 145 
, X-, 49 
, Y-, 49 
, Z-, 49 

B 

Bundle of planes, 94 



Cartesian coordinates, 49 

, oblique, 113 
Center, improper, 176, 177 



Center of quadric surface, 176, 177, 

178, 210 

Center, proper, 176, 179, 180 
Central quadrics, 179, 180 
Circles on ellipsoid, 251 

on hyperboloid of one sheet, 252 

on hyperboloid of two sheets, 
252 

on quadric surfaces, 244, 247, 
250 

Circles, parallel, 131 
Circular cone, 133 

cylinder, 69, 133 

section, 244 

Classification of quadric surfaces, 214, 

220, 227, 229, 230 
c.m., 35 
Coefficient determinant, 36 

matrix, 35 

Cofactor of an element of a deter- 
minant, 12 

Coincident planes, 81, 102, 209 
Collinear points, 57 
Column index, 2 

of a determinant, 2 

of a matrix, 16 
Complementary minor, 17 
Cone, asymptotic, 183 

, circular, 133 

, elliptic, 145 

, imaginary, 145, 185, 220 

, quadric, 166, 175, 179, 180, 185, 
220, 261, 265 

, tangent, 167 
Confocal family of quadrics, 289, 291, 

292 

Conic, focal, 283, 285, 286, 287 
Conical surface, 136, 137, 175 
Conicoid, 159 
Conjugate axis, 140, 141, 145 



303 



304 



INDEX 



Conjugate diameters of ellipsoids, 271, 

274 

Conjugate, harmonic, 164, 165 
Conjugate hyperbolas, 141 

set of diameters, 270 

set of diametral planes, 270 
Contour lines, 138 

Contour map, 138 
Coordinate axes, 49 

parallelepiped of a point, 51, 
118 

parallelepiped of two points, 
54 

planes, 49 

systems, 49, 108 

X-, 50 
-, Y-, 50 

, Z-, 50 

Coordinates, cartesian, 49 
Coordinates, cylindrical, 111 

, elliptic space, 295 

, oblique cartesian, 113 

of a point, 50 
, origin of, 49 
, spherical, 108 

, transformation of , 110, 112, 115, 

120, 121, 123 
Cosines, direction, 55 
Courbes gauches, 157 
c.p. of one point, 51 
c.p. of two points, 54 
Cramer's rule, 37 
Curve, 68 

, equation of, 68 

, meridian, 131 

, twisted, 157 
Cylinder, elliptic, 223, 261, 263, 265 

, enveloping, 271 

, hyperbolic, 223, 261, 263 

, imaginary, 223 

, parabolic, 225, 262, 263 
Cylindrical coordinates, 111 
Cylindrical surface, equation of, 69 

, oblique, 69 

, right, 69 
Cylindrical surfaces, 69 



D 

Degenerate locus, 83 
Degree of a polynomial, 126 
Degrees of freedom, 67 
Derivative of a determinant, 31 
Descartes' rule of signs, 215 
Determinant, 1 

, adjoint of a, 28 

, adjoint of a vanishing, 42 

, coefficient, 36 

, cof actor of an element of a, 12 

, columns of a, 2 

, derivative of a, 31 

, diagonals of a, 2 

, expansion of a, 3 

, Laplace development of a, 20 

, minor of a, 17 

, minor of an element of a, 12 

, notations for a, 2 

of a matrix, 16 
, order of a, 2 

, orientation, 106, 124 

, rows of a, 2 

, symmetric, 33, 42 
Determinant, value of a, 3, 13 
Determinants, product of two, 25 

27 

Development, Laplace, 20 
Diagonals of a determinant, 2 
Diameter of a quadric surface, 268 
Diameters, conjugate, 270 
Diametral plane, 186 
Diametral planes, conjugate, 270 
Direction angles of a line, 55 

, asymptotic, 160 
Direction cosines of a line, 55, 87 
Directions, principal, 187, 189, 194, 

211, 213 
Directrix cylinder, 284, 285, 286, 287 

of a conical surface, 136 

of a cylindrical surface, 69 

of an ellipse, 140 

of an hyperbola, 141 

of a parabola, 141 

of a quadric surface, 281 



INDEX 



305 



Discriminant matrix, 175 

of a quadric surface, 166, 203, 204 
Discriminating equation, 189, 190, 

192 
Discriminating numbers, 189, 192, 

200, 213 
Distance between two points, 51, 54, 

117 

from a plane to a point, 77, 198 



Eccentricity, 140, 141 
Element of a determinant, 2 

, cof actor of an, 12 

, minor of an, 12 
Element of a matrix, 16 
Elementary transformation of a ma- 
trix, 18 
Ellipse, 140 

, axes of an, 140 

, directrices of an, 140 

, eccentricity of an, 140 

, foci of an, 140 

, vertices of an, 140 
Ellipsoid, 142, 215, 217, 251, 256, 264, 
272, 275, 286 

, imaginary, 145, 215, 217 

of revolution, 133, 215 
, semi-axes of an, 142 

Elliptic cone, 145 

Elliptic coordinates, 295 

Elliptic cylinder, 223, 261, 263, 265 

Elliptic paraboloid, 146, 218, 220, 257, 

265 

Enveloping cylinder, 271 
Equation, discriminating, 189, 190, 
192 

, linear, 71 

, locus of an, 68 

of an asymptotic cone, 184, 185 

of a plane, 71 

of a plane, intercept form of the, 
73 

of a plane, normal form of the, 
78 



Equation of a plane, three-point form 

of the, 74 

Equations, equivalent, 78, 137 
, homogeneous, 35 

of a curve, 68 

of cylindrical surfaces, 69 

of a line, 83, 84, 85, 86 

of a line, symmetric, 60 

of surfaces of revolution, 132 
, parametric, 86 

, systems of homogeneous, 38, 

39, 41 
, systems of non-homogeneous, 

36, 38, 44 

Equivalent equations, 78, 137 
Euler's theorem on homogeneous 

functions, 161 
Expansion of a determinant, 3 



F 

Factorability, 209 

Focal conic, 283, 285, 286, 287 

Focal curves, 281 

Focus of the ellipse, 140 

of the hyperbola, 141 

of the parabola, 141 

of the quadric surface, 281 
Form, binary, etc., 221 

, cubic, etc., 221 
, negative definite, 221, 229 
, positive definite, 221, 229 
, quadratic ternary, 221, 229 

Four planes, 101 

Frames of reference, 49, 108 

Freedom, degrees of, 67 

G 

General position of a set of points, 
279, 280, 281 

Generating line of a cylindrical sur- 
face, 69 

Generatrix of a cylindrical surface, 69 

Geometric characterization of quadric 
surface, 232 



306 



INDEX 



H 

Harmonic conjugates, 104, 165 
Homogeneous equation, 35, 136, 137 
Homogeneous equations, system of, 

38, 39, 41 

Homogeneous function, 209 
Homogeneous functions, Euler's the- 
orem on, 161 
Hyperbola, 141 
, asymptotes of the, 141 
, axes of the, 141 
, directrices of the, 141 
, eccentricity of the, 141 
, foci of the, 141 
, vertices of the, 141 
Hyperbolas, conjugate, 141 
Hyperbolic cylinder, 223, 261, 263 
Hyperbolic paraboloid, 147, 218, 220, 

242, 259 

Hyperboloid of one sheet, 140, 142, 
215, 217, 235, 242, 252, 256, 264, 
286 

of revolution, 133, 215 

of two sheets, 144, 215, 217, 252, 
257, 264, 287 



I 

Imaginary cone, 145, 220 

cylinder, 223 

ellipsoid, 145, 215 
Index, column, 2 

, row, 2 

Infinity, plane at, 73 
Intercept form of the equation of a 

plane, 73 

Intercepts of a plane, 73 
Interchange of numbers in a row of 

integers, 6 

Interchange of two columns (rows), 8 
Intersecting planes, 81, 208, 209 
Intersection of 'two planes, 87 

of a surface and a line, 150, 153 
Invariant relations, 197 
Invariants, 197 



Invariants of a quadric surface, 199, 

205, 255 
Inversion, 3 



Laplace development, 20 
Latitude, 108 
Left-handed system, 50 
Line and plane, 97, 99 

and quadric surface, 160 

, angle between a plane and a, 79 
, equations of a, 83, 84, 85, 86 
, intersection of a surface and a, 
150, 153 

normal to a surface, 155, 156 

of intersection of two planes, 87 

of proper centers, 179 

of vertices, 179 

on a quadric surface, 160 

, parametric equations of a, 86 
, symmetric equations of a, 60 

tangent to a surface, 154, 161 
Linear equation, 71 

families of quadrics, 277 

transformation, 126 
Lines, angle between two, 63, 64 

, perpendicular, 64 
, two, 104 
Locus, degenerate, 83 

of an equation, 68 
, symmetric, 137 

Longitude, 108 

M 

Matrix, 16 

a 3 , 177, 178 

a 3 , rank of the, 177, 178, 200, 
208, 209, 217, 220, 223, 225 

a 4 , 173, 178 

a 4 , rank of the, 173, 178, 201, 
208, 209, 217, 220, 223, 225 

, augmented, 36 

b, 178, 186, 225 
, coefficient, 35 
, column of a, 16 



INDEX 



307 



Matrix, discriminant, 175 
, elementary transformation of a, 

18* 

, elements of a, 16 
, minor of a, 17 
, notations for a, 16 

of a determinant, 16 
, rank of a, 16 

, rows of a, 16 

, singular square, 43 

, square, 10, 43 

, symmetric square, 43 
Meridian curve, 131 
Minor, algebraic complement of a, 

17 
Minor of a determinant, 17 

of a matrix, 17 

of an element of a determinant, 
12 

of the adjoint of a determinant, 
29 

, principal, 17 
Minors, complementary, 17 



N 

Nappes of a surface, 133, 136 
Non-homogeneous equations, system 

of, 36, 38, 44 
Normal form of the equation of a 

plane, 78 

Normal to a surface, 155, 156 
Notations for determinants, 2 

for matrices, 16 
Numbers, discriminating, 189, 192 



O 

Oblate spheroid, 133 

Oblique cartesian coordinates, 113 

Order of a determinant, 2 

of a surface, 154 
Orientation determinant, 106, 124 
Origin of coordinates, 49 
Orthogonal transformation, 123 



P 

Parabola, 141 
, directrix of the, 141 
, focus of the, 141 
, vertex of the, 141 
Parabolic! cylinder, 225, 262, 263 
Paraboloid, elliptic, 146, 218, 220, 

257, 265 

, hyperbolic, 147, 218, 220, 242, 
259 

of revolution, 133 
Parallel circles, 131 
Parallel planes, 81, 209 
Parallelepiped, coordinate, 51, 54, 118 

, volume of the, 118, 274 
Parametric equations of the line, 86 
Pencil of planes, 92, 96 
Perpendicular lines, 64 
Perpendicular planes, 82 
Plane and line, 97, 99 

, angle between a line arid a, 79 

at finite distance, 73 

at infinity, 73 
, diametral, 186 

, distance from a, 76, 77 
, equation of a, 71 
, intercepts of a, 73 
, normal form of the equation of 
a, 78 

of proper centers, 179 

of vertices, 179 

, polar, 162, 165, 167 
, principal, 187, 189, 194 

section of a surface, 127, 128 

tangent to a surface, 155, 161, 
162 

, three point form of the equation 

of a, 74 

Planes, angle between two, 82, 198 
, bundle of, 94 
, coincident, 81, 102, 209 
, coordinate, 49 , 
, four, 101 

, intersecting, 81, 208, 209 
, line of intersection of two, 87 



308 



INDEX 



Planes, parallel, 81, 209 

, pencil of, 92, 96 

, perpendicular, 82 

, three, 90 

, two, 82, 206 
Point, distance from a plane to a, 

77 

Polar plane, 162, 165, 167 
Pole, 162 
p-regulus, 236 

Principal diagonal of a determinant, 2 
Principal directions, 187, 189, 194, 

213 

Principal minor, 17 
Principal planes, 187, 189, 194, 211 
Prism, triangular, 96 
Product of two determinants, 25 

, columns by columns, 27 

, columns by rows, 27 

, rows by columns, 27 

, rows by rows, 27 
Projection method, 61, 76, 116, 119 
Prolate spheroid, 133 
Proper quadric cone, 179 

Q 

q-regulus, 236 

Quadratic form, ternary, 221, 229 

Quadric, central, 179 

Quadric cone, 166, 175, 179 

Quadric surface, 159 
, asymptote of a, 160 
, asymptotic direction of a, 160 
, center of a, 176, 177, 178 
, circles on a, 244, 247, 250 
, diameter of a, 268 
, directrix cylinder of a, 284, 285, 

286, 287 

, directrix of a, 281 
, discriminant of a, 166, 203, 204 
, enveloping cylinder of a, 271 
, focal conies of a, 283, 285, 286, 

287 

, focus of a, 281 
, invariants of a, 199, 205, 255 



Quadric surface, line tangent to a, 161 

, normal to a, 161 

, plane tangent to a, 161, 253 

, polar plane of a point with re- 
spect to a, 162 

, pole of a plane with respect to 
a, 162 

, singular, 166, 175 

, umbilics of a, 253 

, vertex of a, 171, 177, 178 

, a line and a, 160 
Quadric surfaces, classification of, 214, 
220, 227, 229, 230 

, confocal family of, 289, 291, 
292 

, geometric characterization of, 
232 

, linear families of, 277 

, ruled, 170 

R 

Radius vector, 108 
Rank of a matrix, 16 
Reference, frames of, 49, 108 
Reguli on the hyperbolic paraboloid, 
242 

on the hyperboloid of one sheet, 

235, 242 ' 
Regulus, 236 
Relations, invariant, 197 
Revolution, axis of, 131 

, ellipsoid of, 133, 215 

, hyperboloids of, 133, 215 

, paraboloid of, 133 

, surface of, 131, 266 
Right-handed system, 50 
Rigid transformation, 126 
Rotation of axes, 118, 201, 202, 211, 

213 

Row index, 2 
Row of a determinant, 2 
Rows of a matrix, 16 
Rule, Cramer's, 37 
Rule of signs, Descartes', 215 
Ruled quadric surfaces, 170, 235, 242 



INDEX 



309 



s 

Section, circular, 244 

Section of a surface, plane, 127, 

128 

Semi-axes of the ellipsoid, 142 
Sheets of a surface, 133, 136 
Singular quadric surface, 166, 175, 

^180, 202 

Singular square matrix, 43 
Solution, trivial, 38 
Sphere, 133, 136 
Spherical coordinates, 108 
Spheroid, oblate, 133 

, prolate, 133 
Square matrix, 16, 43 

, singular, 43 

, symmetric, 43 
Surface, 68 

, conical, 136, 175 

, cylindrical, 69 

, invariants of a, 197 

, line tangent to a, 154, 161 

, nappes of a, 133, 136 

, normal to a, 155, 156 

of order n, 154 

of revolution, 131, 266 

of revolution, equation of, 132 
, plane section of a, 127, 128 
, plane tangent to a, 155, 161, 

162 

, quadric, 159 

, shape of, 137, 138 

, sheets of a, 133, 136 
Surface and line, intersection of, 150, 

153 

Symmetric determinant, 33, 42 
Symmetric equations of a line, 60 
Symmetric locus, 137 
Symmetric square matrix, 43 
Symmetry, 137 
System, left-handed, 50 

, right-handed, 50 
Systems of coordinates, 49, 108 

of homogeneous equations, 38, 
39,41 



Systems of non-homogeneous equa- 
tions, 36, 38, 44 
of planes, 68, 70 



Tangent cone, 167 

Tangent line to a surface, 154, 161 

Tangent plane to a surface, 155, 161, 

162, 241, 253 
Taylor's theorem, 151 
Tetrahedron, 107 
Theorem on homogeneous functions, 

Euler's 161 
, Taylor's, 151 
Three-point form of the equation of a 

plane, 74 
Torus, 135 
Transformation, linear, 126 

of coordinates, 110, 112, 115, 
120, 121, 123 

of a matrix, elementary, 18 
, orthogonal, 123 

, rigid, 126 

Translation of axes, 114, 200, 201, 210 
Transverse axis, 140, 141, 145 
Triangular prism, 96 
Trihedral angle, 96 
Trivial solution, 38 
Twisted curves, 157 
Two lines, 104 

U 

Umbilics, 253, 263, 266 

on the cone, 265 

on the ellipsoid, 264 

on the elliptic cylinder, 265 

on the elliptic paraboloid, 265 

on the hyperboloid of one sheet, 
264 

on the hyperboloid of two sheets, 
264 

on the surfaces of revolution, 
266 

Units, 50, 108, 109, 110, 111, 113, 117, 
118 



310 



INDEX 



Value of a determinant, 3, 13 
Vertex of a conical surface, 136, 137 

of the ellipse, 140 

of the hyperbola, 141 

of the parabola, 141 

of a quadric surface, 171, 175, 
177, 178, 179 



A^-axis, 49 



A"-contour lines, 138 
^-coordinate, 50 



F-axis, 49 

F-contour lines, 138 
^-coordinate, 50 



#-axis, 49 

Z-eontour lines, 138 
2-coordinate, 50