You are on page 1of 455

VECTOR and TENSOR

ANALYSIS
By
LOUIS BRAND, Ch.E., E.E., Ph.D.
PROFESSOR OF MATHEMATICS
UNIVERSITY OF CINCINNATI

New York JOHN WILEY & SONS, Inc.


London CHAPMAN & HALL, Limited
VECTOR AND TENSOR ANALYSIS.
By Louis Brand. 439 pages. 5Y2 by
8%. Cloth.
VECTORIAL MECHANICS. By Louis
Brand. 544 pages. 5% by 8%.
Cloth.
Published by John Wiley & Sons, Inc.
VECTOR and TENSOR
ANALYSIS
By
LOUIS BRAND, Ch.E., E.E., Ph.D.
PROFESSOR OF MATHEMATICS
UNIVERSITY OF CINCINNATI

New York JOHN WILEY & SONS, Inc.


London CHAPMAN & HALL, Limited
COPYRIGHT, 1947
BY
Louis BRAND

All Rights Reserved


This book or any part thereof must not
be reproduced in any form without
the writ+en permission of the putlisher.

THIRD PRINTING, NOVEMBER, 1948

PRINTED IN THE UNITED STATES OF AMERICA


To My Wife
PREFACE

The vector analysis of Gibbs and Heaviside and the more


general tensor analysis of Ricci are now recognized as standard
tools in mechanics, hydrodynamics, and electrodynamics. These
disciplines have also proved their worth in pure mathematics,
especially in differential geometry. Their use not only materially
simplifies and condenses. the exposition, but also makes mathe-
matical concepts more tangible and easy to grasp. Moreover
tensor analysis provides a simple automatic method for construct-
ing invariants. Since a tensor equation has precisely the same
form in all coordinate systems, the desirability of stating physical
laws or geometrical properties in tensor form is manifest.. The
perfect adaptability of the tensor calculus to the theory of rela-
tivity was responsible for its original renown. It has since won a
firm place in mathematical physics and engineering technology
Thus the British analyst E. T. Whittaker rates the discovery of
the tensor calculus as one of the three principal mathematical
advances in the last quarter of the 19th century.
The first volume of this work not only comprises the standard
vector analysis of Gibbs, including dyadics or tensors of valence
two, but also supplies an introduction to the algebra of motors,
which is apparently destined to play an important role in mechanics
as well as in line geometry. The entire theory is illustrated by
many significant applications; and surface geometry and hydro-
dynamics * are treated at some length by vector methods in
separate chapters.
For the sake of concreteness, tensor analysis is first developed
in 3-space, then extended to space of n dimensions. As in the case
of vectors and dyadics, I have distinguished the invariant tensor
from its components. This leads to a straightforward treatment
of the affine connection and of covariant differentiation; and also
to a simple introduction of the curvature tensor. Applications of
tensor analysis to relativity, electrodynamics and rotating elec-
* For a systematic development of mechanics in vector notation see the
author's Vectorial Mechanics, John Wiley & Sons, New York, 1930.
vii
viii PREFACE
tric machines are reserved for the second volume. The present
volume concludes with a brief introduction to quaternions, the
source of vector analysis, and their use in dealing with finite
rotations.
Nearly all of the important results are formulated as theorems,
in which the essential conditions are explicitly stated. In this
connection the student should observe the distinction between
necessary and sufficient conditions. If the assumption of a certain
property P leads deductively to a condition C, the condition is
necessary. But if the assumption of the condition C leads deduc-
tively to the property P, the condition C is sufficient. Thus we-
have symbolically
P -* C (necessary), C (sufficient) -> P.
When P C, the condition C is necessary and sufficient.
The problems at the end of each chapter have been chosen not
only to develop the student's technical skill, but also to introduce
new and important applications. Some of the problems are mathe-
matical projects which the student may carry through step by step
and thus arrive at really important results.
As very full cross references are given in this book, an article
as well as a page number is given at the top of each page. Equa-
tions are numbered serially (1), (2), . . . in each article. A ref-
erence to an equation in another article is made by giving article
and number to the left and right of a point; thus (24.9) means
article 2.4, equation 9. Figures are given the number of the article
in which they appear followed by a serial letter; Fig. 6d, for ex-
ample, is the fourth figure in article 6.
Bold-face type is used in the text to denote vectors or tensors
of higher valence with their complement of base vectors. Scalar
components of vectors and tensors are printed in italic type.
The rich and diverse field amenable to vector and tensor methods
is one of the most fascinating in applied mathematics. It is hoped
that the reasoning will not only appeal to the mind but also im-
pinge on the reader's aesthetic sense. For mathematics, which
Gauss esteemed as "the queen of the sciences" is also one of the
great arts. For, in the eloquent words of Bertrand Russell:
"The true spirit of delight, the exaltation, the sense of being
more than man, which is the touchstone of the highest excellence,
is to be found in mathematics as surely as in poetry. What is
PREFACE is

best in mathematics deserves not merely to be learned as a task,


but also to be assimilated as a part of daily thought, and brought
again and again before the mind with ever-renewed encourage-
ment. Real life is, to most men, a long second-best, a perpetual
compromise between the real and the possible; but the world of
pure reason knows no compromise, no practical limitations, no
barrier to the creative activity embodying in splendid edifices the
passionate aspiration after the perfect from which all great work
springs."
The material in this book may be adapted to several short
Thus Chapters I, III, IV, V, and VI may serve as a
courses.
course in vector analysis; and Chapters I (in part), IV, V, and IX
as one in tensor analysis. I But the prime purpose of the author
was to cover the theory and simpler applications of vector and
tensor analysis in ordinary space, and to weave into this fabric
such concepts as dyadics, matrices, motors, and quaternions.
The author wishes, finally, to express his thanks to his col-
leagues, Professor J. W. Surbaugh and Mr. Louis Doty for their
help with the figures. Mr. Doty also suggested the notation used
in the problems dealing with air navigation and read the entire
page proof.
Louis BRAND
University of Cincinnati
January 15, 1947
CONTENTS
PAGE
PREFACE . . . . . . . . . . . . . . . . . . . . . . . . . . . . Vii

CHAPTER I
VECTOR ALGEBRA
ARTICLE
1. Scalars and Vectors . . . . . . . . . . . . . . . . . . . . . 1
2. Addition of Vectors . . . . . . . . . . . . . . . . . . . . . 3
3. Subtraction of Vectors . . . . . . . . . . . . . . . . . . . 5
4. Multiplication of Vectors by Numbers . . . . . . . . . . . . . 6
5. Linear Dependence . . . . . . . . . . . . . . . . . . . . . 7
6. Collinear Points . . . . . . . . . . . . . . . . . . . . . . 8
7. Coplanar Points 12
8. Linear Relations Independent of the Origin. . . . . . . . . . 18
9. Centroid . . . . . . . . . . . . . . . . . . . . . . . . . 19
10. Barycentric Coordinates . . . . . . . . . . . . . . . . . . . 23
11. Projection of a Vector . . . . . . . . . . . . . . . . . . . . 24
12. Base Vectors . . . . . . . . . . . . . . . . . . . . . . . . 24
13. Rectangular Components . . . . . . . . . . . . . . . . . . 26
14. Products of Two Vectors . . . . . . . . . . . . . . . . . . . 29
15. Scalar Product . . . . . . . . . . . . . . . . . . . . . . . 29
16. Vector Product . . . . .
. . . . . . . . . . . . . . . . . . 34
17. Vector Areas . . . . . . . . . . . . . . . . . . . . . . . . 37
18. Vector Triple Product . . . . . . . . . . . . . . . . . . . . 40
19. Scalar Triple Product . . . . . . . . . . . . . . . . . . . . 41
20. Products of Four Vectors . . . . . . . . . . . . . . . . . . 43
21. Plane Trigonometry . . . . . . . . . . . . . . . . . . . . 44
22. Spherical Trigonometry . . . . . . . . . . . . . . . . . . . 44
23. Reciprocal Bases . . . . . . . . . . . . . . . . . . . . . . 46
24. Components of a Vector . . . . . . . . . . . . . . . . . . . 48
25. Vector Equations . . . . . . . . . . . . . . . . . . . . . . 50
26. Homogeneous Coordinates . . . . . . . . . . . . . . . . . . 51
27. Line Vectors and Moments . . . . . . . . . . . . . . . . . 55
28. Summary: Vector Algebra . . . . . . . . . . . . . . . . . . 57
Problems . . . . . . .
. . . . . . . . . . . . . . . . . . 59

CHAPTER II
MOTOR ALGEBRA
29. Dual Vectors . . . . . . . . . . . . . . . . . . . . . . . . 63
30. Dual Numbers . . . . . . . . . . . . . . . . . . . . . . . 64
31 Motors . . . . . . . . . . . . . . . . . . . . . . . . . . 65
xi
xii CONTENTS
ARTICLE PAGE
32. Motor Sum . . . . . . . . . . . . . . . . . . . . . . . . 67
33. Scalar Product . . . . . . . . . . . . . . . . . . . . . . . 68
34. Motor Product . . . . . . . . . . . . . . . . . . . . . . . 70
35. Dual Triple Product . . . . . . . . . . . . . . . . . . . . . 72
36. Motor Identities . . . . . . . . . . . . . . . . . . . . . . 73
37. Reciprocal Sets of Motors . . . . . . . . . . . . . . . . . . 74
38. Statics . . . . . . . . . . . . . . . . . . . . . . . . . . 75
39. Null System . . . . . . . . . . . . . . . . . . . . . . . . 78
40. Summary: Motor Algebra . . . . . . . . . . . . . . . . . . 80
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 82

CHAPTER III
VECTOR FUNCTIONS OF ONE VARIABLE
41. Derivative of a Vector . . . . . . . . . . . . . . . . . . . . 84
42. Derivatives of Sums and Products . . . . . . . . . . . . . . 86
43. Space Curves . . . . . . . . . . . . . . . . . . . . . . . 88
44. Unit Tangent Vector . . . . . . . . . . . . . . . . . . . . 90
45. Frenet's Formulas . . . . . . . . . . . . . . . . . . . . . 92
46. Curvature and Torsion . . . . . . . . . . . . . . . . . . . 95
47. Fundamental Theorem . . . . . . . . . . . . . . . . . . . 97
48. Osculating Plane . . . . . . . . . . . . . . . . . . . . . . 98
49. Center of Curvature . . . . . . . . . . . . . . . . . . . . 99
50. Plane Curves . . . . . . . . . . . . . . . . . . . . . . . 100
51. Helices . .
. . . .
. . . . . . . . . . . . . . . . . . . . 105
52. Kinematics of a Particle . . . . . . . . . . . . . . . . . . . 108
53. Relative Velocity . . . . . . . . . . . . . . . . . . . . . 110
54. Kinematics of a Rigid Body . . . . . . . . . . . . . . . . . 114
55. Composition of Velocities . . . . . . . . . . . . . . . . . . 120
56. Rate of Change of a Vector . . . . . . . . . . . . . . . . . 121
57. Theorem of Coriolis . . . . . . . . . . . . . . . . . . . . . 123
58. Derivative of a Motor . . . . . . . . . . . . . . . . . . . . 126
59. Summary: Vector Derivatives . . . . . . . . . . . . . . . . 128
Problems . . . . . . . . . .
. . . . . . . . . . . . . . . 130

CHAPTER IV
LINEAR VECTOR FUNCTIONS
60. Vector Functions of a Vector . . . . . . . . . . . . . . . . . 135
61. Dyadics . . . . . . . . . . . . . . . . . . . . . . . . . . 136
62. Affine Point Transformation . . . . . . . . . . . . . . . . . 138
63. Complete and Singular Dyadics . . . . . . . . . . . . . . . . 139
64. Conjugate Dyadics . . . . . . . . . . . . . . . . . . . . . 141
65. Product of Dyadics . . . . . . . . . . . . . . . . . . . . . 142
66. Idemfactor and Reciprocal . . . . . . . . . . . . . . . . . . 144
67. The Dyadic 4) x v . . . . . . . . . . . . . . . . . . . . . . 146
68. First Scalar and Vector Invariant . . . . . . . . . . . . . . . 147
69. Further Invariants . . . . . . . . . . . . . . . . . . . . . 148
CONTENTS xiii
ARTICLE PAGE
70. Second and Adjoint Dyadic . . . . . . . . . . . . . . . . . 151
71. Invariant Directions . . . . . . . . . . . . . . . . . . . . 153
72. Symmetric Dyadics . . . . . . . . . . . . . . . . . . . . . 156
73. The- Hamilton-Cayley Equation . . . . . . . . . . . . . . . 160
74. Normal Form of the General Dyadic . . . . . . . . . . . . . 162
75. Rotations and Reflections . . . . . . . . . . . . . . . . . . 164
76. Basic Dyads . . . . . .
. . . . . . . . . . . . . . . . . . 166
77. Nonion Form . . . . . . . . . . . . . . . . . . . . . . . 167
78. Matric Algebra . . . . . . . . . . . . . . . . . . . . . . . 169
79. Differentiation of Dyadics . . . . . . . . . . . . . . . . . . 171
80. Triadics . . . .
. . . . . . . . . . . . . . . . . . . . . . 172
81. Summary: Dyadic Algebra . . . . . . . . . . . . . . . . . . 172
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 174

CHAPTER V
DIFFERENTIAL INVARIANTS
82. Gradient of a Scalar . . . . . . . . . . . . . . . . . . . . . 178
83. Gradient of a Vector . . . . . . . . . . . . . . . . . . . . 181
84. Divergence and Rotation . . . . . . . . . . . . . . . . . . 183
85. Differentiation Formulas . . . . . . . . . . . . . . . . . . 186
86. Gradient of a Tensor . . . . . . . . . . . . . . . . . . . . 187
87. Functional Dependence . . . . . . . . . . . . . . . . . . . 190
88. Curvilinear Coordinates . . . . . . . . . . . . . . . . . . . 191
89. Orthogonal Coordinates . . . . . . . . . . . . . . . . . . . 194
90. Total Differential . . . . . . .
. . . . . . . . . . . . . 197
91. Irrotational Vectors . . . . . . . . . . . . . . . . . . . . . 198
92. Solenoidal Vectors . . . . . . . . . . . . . . . . . . . . . 201
93. Surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . 203
94. First Fundamental Form . . . . . . . . . . . . . . . . . . 204
95. Surface Gradients . . . . . . . . . . . . . . . . . . . . . . 206
96. Surface Divergence and Rotation . . . . . . . . . . . . . . . 207
97. Spatial and Surface Invariants . . . . . . . . . . . . . . . . 209
98. Summary: Differential Invariants . . . . . . . . . . . . . . . 211
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 213

CHAPTER VI
INTEGRAL TRANSFORMATIONS
99. Green's Theorem in the Plane . . . . . . . . . . . . . . . . 216
100. Reduction of Surface to Line Integrals . . . . . . . . . . . . 218
101. Alternative Form of Transformation . . . . . . . . . . . . . 221
102. Line Integrals . . . . 222
103. Line Integrals on a Surface . . . . . . . . . . . . . . . . . 224
104. Field Lines of a Vector . . . . . . . . . . . . . . . . . . . 226
105. Pfaff's Problem . . . . . . . 230
106. Reduction of Volume to Surface Integrals . . . . . . . . . . . 233
107. Solid Angle . . . . . . . . . . . . . . . . . . . . . . . . 236
xiv CONTENTS
ARTICLE PAGE
108. Green's Identities . . . . . . . . . . . . . . . . . . . . . . 237
109. Harmonic Functions . . . . . . . . . . . . . . . . . . . . 239
110. Electric Point Charges . . . . . . . . . . . . . . . . . . . 241
111. Surface Charges . . . . . . . . . . . . . . . . . . . . . . 242
112. Doublets and Double Layers . . . . . . . . . . . . . . . . . 243
113. Space Charges . . . . . . . . . . . . . . . . . . . . . . . 245
114. Heat Conduction . . . . . 247
115. Summary: Integral Transformations . . . . . . . . . . . . . . 248
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 250

CHAPTER VII
HYDRODYNAMICS

116. Stress Dyadic . . . . . . . . . . . . . . . . . . . . . . . 253


117. Equilibrium of a Deformable Body . . . . . . . . . . . . . . 255
118. Equilibrium of a Fluid . . . . . . . . . . . . . . . . . . . . 257
119. Floating Body . . . . . . . . . . .. . . . . . . . . . . . 257
120. Equation of Continuity . . . . . . . . . . . . . . . . . . . 258
121. Eulerian Equation for a Fluid in Motion. . . . . . . . . . . . 260
122. Vorticity . . . . . . . . . . . . . . . . . . . . . . . . . 262
123. Lagrangian Equation of Motion . . . . . . . . . . . . . . . . 263
124. Flow and Circulation . . . . . . . . . . . . . . . . . . . . 266
125. Irrotational Motion . . . . . . . . . . . . . . . . . . . . . 267
126. Steady Motion . . . . . . . . . . . . . . . . . . . . . . . 268
127. Plane Motion . . . . . . . . . . . . . . . . . . . . . . . 270
128. Kutta-Joukowsky Formulas . . . . . . . . . . . . . . . . . 273
129. Summary: Hydrodynamics . . . . . . . . . . . . . . . . . . 278
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 280

CHAPTER VIII
GEOMETRY ON A SURFACE

130. Curvature of Surface Curves . . . . . . . . . . . . . . . . . 283


131. The Dyadic On . . . . . .. . . . . . . . . . . . . . . . . 285
132. Fundamental Forms . . . . . . . . . . . . . . . . . . . . 289
133. Field of Curves . . . . . . . . . . . . . . . . . . . . . . . 293
134. The Field Dyadic . . . . . . . . . . . . . . . . . . . . . . 293
135. Geodesics . . . . . . . . . .. . . . . . . . . . . . . . . 297
136. Geodesic Field . . . . . . . . . . . . . . . . . . . . . . . 299
137. Equations of Codazzi and Gauss . . . . . . . . . . . . . . . 300
138. Lines of Curvature . . . . . . . . . . . . . . . . . . . . . 302
139. Total Curvature . . . . . . . . . . . . . . . . . . . . . . 306
140. Bonnet's Integral Formula . . . . . . . . . . . . . . . . . . 308
141. Normal Systems . . . . . . . . . . . . . . . . . . . . . . 313
142. Developable Surfaces . . . . . . . . . . . . . . . . . . . . 314
143. Minimal Surfaces . . . . . . . . . . . . . . . . . . . . . . 316
144. Summary: Surface Geometry . . . . . . . . . . . . . . . . . 319
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 321
CONTENTS xv

CHAPTER IX
TENSOR ANALYSIS
ARTICLE PAGE
145. The Summation Convention . . . . . . . . . . . . . . . . . 328
146. Determinants . . . . . .. . . . . . . . . . . . . . . . . 329
147. Contragredient Transformations . . . . . . . . . . . . . . . 332
148. Covariance and Contravariance . . . . . . . . . . . . . . . . 334
149. Orthogonal Transformations . . . . . . . . . . . . . . . . . 336
150. Quadratic Forms . . . . . . . . . . . . . . . . . . . . . . 337
151. The Metric . . .
. . . . . . . . . . . . . .. . . . . . . 339
152. Relations between Reciprocal Bases . . . . . . . . . . . . . . 339
153. The Affine Group . . . . . . . . . . . . . . . . . . . . . . 340
154. Dyadics . . . . . . . . . . . . . . . . . . . . . . . . . . 342
155. Absolute Tensors . . . . . . . . . . . . . . . . . . . . . . 343
156. Relative Tensors . . . . . . . . . . . . . . . . . . . . . . 345
157. General Transformations . . . . . . . . . . . . . . . . . . 346
158. Permutation Tensor . . . . . . . . . . . . . . . . . . . . 350
159. Operations with Tensors . . . . . . . . . . . . . . . . . . . 350
160. Symmetry and Antisymmetry . . . . . . . . . . . . . . . . 352
161. Kronecker Deltas . . . . . . . . . . . . . . . . . . . . . . 353
162. Vector Algebra in Index Notation . . . . . . . . . . . . . . . 354
163. The Affine Connection . . . . . . . . . . . . . . . . . . . 356
164. Kinematics of a Particle . . . . . . . . . . . . . . . . . . . 359
165. Derivatives of e` and E . . . . . . . . . . . . . . . . . . . 360
166. Relation between Affine Connection and Metric Tensor . . . . . 361
167. Covariant Derivative . . . . . . . . . . . . . . . . . . . . 362
168. Rules of Covariant Differentiation . . . . . . . . . . . . . . . 365
169. Riemannian Geometry . . . . . . . . . . . . . . . . . . . 366
170. Dual of a Tensor . . . . . . . . . . . . . . . . . . . . . . 370
171. Divergence . . . . . . . . . . . . . . . . . . . . . . . . 371
172. Stokes Tensor . . . . . . . . . . .
. . . . . . . . . . . . 372
173. Curl . . ... . . . . . . . . . . . . . . . . . . . . . . 374
174. Relation between Divergence and Curl . . . . . . . . . . . . 376
175. Parallel Displacement . . . . . . . . . . . . . . . . . . . . 376
176. Curvature Tensor .. . . . . . . . . . . . . . . . . . . . 380
177. Identities of Ricci and Bianchi . . . . . . . . . . . . . . . . 384
178. Euclidean Geometry . . .. . . . . . . . . . . . . . . . . 385
179. Surface Geometry in Tensor Notation . . . . . . . . . . . . . 388
180. Summary: Tensor Analysis . . . . . . . .. . . . . . . . . . 392
Problems . . . . . . . . . . . .. . . . . . . . . . . . . 396

CHAPTER X
QUATERNIONS

181. Quaternion Algebra . . . . . . . . . . . . . . . . . . . . . 403


182. Conjugate and Norm . . . . . . . . . . . . . . . . . . . . 406
183. Division of Quaternions . . . . . . . . . . . . . . . . . . . 409
xvi CONTENTS
ARTICLE PAGE
184. Product of Vectors . . . . . . . . . . . . . . . . . . . . . 410
185. Roots of a Quaternion . . . . . . . . . . . . . . . . . . . . 412
186. Great Circle Arcs . . . . . . . . . . . . . . . . . . . . . . 414
187. Rotations . . . . . . . . . . . . . . . . . . . . . . . . . 417
188. Plane Vector Analysis . . . . . . . . . . . . . . . . . . . . 421
189. Summary: Quaternion Algebra . . . . . . . . . . . . . . . . 426
Problems . . . . . . . . . . . . . . . . . . . . . . . . . 427,
...
INDEX . . . . . . . . . . . . . . . . . . . . . . . . . . . 431
CHAPTER I

VECTOR ALGEBRA

1. Scalars and Vectors. There are certain physical quantities,


such as length, time, mass, temperature, electric charge, that may
be specified by a single real number. A mass, for example, may
be specified by the positive number equal to the ratio of the given
mass to the unit mass. Similarly an electric charge may be speci-
fied by a number, positive or negative, according as the charge is
"positive" or "negative." Quantities of this sort are called scalar
quantities; and the numbers that represent them are often called
scalars.
On the other hand, some physical quantities require a direction
as well as magnitude for their specification. Thus a rectilinear
displacement can only be completely specified by its length and
direction. A displacement may be represented graphically by a
segment of a straight line having a definite length and direction.
We shall call such a directed segment a vector. Any physical
quantity that involves both magnitude and direction, so that it
may be represented by a line segment of definite length and direc-
tion, and which moreover conforms to the parallelogram law of
addition (§ 2), is called a vector quantity. Velocity, acceleration,
force, electric and magnetic field intensities are examples of vector
quantities. It is customary, however, in applied mathematics, to
speak of vector quantities as vectors.
DEFINITIONS: A vector is a segment of a straight line regarded as
having a definite length and direction. Thus we may represent a
vector by an arrow. The vector directed from the point A to the
point B is denoted by the symbol AB. With this notation AB
--
and BA denote different vectors; they have the same length but
opposite directions.
Besides the proper vectors just defined, we extend the term vector
to include the zero vector, an "arrow" of length zero but devoid
of direction.
1
2 VECTOR ALGEBRA §1

If the initial point of a vector may be chosen at pleasure, the


vector is said to be free. If, however, its initial point is restricted
to a certain set of points, the vector is said to be localized in this
set. When the set consists of a single point (initial point fixed),
the vector is said to be a bound vector. If the vector is restricted
to the line of which it forms a part, it is called a line vector. For
example, the forces acting upon rigid bodies must be regarded as
line vectors; they may only be shifted along their line of action.
Two vectors are said to be equal when they have the same length
and direction.
Vectors are said to be collinear when they are parallel to the same
line. In this sense two parallel vectors are collinear.
Vectors are said to be coplanar when they are parallel to the same
plane. In this sense any two vectors are coplanar.
A unit vector is a vector of unit length.
In addition to the foregoing notation, in which we denote a
vector by giving its end points, we also shall employ single letters
in heavy (bold-face) type to denote vectors. Thus, in Fig. 2b the
vectors forming the opposite sides of the parallelogram are equal
(they have the same length and direction) and may therefore be
represented by the same symbol,

AB=DC=u, AD=BC=v.
A vector symbol between vertical bars, as I AB I or I u 1, de-
notes the length of the vector. We shall also, on occasion, denote
the lengths of vectors, u, v, F by the corresponding letters u, v, F
in italic type. Bars about real numbers denote their positive mag-
nitudes: thus I -3 1 = 3.
The preceding definition of a vector is adequate for the elemen-
tary applications to Euclidean space of three dimensions. For
purposes of generalization, however, it is far better to define a
vector as a new type of number-a hypernumber-which is given
by a set of real numbers written in a definite order. Thus, in our
ordinary space it will be -seen that, when a suitable system of
reference has been adopted, a vector can be represented by a set
of three real numbers, [a, b, c], called the components of the vector.
With this definition the zero vector is denoted by [0, 0, 0]. In
order to complete this definition of a vector, a rule must be given
to enable us to compute the components when the system of refer-
ence is changed.
§2 ADDITION OF VECTORS 3

2. Addition of Vectors. To obtain a rule for adding vectors, let


us regard them, for the moment, as representing rectilinear dis-
placements in space. If a particle is given two rectilinear displace-
ments, one from A to B, and a second from B to C, the result is
the same as if the particle were given a-single displacement from A
to C. This equivalence may be represented by the notation,

(1) AB + BC = AC.
We shall regard this equation as the definition of vector addition.
The sum of two vectors, u, v, therefore is defined by the following
triangle construction (Fig. 2a) :
Draw v from the end of u; then the vector directed from the begin-
ning of u to the end of v is the sum of u and v and is written u + v.
C

FIG. 2a Fia. 2b

Since any side of a triangle is less than the sum of the other
two sides,
lu+vl _ lul +lvl,
the equal sign holding only when u and v have the same direction.
Vector addition obeys both the commutative and associative laws:
(2) u+v = v+u,
(3) (u + v) + w = u + (v + w).
In the parallelogram formed with u and v as sides (Fig. 2b),

u+v=AB+BC=AC, v+u=AD+DC=AC.
This proves (2). In view of this construction, the rule for vector
addition is called the parallelogram law.
To find u + v when u and v are line vectors whose lines intersect
at A, shift the vectors along their lines so that both issue from A
(Fig. 2b) and complete the parallelogram ABCD; then
---3 --3 ---9
u+v=AB+AD= AC.
4 VECTOR ALGEBRA §2

Since the diagonal of the parallelogram on u, v gives the line of


action of u + v, the term "parallelogram law" is especially appro-
priate for the addition of intersecting line vectors. We shall speak
of the addition of line vectors as statical addition.
The associative law (3) is evident from Fig. 2c:
(u + v) + w = (AB -i- BC) -f- CD = AC -i- CD = AD,
u + (v -F w) = AB -f- (BC + CD) = AB + BD = AD.
Since the grouping of the vectors is immaterial, the preceding swa
is simply written u + v + w.

B
Fia. 2c FIG. 2d

From the commutative and associative laws we may deduce the


following general result: The sum of any number of vectors is inde-
pendent of the order in which they are added, and of their grouping
to form partial sums.
To construct the sum of any number of vectors, form a broken
line whose segments, in length and direction, are these vectors
taken in any order whatever; then the vector directed from the
beginning to the end of the broken line will be the required sum.
The figure formed by the vectors and their sum is called a vector
polygon. If A, B, C, , G, H are the successive vertices of a
vector polygon (Fig. 2d), then
(4)
When the vectors to be added are all parallel, the vector "polygon"
becomes a portion of a straight line described twice.
If, in the construction of a vector sum, the end point of the last
vector coincides with the origin of the first, we say that the sum
of the vectors is zero. Thus, if in (4) the point H coincides with A,
we write
-3 -4 -- 3
AB+BC++GA=O.
(5)
§3 SUBTRACTION OF VECTORS 5

This equation may be regarded as a special case of (4) if we agree


--->
that AA = 0.
The zero vector AA (or BB, etc.) is not a vector in the proper
sense since it has no definite direction; it is an extension of our
original vector concept. From
--- f
AB+BB=AB,
----* --3 -* -3 -3
AA+AB=AB
we have, on writing AB = u,
u+0=u, 0+u=u.
We shall refer to vectors which are not zero as proper vectors.
3. Subtraction of Vectors. The sum of two vectors is zero when,
and only when, they have the same length and opposite directions:
AB + BA = 0. If AB = u, it is natural to write BA = -u in
order that the characteristic equation for negatives,
(1) u + (-u) = 0,
will hold for vectors as well as for numbers. Hence by definition:
The negative of a vector is a vector of the same length but opposite
direction.
Note also that - (-u) = u.
The difference u - v of two vectors is defined by the equation,
(2) (u - v) + v = U.
Adding -v to both sides of (2), we have
(3) u-v=u+(-v);
that is, subtracting a vector is the same as adding its negative. The
construction of u - v is shown in Fig. 3a.
A

0
Fia. 3a Fia. 3b
If 0 is chosen as a point of reference, any point P in space may
be located by giving its position vector OP. Any vector AB may
6 VECTOR ALGEBRA §4
be expressed in terms of the position vectors of its end points
(Fig. 3b),
AB = AO + OB = OB + (-OA),
and, from (3),
(4)
-4 -- -4
AB = OB - OA.
4. Multiplication of Vectors by Numbers. The vector u + u is
naturally denoted by 2u; similarly, we write -u + (-u) = -2u.
Thus, both 2u and -2u denote vectors twice as long as u; the
former has the same direction as u, the latter the opposite direc-
tion. This definition is generalized as follows:
The product au or ua of a vector u and a real number a is defined
as a vector a times as long as u, and having the same direction as u,
or the opposite, according as a is positive or negative. If a = 0,
au=0.
In accordance with this definition,
a(-u) = (-a)u = -au, (-a)(-u) = au.
These relations have the same form as the rules for multiplying
numbers. Moreover, the multiplication of a vector by numbers is
commutative (by definition), associative, and distributive:
(1) au = ua,
(2) (a/3)u = a(/3u),
(3) (a + a)u = au + (3u.
The product of the sum of two vectors by a given number is also
distributive:
(4) a(u + v) = au + av.
The proof of (4) follows immediately from the theorem that the
corresponding sides of similar triangles are proportional. Figure 4
applies to the case when a > 0.
The quotient u/a of a vector by a number
u+V av a (not zero) is defined as the product of u
v
U au by 1/a.
Fia. 4 The developments thus far show that:
As far as addition, subtraction, and multi-
plication by numbers are concerned, vectors may be treated formally
in accordance with the rules of ordinary algebra.
§5 LINEAR, DEPENDENCE 7

5. Linear Dependence. The n vectors u1, u2i , un are said


to be linearly dependent if there exist n real numbers X1, X2, ,
An, not all zero, such that
(1) X1U1 + X2U2 + ... + XnUn = 0.
If the vectors are not linearly dependent, they are said to be
linearly independent. Consequently, if a relation (1) exists be-
tween n linearly independent vectors, all the constants must be
zero.
If m vectors ul, U2, , um are linearly dependent, any greater
number n of vectors including these are also linearly dependent. For
if u1i u2, , um satisfy
X1U1 + A2U2 + ... + XrUm = 0,
we can give ?1, X2, , Xm the preceding values (at least one of
these is not zero) and take X,n+1 = X m+2 = _ X n = 0. Then
(1) is satisfied, and the n vectors ui are linearly dependent.
If Xu = 0 and X 0, u = 0; hence one vector is linearly depend-
ent only when it is the zero vector. Hence the vectors of any set
that includes the zero vector are linearly dependent. Conse-
quently, we need only consider sets of proper vectors in the theo-
rems following.
If X1u1 + X2u2 = 0 and X1 0 0, we can write ul = au2 i hence
ul and u2 are collinear. Conversely, if ul and u2 are collinear,
U1 = au2 (a 0 0). Therefore:
A necessary and sufficient condition that two proper vectors be lin-
early dependent is that they be collinear.
If X1u1 + A2u2 + X3u3 = 0 and X1 0, we can write ul =
aU2 + ,13u3; the parallelogram construction (Fig. 5a) now shows
that ul is parallel to the plane of u2 and u3. Conversely, if
U1, U2, u3 are coplanar, they are linearly dependent. For (a) if
two of the vectors are collinear, they are linearly dependent, and
the same is true of all three; and (b) if no two vectors are col-
linear, we can construct a parallelogram on ul as diagonal whose
sides are parallel to u2 and u3 (Fig. 5a), so that
ul =AC=AB+BC=au2+$u3.
Therefore:
A necessary and sufficient condition that three proper vectors be
linearly dependent is that they be coplanar.
8 VECTOR ALGEBRA §6

In space of three dimensions, four vectors ul, u2i u3, u4 are al-
ways linearly dependent. For (a) if three of the vectors are co-
planar, they are linearly dependent, and the same is true of all
four; and (b) if no three vectors are coplanar", we can construct a

FIG. 5a FIG. 5b

parallelepiped on ut as a diagonal whose edges are parallel to


U2i n3, U4 (Fig. 5b), so that
-3 ---3 --a - 3
ut = AD = AB + BC -}- CD = au2 + 9u3 -I- yu4.
Therefore: Any four vectors are linearly dependent.
6. Collinear Points. If A, B, P are points of a straight line,
P is said to divide the segment AB in the ratio X when
--> -3
(1) AP=APB.
As P passes from A to B (Fig. 6a), X increases through all positive
values from 0 to infinity. If P describes the line to the left of A,
X varies from 0 to -1; and, as P de-
scribes the line to the right of B, X
varies from -oo to -1. Thus X = 0,
X = =L00 , A = -1 correspond, respec-
tively, to the points A, B, and the in-
finitely distant "point" of the line. The
0 ratio A is positive or negative, according
FIG. 6a as P lies within or without the segment
AB.
To find the position vector of P, relative to an origin 0, write
(1) in the form,
OP - OA = X(OB - OP).
Then

(2)
--
OP =
OA + A OB
1 + A
§6 COLLINEAR POINTS 9

or, if we write X = (3/a,

(3)
-OPaOA
=
+aOB
a+0
-
In particular, if X = 1, P is the mid-point of AB.
In the following we shall denote the position vectors of the
points A, B, C, , P by a, b, c, , p. Thus if C divides AB
in the ratio /3/a,
c=ash-(3b
(4)
a+0
Thus the mid-point of AB has the position vector (a + b).
2
When the points C, D divide a segment AB internally and ex-
ternally in the same numerical ratios ±X, we have
a+Ab a - Xb
c =1+x, d=
1-x
If we solve these equations for a and b, we find that the points
A, B also divide the segment CD in the same numerical ratios
f(1 - X)/(l + X). Pairs of points A, B and C, D having this
property are said to be harmonic conjugates; either pair is the
harmonic conjugate of the other.
A useful test for collinearity is given by the following:
THEOREM. Three distinct points A, B, C lie on a straight line
when, and only when, there exist three numbers a, /3, y, different from
zero, such that
(5) aa+/3b-+-yc=0, a+$+y=0.
Proof. If A, B, C are collinear, C divides AB in some ratio /3/a;
hence on putting y = - (a + ,B) in (4) we obtain (5). Conversely,
from (5) we can deduce (4) since a + -y 0; hence C lies
on the line AB.
From (5) we conclude that C, A, B divide AB, BC, CA, re-
spectively, in the ratios S/a, y/l3, a/y whose product is 1.
If an equation of the form (5) subsists between three distinct non-
collinear points, we must conclude that a = (3 = y = 0. For at
least one coefficient y = 0; and from
as+3b=0, a+/3=0,
we have a = b (A coincides with B) unless a = 0 = 0.
10 VECTOR ALGEBRA §6
Another criterion for collinearity may be based on the statical
addition of line vectors (§ 2).
THEOREM. The points A, B, C are collinear when the line vectors
AB, BC, CA are statically equal to zero:

(6) AB +BC +CA =O.


Proof. If we use = to denote statical equivalence, AB + BC
BD, a vector through B; and, since BD + CA = 0, B, C, A
are collinear.
Example 1. In the parallelogram ABCD, E and F are the middle points
of the sides AB, BC. Show that the lines DE, DF divide the diagonal AC
into thirds and that AC cuts off a third of each line (Fig. 6b).

E " B
Fia. 6b Fia. 6c

The hypotheses of our problem are expressed by the equations:


d - a=c - b, 2e=a+b, 2f =b+c.
Let DE cut AC at X. To find x, eliminate b from the first and second equa-
tions. Thus
d + 2e 2a + c
d-a+2e=a+c and 3
=
3
=a;
for the first member represents a point on DE, the second member a point
on AC, and, since the points are the same, the point is at the intersection X
of these lines. Comparison with (4) now shows that X divides DE in the
ratio 2/1, AC in the ratio 1/2.
Let DF cut AC at Y. To find y, eliminate b from the first and third equa-
tions. Thus
d-a+2f=2c and d+2f-a+2c_
3 3

Hence Y divides DF in the ratio 2/1, AC in the ratio 2/1.


Example 2. In a plane quadrilateral ABCD, the diagonals AC, BD inter-
sect at P, the sides AB CD intersect at Q (Fig. 6c). If P divides AC and
BD in the ratios 3/2 and 1/2, respectively, in what ratios does Q divide the
segments A B, CD?
§6 COLLINEAR POINTS 11

By hypothesis
2a + 3c 2b + d
p 5 3
hence

- 5d = lOb - 6a _
6a + 9c = 10b + 5d and
4

for the first fraction represents a point


on CD, the second a point on AB,
and both points are the same, that is,
the point Q. Therefore Q divides CD
in the ratio -5/9, AB in the ratio
-10/6.
Example 3. Prove that the mid-
points of the diagonals of a complete
quadrilateral are collinear.
In the complete quadrilateral
ABCDLM (Fig. 6d) let P, Q, R be the
mid-points of the diagonals AC, BD,
LM. The sum of the line vectors,
--, ---> --, --- --i --> --->
AB + AD = 2 AQ, CB + CD = 2 CQ, QA + QC = 2 QP;
hence we have the statical equivalence,

AB +AD +CB +CD =4PQ,


for the quadrilateral ABCD with diagonals AC, BD. Similarly, for the ad-
rilateral BLDM with diagonals BD, LM,
--> --9 - 4 --+
BL + BM + DL + D31 = 4QR;
and, for the quadrilateral LA.11C with diagonals LM, AC,
--4 --4 ----+ --+
LA + LC + MA + MC = 4 RP.
-,
On adding these three equations, we find that the entire left member is stati-
cally equal to zero; for
--+ --- --> ---*
=0,
--- -9
DM+MA = 0,
-+ -4 -+
CB +BM +MC-0, CD +DL +LC -O,
are statical equations, since the vectors in each are collinear. Hence

PQ +QR+RP =0,
and I', Q, R are collinear.
12 VECTOR ALGEBRA §7

7. Coplanar Points. THEOREM. If no three of the points A, B,


C, D are collinear, they will lie in a plane when, and only when,
there exist four numbers a, /3, y, S, different from zero, such that
(1) as+$b+yc+Sd = 0, a+0+y+S = 0.
Proof. If A, B, C, D are coplanar, either AB is parallel to CD,
or AB cuts CD in a point P (not A, B, C, or D). In the respective
cases, we have
a+Ab c+A'd
b - a=K(d - c);
1+A = 1+A, -P,
where A, A' are neither 0 nor -1. In both cases, a, b, c, d are con-
nected by a linear relation of the form (1). Conversely, let us
assume that equations (1) hold good. If a + /3 = 0 (and hence
y + S = 0), we have
a(a - b) + y(c - d) = 0
and the lines AB and CD are parallel. If a + /3 5,16 0 (and hence
y+6 0),
as+/3b yc+Sd
(2)
a+/i y+6 -P
where P is a point common to the lines AB and CD. In both
cases, A, B, C, D are coplanar.
Note that (2) states that the point P in which AB, CD inter-
sect divides AB and CD in the ratios #/a, 6/y. Similarly, if
a+y; 0,
as + yc /3b + 6d
(3)
a -f-.y #+ 6 - qi
thus Q, the point in which AC and BD intersect, divides AC and
BD in the ratios y/a, S//i.
What conclusion can be drawn if a + 3 F6 0?
If an equation of the form (1) subsists between four distinct non-
coplanar points, we must conclude that a = a = y = S = 0. For
at least one coefficient 6 = 0; then, from
as+/3b+yc = 0, a+0+y = 0,
and the fact that A, B, C are not collinear, we deduce (§ 6) that
a=0=y=0.
§7 COPLANAR POINTS 13

Example 1. The Trapezoid. If ABCD is a trapezoid with AB parallel to


DC (Fig. 7a), then
AB=XDC or b-a=A(c-d).
Hence we have
b+Xd=a+Xc or b-Xc=a-Xd.
These equations may be written
b+ad_a+Xc_-P b - Ac_a - Ad
+a 1+a -q;
for the former expressions represent the point P where the diagonals BD, AC
meet, and the latter the point Q where the sides BC, AD meet. Evidently
P divides both BD and AC in the same ratio A; and Q divides both BC and
AD in the same ratio -A. In what ratio does the line PQ divide AB?
In particular, if A = 1, the trapezoid becomes a parallelogram. The diag-
onals then bisect each other at P, while Q recedes to infinity.
Q

Fia. 7a

Example 2. Theorem of Menelaus. If a line s cuts the sides BC, CA, AB of


the triangle ABC in the points P, Q, R, respectively, the product of the ratios in
which P, Q, R divide these sides equals - 1. Conversely, if P, Q, R divide the
sides of the triangle in ratios whose product is - 1, the points are collinear.
Proof. (Fig. 7b.) We lose no generality if we assume that P, Q divide
BC, CA in the ratios -y//3, -a/y:
(i) (/3-y)P=$b-yc,
(ii) (y-a)q=yc-aa.
In order to locate R, which lies on the lines PQ, AB, we seek a linear relation
between p, q, a, b. Add (i) and (ii) to eliminate c and divide by S - a; then
(R - y)P + (y - a)q Sb - as

Thus R divides AB in the ratio -S/a. The product of the division ratios
-y//3, -a/y, -/3/a is -1.
14 VECTOR ALGEBRA §7

Conversely, let us assume that P, Q, R divide BC, CA, AB in the ratios


-y/S, -a/y, -a/a whose product is -1. Then we have equations (i), (ii)
and also
(iii) (a - fl)r = as - 13b.
From these we deduce the linear relation,
(iv) (0-y)p+(y-a)q+(a-$)r=0,
in which the sum of the coefficients is zero. The points P, Q, R are therefore
collinear. *
Note. From (i), (ii),
(iii) it is easily proved that the three pairs of lines
BQ, CR; CR, AP; AP, BQ meet in the points A', B', C' given by
(-a+0+y)a' _ -aa+ftb+yc,
(a-S+y)b' =aa-$b+yc,
(a+9-y)c'=as+9b-yc.
If we add 2aa, 2gb, 2yc, respectively, to these equations, we find that the
point S given by
s =
as +Ab+yc
(v)
a+#+y
is common to the lines AA', BB', CC'. Thus to every line s given by (iv)
we have a corresponding point S given by (v)
C -the pole of s relative to the triangle ABC.
Example 3. Theorem of Ceva. If S is a
point in the plane of the triangle ABC, and the
lines SA, SB, SC cut the sides opposite in the
points A', B', C', then the product of the ratios
in which A', B', C' divide the sides BC, CA,
AB equals 1. Conversely, if A', B', C' divide
the sides BC, CA, AB in ratios whose product
C, B is 1 , the lines AA', BB', CC meet in a p oint .
Proof. (Fig. 7c.) Since A, B, C, S are
Fte.7c coplanar,
aa+pb+yc+as =0, a+0+y+S =0.
Hence
_ Sb + ye as + SS
(i)
a 0+y a+s
yc as _ Rb + as
(ii) b' =
y+a +b '
c'
=as+Sb -yc+Ss.
a+R y+S
*This conclusion is obvious; for the line PQ must meet AB in the point R
for which the product of the division ratios is -1.
§7 COPLANAR POINTS 15

These equations state that A', B', C' divide BC, CA, AB in the ratios -y/fl,
a/y, fl/a, whose product is 1. Incidentally, A', B', C' divide SA, SB, SC in
the ratios a/a, fl/S, y/3 whose sum is -1.
Conversely, let us assume that A', B', C' divide BC, CA, AB in the ratios
y/$, a/-j, fl/a whose product is 1. Then we have equations (i), (ii), (iii).
From these we find that the vectors as + (0 + y)a', fib + (y + a)b', yc +
(a + fl)c' are all equal to as + lb + yc; the point,
s
as+pb+yc
(iv)
a + fl + y
is therefore common to the lines AA', BY, CC'.
Note. From (i), (ii), (iii) it is easily proved that the three pairs of lines
BC, B'C'; CA, C'A'; AB, A'B' meet in the points P, Q, R given by
(Q-y)p=Rb-yc,
(v) (y - a)q = yc - aa,
(a-fl)r=as-/3b.
From these equations we deduce the linear relation,
(vi) (fi-y)p+(y-a)q+(a-a)r=0,
in which the sum of the coefficients is zero. The points P, Q, R therefore lie
on a line s. Thus to every point S given by (iv) we have a corresponding
line s whose points P, Q, R are given by (v)-the polar of S relative to the
triangle ABC.
Example 4. Let ABC and A'B'C' be two triangles, in the same or different
planes, so that the vertices A, B, C correspond to A', B', C', and the sides
AB, BC, CA correspond to A'B', B'C', C'A'. We then have (Fig. 7d)
P

FIG. 7d

DESARGUES' THEOREM. If the lines joining the corresponding vertices of two


triangles are concurrent, the three pairs of corresponding sides intersect in col-
linear points, and conversely.
Let the lines AA', BY, CC' intersect at S; then
as + a'a' = fib + fl'b' = yc + y'c' = s,
a+a =t3+fl' ='Y+y'=1.
16 VECTOR ALGEBRA §7

From these equations we find in the usual manner the points P Q, R in which
BC, B'C'; CA, C'A'; AB, A'B' intersect:

(i) Jr =
fib
- yc =
fl_y '6',
fl'b'
- y'c/
yC - as y'C' - a'a'
(ii) q = y_a y,_«,
1

(iii)
_«a-3b «a'-fl'b'
at - of
Hence
(iv) (l; - 'y)p + (-y - a)q + (a - 6)r = 0,
(v) 'y')p + ('y' - a')q + (a' - 6')r = 0;
either equation shows that P, Q, R are collinear.
To prove the converse, we may start with the expressions (i), (ii), (iii) for
p, q, r. These ensure that P, Q, R are collinear; but, in order that (iv) and
(v) determine the same division ratios for P, Q, R, we must have
,-.y' 'y'-a'
= h,
y y - a a -
or

(vii) a'-ha =fl'-hfi=y'-hy=k,


h and k representing the values of the equal members of (vi) and (vii). In
the usual way we now find from (i), (ii), (iii) that
aa'-haa=6'b'-hfib=y'c'-hyc=ks,
where S is a point common to AA', BB', CC'.

FIG. 7e

Example 5. The Complete Quadrangle. A complete quadrangle consists of


four coplanar points, its vertices, no three collinear, and the six lines, its sides,
which join them. The three pairs of sides which do not meet at a vertex
37 COPLANAR POINTS 17

arc said to be opposite; and the three points in which they meet are called
rliagon<zl points. In the complete quadrangle ABCD (Fig. 7e) the pairs of
opposite sides (BC, AD), (CA, BD), (AB, CD), meet at the diagonal points
L, :11, N, respectively.
The properties of this configuration of points and lines must all be con-
sequences of the fact that A, B, C, D are coplanar points, that is,

(i) CA +(3b+yc+Sd =0, a+$+y+a =0.


Since no three vertices are collinear, none of the scalars a, ,, y, S are zero.
From equations (i) we locate at once the diagonal points at the intersections
of the opposite sides:
Go +yc-aa+ad
a+S
0+y
m
_yc+aa /3b+5d
y+a 13+S
,

_aa+13b yc+Sd
a+13 'Y+s
We here assume that no two of the scalars a, fl, y, S have a zero sum; the
diagonal points L, M, N are then all in the finite plane. But if for example,
a + 0 = y + S = 0, AB and CD are parallel, and N is the point at infinity
in their common direction.
To find the points X1, X2 where LM cuts AB and CD, we seek linear rela-
tions connecting 1, m, a, b and 1, m, c, d, respectively.
Thus, from (ii) and (iii),
('Y+a)m- (S +y)1 as -/3b
a-/3 a-0 '
(0+y)1-(S+S)m yc - Sd
(vi) X2
y-S y-S
Equations (iv) and (v) show that N and X1 divide AB in the ratios '6/a
and -/3/a; hence N, X1 are harmonic conjugates of A, B.
Equations (iv) and (vi) show that N and X2 divide CD in the ratios S/y
and -S/y; hence N, X2 are harmonic conjugates of C, D.
Equation (v) shows that X1 divides LM in the ratio -(y + a)/(3 + y);
and (vi) shows that X2 divides LM in the ratio -(S + 5)/(3 + y) _
(y + a)/(3 + y); hence X1, X2 are harmonic conjugates of L, M.
We may now state the following harmonic properties of the complete quad-
rangle in the
THEOREM. Two vertices of a complete quadrangle are separated harmonically
by the diagonal point on their side and by a point on the line joining the other
two diagonal points.
Two diagonal points are separated harmonically by points on the sides passing
through the third diagonal point.
18 VECTOR ALGEBRA §8
From (ii), (iii), (iv) we next find the points X1, Y1, Z1 in which AB, BC, CA
are cut by LM, MN, NL, respectively; thus
as - $b 3b - yc yc - as
$1 = y1 = Z1 =
a-/3 I
0-y ,
y-a
Since

(a-/3)xl+(f -y)yl+(y-a)Z1 =0, a-8+0-y+y-a =0,


the points Xl, Y1, Zl are collinear. This is also a consequence of Desargues'
Theorem applied to the triangles ABC, LMN, which are in perspective
from D.
Each vertex of the quadrangle is the center of perspective of the triangles
formed by the other three vertices and by the three diagonal points. Thus,
with careful regard to exact correspondence, the triangle LMN is in perspec-

tive with DCB, CDA, BAD, ABC from A, B, C, D, respectively. Hence the
coliesponding sides of these triangles intersect in four lines a, b, c, d (only d,
the line X1Y1Z1, is shown in Fig. 7e). These lines are the polars of A, B, C, D
with respect to the triangle LMN (ex. 3).
Corresponding to the complete quadrangle given by four points A, B, C, D,
we now have a complete quadrilateral given by four lines a, b, c, d. In these
configurations the roles of points and lines are interchanged (Fig. 7f):
The quadrangle has four vertices A, B, C, D and six sides consisting of three
opposite pairs (BC, AD), (CA, BD), (AB, CD) which meet in the three diag-
onal points L, M, N.
The quadrilateral has four sides a, b, c, d and six vertices consisting of three
opposite pairs (bc, ad), (ca, bd), (ab, cd) which lie on the three diagonal lines
1, m, n.
The quadrangle and quadrilateral have the same diagonal triangle LMN or
lmn; the sides 1, m, n are opposite the vertices L, M, N.

8. Linear Relations Independent of the Origin. We have seen


that the position vectors of collinear or coplanar points satisfy a
§9 CENTROID 19

linear equation in which the sum of the scalar coefficients is zero.


The significance of such relations is given by the
THEOREM. A linear relation of the form,

(1) XiP1 + X2P2 + ... + XnPn = 0,


connecting the position vectors of the points P1, P2, , Pn will be
independent of the position of the origin 0 when, and only when, the
sum of the scalar coefficients is zero:

(2) X1 + X2 -{- ... + an = 0.

Proof. Change from 0 to a new origin 0'. Writing OPi = pi,


O'Pi = pi, 00' = d, we have pi = d + pi; hence (1) becomes
(X1 + t2 +. +Xn)d+X1p1 +X2P2 XnPn' = 0.
This equation will have the same form as (1) when and only when
(2) is satisfied.
For two, three, and four points relations independent of the
origin have a simple geometric meaning; namely, the points are
coincident, collinear, or coplanar, respectively. The question now
arises: What geometric property relates five or more points whose
position vectors satisfy a linear relation independent of the origin?
9. Centroid. We shall encounter problems in which each point
of a given set is associated with a certain number. The points,
for example, may represent particles of matter and the numbers,
their masses or electric charges. In the latter case, the numbers
may be positive or negative. We shall now define a point P*
called the centroid of a set of n points P1, P2, , Pn associated

with the numbers ml, m2i , mn, respectively. Denote any one
of these "weighted" points by the symbol m.iPi; then, if the sum
of the numbers mi is not zero, the centroid of the entire set is defined
as the point for which the sum of all the vectors m1P*Pi is zero. The
defining equation for the centroid is thus

(1) 2;miP*Pi = 0, provided Mmi Fl- 0.


As we wish P* to be uniquely defined, the case Emi = 0 must
be excluded; for then (1) is a relation independent of the position
of P* (j 8).
20 VECTOR ALGEBRA i9
When 2 mi 0 0 there is always a unique point P* which satis-
fies (1). For if we choose an origin 0 at pleasure, (1) can be
written
---4 -4
F.mi(OPi - OP*) = 0,
or

(2) (2mi)OP* = 1m1OPi.


This relation, independent of the origin, fixes the position of P*
relative to 0. The point P* thus determined is the only point
that satisfies (1) ; for if Q is a second point for which

EmiQPi = 0,
we have, on subtraction from (1),
2;mi(P*Pi
- QPi) = Timi(P*Pi + PQ) = (2;mi)P*Q = 0;
hence P*Q = 0, and Q coincides with P*.
If the position vectors of P*, Pi are written p* and pi, (2) be-
comes
(3) (2;mi)P* = Emipi.
The centroid of the points miP1 is not altered when the numbers
mi are replaced by any set of numbers cmi proportional to them;
for, in (2), the constant c may be canceled from numerator and
denominator. In particular, if the numbers mi are all equal, we
may replace them all by unity; the centroid of the n points then
is called their mean center and is given by the equation:

(4)
-3 1 --
OP* _ - ZOP1, or p* _ - Zpi.
1

n n
In finding the centroid P* of any set of weighted points miP1,
we may replace any subset of points for which the sum of the
weights is a number m' 0 0 by their centroid P with the weight
m'. For, if V and 1", respectively, denote summations extended
over the points of the subset and over all the remaining points,
we may write (3) in the form,
(2;'mi + Z"mi)P* = 7,'mipi + E"miPi,
or
(m' + I"mi)P* = m'p' + E"mipi.
§9 CENTROID 21

This equation shows that P* is also the centroid of the point m'P'
and the points miPi not included in the subset.
Finally, let us consider the nature of a set of n points miPi
(?Pi F4- 0) for which Ximi = 0.If we attempt to find a point P*
which satisfies
(5) ZmiP*Pi = 0 when Xmi = 0,
we find that there are two possibilities.
(a) For any arbitrary choice of the points miPi, subject only to
the condition Emi = 0, there will in general be no point P* that
will satisfy (5). Thus in the cases n = 2, 3, 4, equations (5) imply,
respectively, that P1 and P2 coincide; P1, P2, P3 are collinear;
P1, P2, P3, P4 are coplanar. Hence, if these conditions are not
fulfilled, P* does not exist.
(h) If, however, a point P* can be found which satisfies (5), any
point whatever will serve (§ 8). In fact, we see, from (2), that
the existence of P* implies that
(6) ,6dmiOPi = 0 1mi = 0,
for any choice of 0. In particular, we may take any one of the
given points Pi as origin. Thus, if we take 0 at P1, (6) becomes
V
miP1 Pi = 0, 2; mi = -m1 0.
i=2 i=2
Hence, from the defining equation (1), we see that P1 is the cen-
troid of the points m2P2, m3P3, . . . , mnPn. Precisely the same
conclusion may be drawn for P2, P3, , Pn. We now can answer

the question raised at the end of § 8. Any set of weighted points


miPi (mi F6 0) whose position vectors satisfy a linear relation (6)
independent of the origin has the intrinsic property that any point of
the set is the centroid of all the remaining weighted points.
A set of n points having this property may be readily con-
structed. Take any set of n - 1 weighted points miPi, such that
m1 + m2 + + mn_1 0, and adjoin to the set their centroid
P* = Pn with the weight Mn = -(m1 + m2 + + mn-1).
Then, since
-mnPn = m1P1 + m2P2 + + mn-1Pn-1,
we have
n n
X mipi = 0, 2;mi = 0.
1 t
22 VECTOR ALGEBRA §9

Thus the relation,


(7) as+Sb+yc=0, a+$+-y=0,
between three collinear points shows that each point of the set
aA, ,BB, yC is the centroid of the other two. Any point C on the
line of A and B satisfies a relation of the type (7) and hence is the
centroid of these points when suitably weighted.
The relation,
(8) as+,9b+yc+8d = 0, a+a+7+S = O,
between four coplanar points shows that each point of the set
aA, ,BB, yC, 8D is the centroid of the other three. Any point D
in the plane of A, B, C satisfies a relation of the type (8) and hence
is the centroid of these points when suitably weighted.
Example 1. Centroid of Two Points. The centroid of aA and pB is given by
* as+Rb
P
a+13
Hence P* divides AB in the inverse ratio f/a of their weights. In particular,
if a = S, P* divides AB in half. The mean center of two points lies midway
between them.
Example 2. Mean Center of Three Points A, B, C. Let L, M, N be the mid-
points of BC, CA, AB (Fig. 9a). Then the mean center P* of A, B, C is the
centroid of A and 2L, B and 2M, C and 2N. Hence (ex. 1) the segments
AL, BM, CN are all divided by P* in the ratio of 2/1.
If A, B, C are not collinear, they are the vertices of a triangle. Its medians
AL, BM, CN intersect at P* and are there divided in the ratio of 2/1.
C

N a M2
Fia. 9a FIG. 9b
Example 3. Mean Center of Four Points A, B, C, D. Let L1, L2; M1, M2;
N1, N2 be the mid-points of A B, CD; BC, DA; AC, BD, respectively (Fig. 9b).
To find P* we may replace A, B, C, D by 2L1 and 2L2, or by 2M1 and 2M2,
or by 2N1 and 2N2. Therefore P* is the mid-point of the segments L1L2,
M1M2, N1N2. The bisectors of the three pairs of opposite sides of the quad-
§ 10 BARYCENTRIC COORDINATES 23

rangle ABCD (which may be plane or skew) intersect at P* and are there
divided in half.
If A, B, C, D are not coplanar, they determine a tetrahedron. Let A', B',
C', D' be the mean centers of the triads BCD, CDA, DAB, ABC. Then P*
is the centroid of A and 3A', B and 3B', C and 3C', D and 3D', and hence
divides each of the segments AA', BB', CC', DD' in the ratio of 3/1. The
preceding result also shows that the bisectors of the three pairs of opposite
sides of the tetrahedron meet at P*.
Example 4. The sum of n vectors AiB1 (i = 1, 2, , n) is given by
Z(bi-a;) =nb*-na* = n(b* - a*)
where A* and B* are the mean centers of the initial points Ai and the terminal
points Bi, respectively; hence
2;A;Bi = n A*B*.
In particular,
-4 --
A1B1 + A2B2 = 2 A*B*,
where A*, B* are the mid-points of A1A2 and B1B2, respectively.
10. Barycentric Coordinates. If P is any point in the plane of
-3-3 --3
the reference triangle ABC, the vectors AP, BP, CP are linearly
dependent (§ 5) ; hence
«(p - a) + /3(p - b) + y(p - c) = 0,
p=as+9b+-ic
(1)
«+6 + y
The denominator is not zero; for, if
a+a+y=0, then aa+(3b+yc=0,
and A, B, C would be collinear, contrary to hypothesis. Thus P
is the centroid of the weighted points aA, $B, yC. The three
numbers a, /3, y are called the barycentric coordinates of P; as they
all may be multiplied by the same number without altering p,
their ratios determine P. The vertices of the reference triangle
have the coordinates A(1, 0, 0), B(0, 1, 0), C(0, 0, 1); the unit
point (1, 1, 1) is the mean center of A, B, C.
If P is a point in space and ABCD a reference tetrahedron, the
vectors AP, BP, CP, DP are linearly dependent (§ 5); hence
«(p-a)+$(pp-b)+y(p-c)+a(p-d) =0,
as+Ab+yc+Sd
(2) p
a+O+y+S .
24 VECTOR ALGEBRA § 12

The denominator is not zero; for, if


a+8+y+S = 0, then as+,lib+yc+Sd =0
and the points A, B, C, D would be coplanar, contrary to hypoth-
esis. Thus P is the centroid of the weighted points aA, $B, yC, W.
The four numbers a, 0, y, S are called barycentric coordinates of P;
as they all may be multiplied by the same number without alter-
ing p, their ratios determine P. The vertices of the reference tetra-
hedron have the barycentric co-
ordinates A(1, 0, 0, 0), B(0, 1, 0,
0), C(0, 0, 1, 0), D(0, 0, 0, 1);
the unit point (1, 1, 1, 1) is the
mean center of A, B, C, D.
11. Projection of a Vector. To
find the projection of a vector
Fla. 11 PQ upon a line x, a director plane
7r must be specified. Pass planes
through P and Q parallel to 7r and let them cut x in the points
P1i Ql (Fig. 11). Then the vector P1Q1 is the 7r-projection of PQ
upon x.
. If no director plane 7r is specified, we tacitly assume that IT is
perpendicular to x; the projection is then orthogonal.
Let the vectors PQ, QR and their sum PR have the projections
P1Q1, Q1R1 and P1R1 on the line x; then, since

P1Q1 + Q1R1 = P1R1,


the projection of the sum of two vectors on a line is equal to the sum
of their projection on this line.
12. Base Vectors. Let e1, e2, e3 be three linearly independent
vectors; they are then non-coplanar. If the vectors are drawn
from a common origin 0 (Fig. 12a), we may pass a closed plane
curve through their end points El, E2, E3. If this curve is viewed
from the side of its plane opposite to that on which 0 lies, the
order E1E2E3 defines a sense of circulation. If this sense is counter-
clockwise, the set e1, e2, e3 is said to be right-handed or dextral; for
it is then possible to extend the thumb, index, and middle fingers
of the right hand so that they have the directions of e1, e2, e3,
§ 12 BASE VECTORS 2'r

respectively. If the sense defined by E1E2E3 is clockwise the set


is said to be left-handed or sinistral.
Any vector u = PQ may be expressed as the sum of three vec-
tors parallel to el, e2, e3, respectively. For, if we construct a
parallelepiped on PQ as diagonal by passing planes through P and

R
Fia. 12a FIG 12b

Q parallel to e2 and e3, e3 and el, el and e2 (Fig. 12b), its edges
will be parallel to el, e2, e3, and

PQ = PQ1 + Q1R + RQ = PQ1 + PQ2 + PQ3


_ y
Since PQi is a scalar multiple of ei, say u`ei,
(1) u = ulel + u2e2 + u3e3,
where ul, u2, u3 are numbers called the components of u with re-
spect to the basis el, e2, e3. Their indices are not exponents but
mere identification tags; they are written as superscripts for reasons
given in Chapter IX. The vector u is often written [ul, u2, u3],
with brackets to enclose its components.
If u is the position vector OP, the components ui are called the
Cartesian coordinates of the point P with respect to the basis
e1, e2, e3-
A straight line upon which two directions are distinguished is
called an axis. One direction is called positive, the other negative.
In a figure the positive direction is marked with an arrowhead.
With a given basis, lines drawn through the origin 0 parallel to
26 VECTOR ALGEBRA § 13

el, e2, e3 and with their directions positive are called the coordi-
nate axes. The numbers u', u2, u3 often are called the components
of u on these axes.
The components of a zero vector are all zero; for, since el, e2, e3
are linearly independent, u = 0 implies u' = u2 = u3 = 0.
If A is any scalar,
(2) Au = Aulel + Au2e2 + Au3e3.
If v = vie1 + v2e2 +.v3e3,
(3) u + v = (u1 + v')el + (u 2 + v2)e2 + (u3 + v3)e3
With the bracket notation, (2) and (3) become
(4) A[ul, u2, u3] = [Au', Au2, Au3],
(5) [ul, u2, u3] + [v', v2, v3] = [u' + v1, u2 + v2, u3 + v3].
To multiply a vector by a number, multiply its components by that
number; to add vectors, add their corresponding components. In par-
ticular,
(6) -[u', u2, u3] = [-u', -u2, -u3],
(7) [ul, u2, u3] - [v', v2, v3] = [u' - v', u2 - v2, u3 - v3].

If u = v, then u - v = 0; hence
(8) U=V implies u' = v', u2 = v2, u3 = v3.
When referred to the same basis, equal vectors have their corresponding
components equal.
13. Rectangular Components. Let i, j, k denote a dextral sys-
tem of mutually perpendicular unit vectors. From an origin 0
draw the coordinate axes x, y, z with positive directions given by
i, j, k (Fig. 13). Any vector u now is determined by giving its
components u1, u2i u3 on these rectangular axes:
(1) u = uli + u2j + u3k.
Draw OP = u from the origin, and let OP1, OP2i OP3 be its
orthogonal projections on the axes. Then ul is the length of OPI,
taken positive or negative according as OPI has the direction of
i or -i; hence
u1 = I u I cos (i, u),
§ 13 RECTANGULAR COMPONENTS 27

where j u j denotes the length of u and (i, u) the angle between


i and u. The rectangular components of u thus are obtained by
multiplying its length by the corresponding direction cosines:
(2) ul = I u I cos (i, u), u2 = I U I cos (j, u),
u3 = I u I cos (k, u).

R
Fic. 13

The Pythagorean Theorem gives the length of u in terms of its


components; from
0P2 = (OR)2 + (RP)2 = (OPl)2 + (OP2)2 + (OP3)2,
(3) I U 12 = (ul)2 + (u2)2 + (u3)2.
If we substitute from (2) in (3), we obtain the relation,
cost (i, u) + cost (j, u) + cost (k, u) = 1,
satisfied by the direction cosines of any vector.
The rectangular coordinates of any point P are defined as the
components of its position vector OP on the x-, y-, and z- axes.
Thus, if

(4) OP=xi+yj+zk,
P has the rectangular coordinates (x, y, z). Since P1P2 = OP2
28 VECTOR ALGEBRA §13

- OPI, this gives

(5) PIP2 = (x2 - xl)i + (Y2 - yl)j + (z2 - zl)k


if (xi, y,, zi) are the coordinates of Pi. The components of PIP2
are found by subtracting the coordinates of PI from the corresponding
coordinates of P2.
With an orthogonal basis i, j, k, the components may be written
with subscripts (as previously) or superscripts. Separately, they
are called the x-, y-, and z- components and often are written
ul, U2, U3- Just as with a general basis, vectors are specified by
giving their components: u = -[ul, u2, u3]. Thus the point
(x, y, z) has the position vector [x, y, z]; and (5) may be written

PIP2 = [x2 - xl, Y2 - yl, z2 - Z11-


The unit vectors i, j form a basis for all vectors u in their plane:
(6) u = uli + u2j.
If e is a unit vector in the plane and 0 is the angle (i, e), reckoned
positive in the sense from i to j, we define cos 0 and sin 0 as the
components of e :
(7) e=icos0+jsin0.
Any vector in tine plane may be written

(8) u= Iule= Jul {icos(i,u)+jsin(i,u)};


its rectangular components are
(9) ul = I U I cos (i, u), u2 = I u I sin (i, u), u3 = 0.
Example. Addition Theorems for the Sine and Cosine. Let a and b be two
unit vectors such that the angles (i, a) = a, (a, b) = 0; then the angle (i, b)
= a + 14, and, from (7),
a =icosa+jsina,
b = i cos (a + fl) + j sin (a + 0).
If we refer b to the new basis i = a, j (a unit vector 90° ahead of a), we have
b=icos0+isin 0
= (i cos a + j sin a } cos l3 + I i cos (a a + 2) + j sin (a + 2)1 sin 6.
15 SCALAR PRODUCT 29

Comparing the components of b in the two preceding expressions, we find


cos (a + 13) = cos a cos 8 + cos (a + sin [3,
2)
sin (a + /3) = sin a cos l3 + sin (a + 2) sin
With a =7r/2, these equations give

cos (/3 + 2) sin $, sin (p + 2) = cos 0;


hence
cos (a + 0) = cos a cos 0 - sin a sin 0,
sin (a + )3) = sin a cos g + cos a sin g.
These addition theorems hold for all values of a and fl, positive or negative.
14. Products of Two Vectors. Hitherto we have considered
only the products of vectors by numbers. Next we shall define
two operations between vectors, which are known as "products,"
because they have some properties in common with the products
of numbers. These products of vectors, however, will also prove
to have properties in striking disagreement with those of numbers.
Since one of these products is a scalar and the other a vector,
they are called the scalar product and vector product, respectively.
The definitions of these new products may seem rather arbitrary
to one unfamiliar with the history of vector algebra. We present
this algebra in the form and notation due to the American mathe-
matical physicist, J. Willard Gibbs (1839-1903).t It is an offshoot
of the algebra of quaternions, adapted to the uses of geometry and
physics. In Chapter X quaternion algebra is developed briefly,
and the origin of the foregoing products is revealed.
15. Scalar Product. The scalar product of two vectors u and v,
written u v, is defined as the product of their lengths and the cosine
of their included angle:
(1) Jul IvI cos(u,v).
The scalar product is therefore a number which for proper (non-
zero) vectors is positive, zero, or negative, according as the angle
(u, v) is acute, right, or obtuse. Hence, for proper vectors,
(2) means u1v.
t Professor of mathematical physics at Yale University. His pamphlet on
the Elements of Vector Analysis was privately printed in 1881. A more com-
plete treatise on Vector Analysis (New Haven, Yale University Press, 1901)
based on Gibbs's lectures, was written by Professor E. B. Wilson.
30 VECTOR ALGEBRA 115
When u and v are parallel,
Jul lvl, or -Jul lvI,
according as the vectors have the same or opposite directions; thus

From (1) we see that

(3) (au) ($v) = ag u . V.


The last result is obvious when a and i3 are positive numbers; the
other cases then follow from the equations preceding.
Besides the components of a vector on the coordinate axes
(§§ 12, 13) we shall also use the orthogonal projection and compo-
nent of a vector on an arbitrary directed line 1. If the positive
direction of 1 is given by the unit vector e, the projection of u
upon 1 (proj 1 u) is a scalar multiple of e. This scalar is called the
component of u upon 1 (comp1 u) ; its defining equation is therefore
(4) e comp j u = proj 1 u.
As in § 13, we compute comp j u as
(5) comps u = l u i cos (e, u);
since I e 1, this may also be written
(6) comps u = e u.
From the projection theorem of § 11,
(7) proj 1 (u + v) = proj 1 u + proj 1 v;
hence, from (3),
(8) comps (u + v) = comp] u + comps v.
The operations expressed by proj1 and comp] are distributive with
respect to addition.
The definition (1) of u v now may be written
(9) U V= I u l compu V=IV l comp u.
Scalar or "dot" multiplication is commutative and distributive:
(10)
(11) w(u+v) =wu+wv.
§ 15 SCALAR, PRODUCT 31

The last equation is nothing more than (8) multiplied by


when 1 is taken along w.
If c 0 in the equation,
or
we can conclude either that a - b = 0 or that a - b and c are
perpendicular. We cannot "cancel" c to obtain a = b unless
a - b and c are not perpendicular.
It is obvious that, in general, (u v)w u(v w).
Since i, j, k are mutually perpendicular unit vectors,
(12) i
Hence, if we expand the product,
u ' v = lull + u2j + u3k) (v1i + v2j + v3k)
we obtain
(13) u v = u1v1 + u2v2 + u3v3
The scalar product of two vectors is equal to the sum of the products
of their corresponding rectangular components.
Example 1. If u = [2, -1, 3], v = [0, 2, 4], u v = 0 - 2 + 12 = 10.
Moreover,
comp,, u =
urv7
v =
10
= 2 236, compu v =
uuv = 10
= 2.673;
/20 luI 14
v 10
cos (u, v) = = 0.5979, angle (u, v) = 530 18'.
I
uU v /280 =
Example 2. Identities involving scalar products may be given a geometric
i.iterpretation. Thus (Fig. 15a)
gives cdcosw=a2-b2,
and, if we write c = a + b, d = a - b, w = an-
L,Ie (c, d). If a = b, c d = 0; then PQRS is a
rhombus and the angle PRT may be inscribed
in a semicircle about Q. We thus have two geo-
metric theorems:
1. The diagonals of a rhombus cut at right
angles.
2. An angle inscribed in a semicircle is a right P a Q
angle. Fro. 15a
Moreover, from
(a-b) (a - b) =
we have the cosine law: d2 = a2+0 - 2ab cos 0.
32 VECTOR ALGEBRA §15

Example 3. We also may interpret identities involving scalar products by


regarding the vectors a, b, as position vectors OA, OB, from an
arbitrary origin O. Then a - b = BA and a + b = 2 011 where M is the
mid-point of AB.
Consider, for example, the identity,
(b-a)2+(c-b)2+(d-c)2+(a-d)2
= (c-a)2+(d-b)2+(a+c-b-d)2,
which can be verified on expansion; u2 means u u. If we regard a, b, c, d
as the position vectors of the vertices of a space quadrilateral ABCD, p =
(a + c), q = (b + d) locate the mid-points P, Q of the diagonals AC, BD;
hence 2

(AB)2 + (BC)2 + (CD)2 + (DA)2 = (AC)2 + (BD)2 + 4(QP)2.


The sum of the squares of the sides of any space quadrilateral equals the sum of
the squares of its diagonals plus four
C times the square of the segment joining
their middle points.
Example 4. The identity,
(a-b) (h-c)
0,
B
Fia. 15b shows that the altitudes of a triangle
ABC meet in a point H, the ortho-
center of the triangle (Fig. 15b); for, if two terms of this equation are zero, the
third is likewise.
Similarly the identity,

(a-b) k- a +2 b \ b+c
2
c+a\
C

shows that the perpendicular bisectors of the triangle ABC meet in a point K,
the circumcenter of the triangle.
For the orthocenter H and circumcenter K the individual terms of the fore-
going equations vanish; for example,
(a - 0, (a - 0.

On adding these, we obtain


=0;
or on writing g = 3 (a + b + c) for the position vector of the mean center G
of the triangle ABC.
(a-b) (h+2k-3g) =0.
§ 15 SCALAR PRODUCT 33

Since this equation also holds when a - b is replaced by b - c and c - a,


we conclude that
h + 2k - 3g = 0.
Therefore the mean. center of a triangle lies on the line joining the orthocenter to
the circumcenter and divides it in the ratio of 2/1. This line is called the Euler
line of the triangle.
Example 5. In order to interpret the identity,
(a + b - c - d)2- (a - b - c+ d)2 = 4(a - c) - (b-d),
with reference to the plane quadrilateral ABCD, let P, Q, R, S denote the
mid-points of AB, BC, CD, DA; then (Fig. 15e) we have
(PR)2 - (QS)2 = CA DB.
Hence, if the diagonals of a quadrilateral cut at right angles, the lines joining
the mid-points of opposite sides are equal. Since the lines PR, QS intersect
D

P R'
Fta. 15c Fia.15d

in the mean center N of the points A, B, C, D and are bisected there (§ 9,


ex. 3), a circle with N as center will pass through P, Q, R, S. If perpendiculars
from P, Q, R, S are dropped upon the sides opposite, their feet, P', Q', R', S
also will lie on this circle. Thus we have proved the
THEOREM. When the diagonals of a quadrilateral are perpendicular, the mid-
points of its sides and the feet of the perpendiculars dropped from them on the
opposite sides all lie on a circle described about the mean center of the vertices.
In the figure formed by a triangle ABC and its three altitudes meeting at
the orthocenter H (Fig. 15d), three quadrilaterals, ABCH, BCAH, CABH,
all have perpendicular diagonals. Their three eight-point circles are all the
same. This circle, whose center is at the mean center N of A, B, C, H, is the
famous nine-point circle of the triangle.
THEOREM. For any triangle ABC whose orthocenter is H, a circle whose center
is the mean center of A, B, C, H, passes through the nine points: the mid-points
of the sides, the feet of its altitudes, and the mid-points of the segments joining H
to the vertices.
34 VECTOR ALGEBRA § 16

The center N of the nine-point circle has the position vector n given by
4n =a+b-}-c+h=3g+h=2h+2k,
in view of ex. 4. Therefore the center N of the nine-point circle is collinear
with the mean center G of the triangle, its orthocenter H, and circumcenter K.
Moreover N bisects the segment HK.
16. Vector Product. The vector product of two vectors u and v,
written u x v, is defined as the vector,
(1) uxv= Jul Ivlsin(u,v)e,
where e is a unit vector perpendicular to both u and v and forming
with them a dextral set u, v, e. If u and v are not parallel, a right-
handed screw revolved from u towards v will advance in its nut
towards u x v.
When u and v are parallel, e is not defined; but in this case
sin (u, v) = 0 and u x v = 0. Moreover, if u and v are not zero,
u x v = 0 only when sin (u, v) = 0. Hence, for proper vectors,
(2) uxv = 0 means u I I v.
In particular u x u = 0.
From (1) we see that
(-u)xV = ux(-v) _ -uxv, (-u). (-v) = uxv,
(3) (au) x ((3v) = a# u x v.
The last result is obvious when a and /3 are positive numbers; the
other cases then follow from the equations
preceding.
If u and v are interchanged in (1), the
scalar factor is not altered, but e is reversed;
hence
(4) vxu = -uxv.
Vector multiplication is not commutative.
FIG. 16 Draw u and v from the point A, and let p
be a plane perpendicular to u at A (Fig. 16).
Then u x v may be formed by a sequence of three operations:
(P) Project v on p, and obtain v';
(M) Multiply v' by I u I, and obtain I u I v';
(R) Revolve I u I v' about u through +90°
316 VECTOR PRODUCT
The resulting vector agrees with u x v in magnitude, for I V,
I v I sin (u, v), and also in direction (upward in the figure). We is
dicate this method of forming u X v by the notation,
(5) u X v = RMPV.
This means that v is projected, and the projection multiplied, an
finally revolved as previously described. Now each of these ope
ators is distributive: operating on the sum of two vectors is the sarr.
as operating on the vectors separately and adding the result:
hence
RMP(v + w) = RM(Pv + Pw)
= R(MPv + MPw) = RMPV + RMPvc
Thus, from (5),
(6) ux (v + w) = uxv + uxw, (V + W) Xu = vxu + wxu
Vector or "cross" multiplication is distributive. By repeated appli
cations of (6) we may expand the vector product of two vecto
sums just as in ordinary algebra, provided that the order of th,
factors is not altered. For example,
(a+b)x(c+d) =axc+axd+bxc+bxd.
If c 7d 0 in the equation,
axc=bxc, or (a-b)xc=0,
we can conclude either that a - b = 0 or that a - b and c are
parallel. We cannot "cancel" c to obtain a = b unless a - b and
c are not parallel.
We shall see in § 18 that in general, (u x v) x w P6 u x (v x w).
Since the unit vectors i, j, k form a dextral orthogonal set, we
have the cyclic relations.
(7) ixj=k, jxk=i, kxi=j; ixi=jxj=kxk=0.
Hence, if we expand the product,
u x v = (u1i + u2j + u3k) x (v1i + v2j + v3k),
we obtain
(8) u x V = (u2v3 - u3v2)i + (u3v1 - u1v3)j + (ulv2 - u2v1)k.
36 VECTOR ALGEBRA § 16

The components of u x v are the determinants formed by columns


ul U2 u3
°l and 3, 3 and 1 (not I and 3), 1 and l of the array
(VI V2 V3
hence we may write
i j k
(9) u1 U2 U3

V1 V2 V3

For example, if u = [2, -3, 5], v = [-1, 4, 2], we compute the


components of u x v from the array,
2 -3 5 -3 5 5 2 2 -3
= - 26, _ -9, =5.
C -1 42 ' 42 2 -1 -1 4

Thus u x v = [ -26, -9, 5 ]. As a check, we verify. that u x v is


perpendicular to both u and v: -52 + 27 + 25 = 0, 26 - 36 +
10 = 0.
Example 1. To find the shortest distance d from a point A to the line BC.
Method. Let e be the unit vector along BC. Then, if u is any vector from
A to the line (as AB or AC),
d = I uI sin(u,e) =Iuxe1.
Computation. If the points are A(3, 1, -1), B(2, 3, 0), C(-1, 2, 4),
a=[-3, -1,4]
u=AB=[-1,2,1], BC=[-3,-1,4], x/26
[9, 1, 7]
u x e = x/26 , d = IuxeI i s -_ -2.245

-4
Check. If we take u = AC = [-4, 1, 5], u x e again has the preceding value.

Example 2. To find the shortest distance d from a point A to the plane


BCD.
Method. Find a vector normal to the plane, such as BC x BD, and let n
be the unit vector in its direction. Then, if u is any vector from A to the
plane (as AB or AC), d = n u
Computation. If the points are A(1, -2, 1), B(2, 4, 1), C(-1, 0, 1),
D(-1, 4, 2),
-- 4
BC = [-3, -4, 0],
-,
BD = [-3, 0, 1],
----> ---4
BC-BD = [-4,3, -12];
-i
u=AB=[1,6,0]; n= [-4,3, -121 _ 14
13
-4
If we take u = AC = [-2, 2, 0], n u =.
Check.
17 VECTOR AREAS 37

Example 3. To find the shortest distance d between two non-parallel lines


AB, CD; and to locate the shortest vector PQ from AB to CD.
Method. Find the vector AB x CD which is perpendicular to both lines,
and let n be a unit vector in its direction. Then, if u is any vector from AB
to CD (as AC orBD), d = In-
To find P and Q, write AP = a AB, CQ = y CD, and find the scalars a, y
from the condition that PQ = PA + AC + CQ is parallel to AB x CD. The
length PQ = d.
Computation. If the lines AB, CD are given by the points A(1, -2, -1),
B(4, 0, --3); C(1, 2, -1), D(2, -4, -5);
-4 --4
= [1, -6, -4],
--> --->
AB x CD = 10[-2, 1, -2];
AB = [3, 2, -2], CD

n= 3[-2, 1, -21, AC = [0, 4, 0], d = n AC =


To find P and Q, we have
-- --4 -+ -4
PQ = -aAB +AC +yCD
= -a[3, 2, -2] + [0, 4, 0] + y[l, -6, -4]
= [-3a+y, -2a+4-6y,2a-4y];
and, since PQ is parallel to [-2, 1, -2],
-3a + y -2a + 4 - 6y 2a - 4y
-2 1 -2
These equations give a = y = $; hence

OP=OA+AP=[1,-2,-1]+$[3,2,-2]=$[21,-10,-17],
OQ = OC + CQ = [1, 2, -1] + $[1, -6, -4] _ $[13, -6, -25].
Check. The distance PQ = .

17. Vector Areas. Consider a plane area A whose boundary is


traced in a definite sense-shown by arrows in Fig. 17a. We shall

V
U
FIG. 17a Fia. 17b
associate such an area with a vector of magnitude A, normal to its
plane, and pointing in the direction a right-handed screw would move
if turned in the given sense. Thus the parallelogram whose sides
38 VECTOR ALGEBRA $ 17

are the vectors u, v (Fig. 17b) and whose sense agrees with the
rotation that carries u into v is associated with the vector u x v;
for its area is
A = Iul Ivlsin(u,v) = Iuxvl.
All plane areas associated with the same vector will be regarded
as equal. The sum of two plane areas associated with the vectors,
u, v, is defined as the plane
area associated with u + v.
A directed plane is a plane
associated with a definite nor-
mal vector, say the unit normal
n. A circuit in a directed plane
is positive when it is counter-
clockwise relative to n; a clock-
wise circuit is negative.
Consider now a plane area A
associated with the vector u.
The projection of A on a di-
Fm. 17c rected plane p is A' = A cos 8
(Fig. 17c).The circuital sense
of A is projected on A'. We shall give A' the sign corresponding
to its circuit on p, and call this signed area the component of A
on p. Since I u = A and (n, u) = 8 or ir - 0,
n- u = u I cos (n, u) = A cos 0 or -A cos 8
according as (n, u) is acute or obtuse. In both cases n u gives
the component of A on p:
(1) compp A = n u.
THEOREM. When the vector areas of the faces of a closed poly-
hedron are all drawn in the direction of the C
outward normals, their sum is zero.
Proof. Any polyhedron may be subdi-
vided by planes into a finite number of A
tetrahedrons. Let ABCD (Fig. 17d) be one FIG. 17d
such tetrahedron. If we imagine ABC to
lie in the plane of the paper while D is above the paper, the out-
ward vector areas of the triangular faces,
DAB, DBC, DCA, ABC,
§ 17 VECTOR AREAS 39

are given by one half of the respective vectors,

DA X DB, DB X DC, DC X DA, -AB-AC.


If we choose D as origin, the sum of these vectors is given by
axb+bxc+cxa - (b - a)X(c - a) = 0.
Thus the theorem is true for a tetrahedron.
Now write such an equation for all the tetrahedrons that make
up the polyhedron, and add the results. The vector areas over
all inner faces cancel, for each appears twice but with opposed
directions. The net result on the left is double the sum of the
outward vector areas of the polyhedron's faces. Since this sum is
zero, the theorem follows.
Consider now a polygon P1P2 . Ph of area A lying in a di-
.

rected plane of unit normal n (Fig. 17e), and suppose that the
circuit P1P2 . Ph is positive. In the D
figure n points up from the plane of
the paper, so that a positive circuit is
counterclockwise. Choose an origin 0
at pleasure above the plane (towards P
the reader), and let r1, r2, , rh ''
denote the position vectors of the ver-
tices. These vectors are the edges of P,
a pyramid having 0 as vertex and the Fla. 17e
given polygon as base. The triangular
faces of this pyramid have as their outward vector areas 2r1 x r2i
ire X r3, , irh x r1, while the outward vector area of the base is
-An. From the preceding theorem the sum of these vector areas
is zero; hence
(2) An = (r1 x r2 + r2 x r3 + ... + rh x r1).
i
As the vertex 0 approaches the plane of the base, the terms on
the right of (2) vary continuously, but their sum is always An.
Hence when 0 is in the plane of the polygon, (2) remains valid.
If 0 is any origin of rectangular coordinates in the plane, let the
vertices of the polygon be (x1, yl), (x2, y2), , (xh, yh) taken in
counterclockwise order. Now n = k, and
x1 x2
r1 X r2 = (x1i + yll) X (x2i + y2i) = k,
Y1 Y2
40 VECTOR ALGEBRA § 18

so that (2) gives

(3) 2A = I xl
x2 + I x2 x3 -+ -. ... + Xh
xl
Y1 Y2 I f y2 Y3 I I Yh Y1 J

Example 1. The sum of the outward vector areas of the triangular prism
(Fig. 17f) is
uxw +vXw - (u +v) xw = 0.
The vector areas of the triangular bases cancel, for
they are equal in magnitude but opposite in direction.
Example 2. To find the area A of a triangle
u whose vertices are (a, 0, 0), (0, b, 0), (0, 0, c), put
FIG. 17f r1 = ai, r2 = bj, r3 = ck in (2); then
An = (abk + bci + caj), A = 11/a2b2 + b2c2 + OA
i
Example 3. To find the area of the polygon whose vertices in counter-
clockwise order are (1, 2), (5, 4), (-3, 7), (-5, 5), (-1, -3), we form the
array,
(x) 1 5 -3 -5 -1 1

(y) 2 4 7 5 -3 2,
repeating the first column. Then, from (3),
2A = -6+47+20+20+1 =82, A =41.
18. Vector Triple Product. The vector (u x v) x w is perpendic-
ular to u x v and therefore coplanar with u and v; hence (§ 5),
(u x v) x w = au + /3v.
But, since (u x v) x w is also perpendicular to w,
aU 0.
All numbers a, that satisfy this equation must be of the form
a = -A v w, = X u w, where X is arbitrary. Thus we have
(uxv)xw =
In order to determine X, we use a special basis in which i is col-
linear with u, j coplanar with u, v; then
u = u1i, V = v1i + v2j, w = wli + w2j + w3k.
Substituting these values gives, after a simple calculation, A = 1
We therefore have the important expansion formulas,
(1) (uxv)xw = u-WV
wx(uxv) = w-vu -wuv.
§ 19 SCALAR TRIPLE PRODUCT 41

In the left-hand members of (1), one of the vectors in parenthesis


is adjacent to the vector outside, the other remote from it. The
right-hand members may be remembered as
(Outer dot Remote) Adjacent - (Outer dot Adjacent) Remote.
In general (u x v) x w -/- u x (v x w); for the former is coplanar
with u and v, the latter with v and w. Cross multiplication of vec-
tors is not associative.
From (1) we see that the sum of a vector triple product and its
two cyclical permutations is zero:
(2) (axb)xc -{- (bxc) x a + (cxa)xb = O.
If 1 is a directed line carrying the unit vector e, we may express
any vector u as the sum of its orthogonal projections on 1 and on
a plane p perpendicular to 1:
(3) projl u = e comps u = (e u) e (15.6),
(4) projp u = u - (e u) e = e x (u x e).
19. Scalar Triple Product. The scalar product of u x v and w
is written u x v - w or [uvw]. No ambiguity can arise from the pa-
rentheses being omitted, since u x (v w)
uXv
is meaningless.
THEOREM. The product u x v - w is
numerically equal to the volume V of a
parallelepiped having u, v, w as concur-
rent edges. Its sign is positive or negative
according as u, v, w form a right-handed
or left-handed set. Fic. 19
Proof. The volume V (Fig. 19) may be computed by multiply-
ing the area of a face parallel to u and v,
A = J u l H v l sin (u, v) = l u x e 1,
by the corresponding altitude h = I w I cos B:
V= IuxvIIwIcos6.

Now the angle between u x v and w is B or 7r - B according as


u, v, w form a dextral (as in the figure) or a sinistral set (§ 12).
The definition of a scalar product now shows that
V dextral.
(1) uxvw= when the set u, v, w is
V} sinistral.
42 VECTOR ALGEBRA § 19

The dextral or sinistral character of a set u, v, w is not altered


by a cyclical change in their order. Hence, from (1),
(2)
But a dextral set becomes sinistral and vice versa, when the cy-
clical order is changed :
(3)
Thus, if the set u, v, w is dextral, the products in (2) all equal V,
while the products,
U"W V =
all equal - V.
On account of the geometric meaning of the scalar triple product,
we shall call it the "box product." $
If u, v, w are proper vectors, V = 0 only when the vectors are
coplanar. Therefore: three proper vectors are coplanar (parallel to
the same plane) when and only when their box product is zero. In
particular, a box product containing two parallel vectors is zero;
for example u x v u = 0.
The value of a box product is not altered by an interchange of the
dot and cross. For
(4)
as dot multiplication is commutative. The notation [uvw] often
is used for the box product, as the omission of dot and cross causes
no ambiguity.
From (15.3) and (16.3), we have
(5) (au) x (av) . (tiw) = a(3y u x v w.
Finally, the distributive law for scalar and vector products
shows that a box product of vector sums may be expanded just
as in ordinary algebra, provided that the order of the vector fac-
tors is not altered. Thus, if we expand the product u x v w when
the vectors are referred to the basis i, j, k, we obtain 27 terms of
which all but six vanish as they contain box products with two
or three equal vectors. The remaining six terms are those con-
taining
[ijk] = [jki] = [kij] = 1, [ikj] = [kji] = [jik] = -1,
$ The name proposed by J. H. Taylor, Vector Analysis, New York, 1939,
p. 46.
3 20 PRODUCTS OF FOUR VECTORS 43

and constitute the expansion of the determinant,


261 U2 U3

(6) [uvw] = v1 v2 v3 .

W1 W2 W3

This result also follows at once from (16.8).


Example. To find the point P where the line AB pierces the plane CDE.
Method. AP = it AB; the scalar it is then determined by the equation,

(i) CP CD X CE = (CA + it AB) . CD x CE = 0


which expresses that P is coplanar with C, D, E.
Computation. With the points,
A(1, 2, 0), B(2, 3, 1); C(2, 0, 3), D(0, 4, 2), E(-1, 2, -2);
CP =CA+XAB =[-1,2, -3]+1`[1,1,1] =[7,-1,it+2,A-3];
CD xCE = [-2,4, -1]x[-3,2, -5] = [-18, -7,8];
hence, from (i),
-18(X - 1) -7(X+2)+8(x-3) =17x-20=0, 7;
OP = OA + X AB = [1, 2,0] - -4[1, 1, 1] = .[-3, 14, -20].
-> --4 ---,
Check. DP CD X CE = 0; for
DP = --[1,18,18],
20. Products of Four Vectors. Since
(a x b) (c x d) = a b x (c x d) = a (b d c - b c d),
(1)

x d) may be regarded as a triple product


(a x b) x
of a x b, c, d or of a, b, c x d. We thus are led to the two expan-
sions,
(2) (a x b) x (c x d) = [acd]b - [bcd]a = [abd]c - [abc]d,
which give, in turn, the following equation connecting any four
vectors:
(3) a[bcd] - b[cda] + c[dab] - d[abc] = 0.
44 VECTOR ALGEBRA § 22

When [abc] 0,
(4) [abc]d = [dbc]a + [adc]b + [abd]c
gives d explicitly in terms of the basis a, b, c.
21. Plane Trigonometry. If the vectors a, b, c form a closed
triangle when placed end to end (Fig. 21),
(1) a+b+c=0.
We shall denote the lengths of the sides by a, b, c and the interior
angles opposite them by A, B, C; then the
angles (b, c), (c, a), (a, b) are equal, respec-
tively,to7r- A,7r-B,,r-C.
From the identity,
a
FIG. 21 (a+b) (a+b) =
we obtain the cosine law for plane triangles:
(2) c2 = a2 + b2 - tab cos C.
Cyclical interchanges of the letters in (2) give two other forms of
this law.
On multiplying (1) first by a x, then by b x, we find
(3) bxc = cxa = axb.
Division by abc gives the sine law for plane triangles:
sin A sin B sin C
(4)
a b c

Since each product in (3) is


double the vector area of the
triangle, its
(5) Area = be sin A
2
= I ca sin B = 2 ab sin C.
22. Spherical Trigonometry.
Consider a spherical triangle
ABC on a sphere of unit radius, FIG. 22
and let a, b, c, denote the position
vectors of the vertices referred to its center (Fig. 22). The nota-
tion is so chosen that a, b, c, form a dextral set: then [abc] > 0.
22 SPHERICAL TRIGONOMETRY 45

Let a, /3, y denote the sides (arcs of great circles) opposite the
vertices, A, By C. The interior dihedral angles at these vertices
also are denoted by A, B, C. We shall consider only spherical
triangles in which the sides and angles are each less than 180°.
Now
(1)

(2) bxc = sin a a', cxa = sin a b', axb =


where a', b', c' are unit vectors. The vectors b' and c' are perpen-
dicular to the planes of c, a and a, by respectively, and include an
angle a' = it - A, the exterior dihedral angle at A. Moreover,
from (2),
(c x a) x (axb) [abc]
b'xc' _ a,
sin /3sin y sin f3sin y
a positive multiple of a. Hence if a', 0', y' denote the exterior
dihedral angles at A, By C, we have
(3) b' c' = cos all Ca= cos a' b' = cos y'
(4) b'xc' = sin a' a, c'xa' = sin#'b, a' b' = sin y'c.
Thus there is complete reciprocity between the vector sets
a, by c, and a', b', c', and also between the spherical triangles they
determine. While ABC has a, S, y for sides and a', (3', y' for ex-
terior angles, A'B'C' has a', 0', y' for sides and a, /3, y for exterior
angles. Since the vertices of one triangle are poles of the corre-
sponding sides of the other, the triangles are said to be polar.
From (2) and (4), we have
[abc] = sin a a a' = sin/3b - b' =
sin a' sin /3'b - by = sin y' c. c';
hence, on division,
[abc] _ sin a _ sin / _ sin y
(5)
[a'b'c'] sin a' sin S' sin y'
This is the sine law for spherical triangles.
Again, from (2), we have
sin /3 sin y by c' = (c x a) (axb) = (c a) (a b) - b c;
46 VECTOR ALGEBRA § 23

hence, from (1) and (3),


(6) cos a = cos 0 cos y - sin 0 sin y cos a'.
Similarly, from (4), we deduce
(7) cos a' = cos 0' cos y' - sin f3' sin y' cos a.
Equations (6) and (7) and the four others derived from them by
cyclical permutation constitute the cosine laws for spherical tri-
angles.
By replacing a', 0', y' in (5), (6), and (7) by 7r - A, 7r - B,
it - C we may express the sine and cosine laws in terms of the
sides and interior angles. Thus, we find, for the sine law,
sin a sin j sin y
sin A sin B sin C '
and, for the cosine laws,
cosa = cos,3cosy + sin $ sin -y cos A,
cos A = - cos B cos C + sin B sin C cos a.
In this version of the cosine laws, the structural similarity ex-
hibited by (6) and (7) is lost.
23. Reciprocal Bases. Two bases, el, e2, e3 and e', e2, e3, are
said to be reciprocal when they satisfy the nine equations:
el e'=1, el e2=0, el e3=0,
(1) e2 e' = 0, e2 e2 = 1, e2 e3 = 0,

e3 e' = 0, e3 e2 = 0, e3 e3 = 1.
The superscripts applied to the base vectors are not exponents,
but mere identification tags. By use of the Kronecker delta 6j, de-
fined as
when i=j
(2) ' 0 when i 5 j,
equations (1) condense to
(3) ei e' = Si (i, j = 1, 2, 3).
Consider the three equations in the first column of (1). The
second and third state that e' is perpendicular to both e2 and e3,
§ 23 RECIPROCAL BASES 47

that is, parallel to e2 x e3. Hence el = X e2 x e3; and, from the


first equation, 1 = X el e 2 x e3. We thus obtain

(4) el = e2 x e3 e2 = e3 x el e3 = el x e2
[ele2e3] [ele2e3] [ele2e3]
e2 and e3 being derived from el by cyclical permutation. From
the symmetry of equations (1) in the two sets e1, ei, we have also
e2xe3 e3xel elxe2
(5) el = [ele2e3] , e2 = [ele2e3] , e3 = [ele2e3] .

Thus, either basis is expressed in terms of the other by precisely


the same formulas.
From (4) and (5), we have
1
(e2 x e3) (e2 x e3) _
e el =
[ele2e3][ele2el] [eIe2e3][ele2e3]

on making use of (20.1) and equations (1); hence


(6) [ele2e3][ele2e3] = 1,

an equation which gives further justification for the name recip-


rocal applied to the sets. Since the box-products in (6) must have
the same sign, a basis and its
reciprocal are both right-
handed or both left-handed.
As to the orientation of re-
ciprocal sets, we see from (1)
that e' is perpendicular to the
plane of e2 and e3 in the di-
rection which makes an acute
angle with el. Similar state-
ments apply to e2 and e3. If
the vectors et, et are all drawn
from the same point 0 and cut
by a sphere s about 0 in the Fla. 23
points Ei, Et, respectively,
the three planes OEiE; cut a spherical triangle E1E2E3 from s; and
the three planes OEE' cut out a second spherical triangle E'E2E3
(Fig. 23). Either triangle is the polar of the other; for E', E2, E3
are poles of the great-circle arcs, E2E3, E3E1, E1E2i and similarly
for El, E2, E3.
48 VECTOR ALGEBRA § 24

On dividing the identity,


ei x (e2 x e3) + e2 x (e3 x el) + e3 x (el x e2) = 0 (18.2),
by [ele2e3], we obtain the relation,
(7) elxel +e2xe2 +e3xe3 = 0.
This has an interesting geometric interpretation. The great circle
EIEI is perpendicular to both great circles E2E3 and E2E3; it thus
contains the altitudes of both triangles through the vertex labeled 1.
Similarly, the great circles E2E2, E3E3 contain the altitudes of
both triangles through the vertices labeled 2 and 3. The planes
of the great circles EIEI, E2E2, E3E3 are perpendicular to el x el,
e2 x e2, e3 x e3, respectively; and these vectors, in view of (7), lie
in a plane p. The poles of this plane (the ends of the diameter of
s perpendicular to p) are points common to all of the great circles
EiEi. Therefore, the three altitudes EiEi of the polar triangles
EIE2E3, E1E2E 3 meet in a pole P of the plane p.
When a basis and its reciprocal are identical, the basis is called
self-reciprocal. The equations (3) then become
(8) ei - e; = bi;. j'
These equations characterize an orthogonal triple of unit vectors.
Hence a basis is self-reciprocal when and only when it consists of a
mutually orthogonal triple of unit vectors. The triple will be dex-
tral if [ele2e3] = 1 (§ 19), sinistral if [ele2e3] = -1. Thus i, j, k
and i, j, -k are typical dextral and sinistral orthogonal bases.
24. Components of a Vector. We now supplement the notation
of § 12 by writing the components of a vector u ui or ui according
as the basis is ei or et. The components ui are called contravariant,
the components ui covariant, for reasons given in § 148. Any vec-
tor now may be written in two forms:
(1) (2)u = ulel + u2e2 + u3e3i u = ulel + u2e2 + u3e3.
From (1) and (2), we obtain equations of the type u - e' = ul,
u - el = ul ; all six are included in
(3) (4) ui = u ei, (i = 1, 2, 3).
ui = u - ei
When a basis is given, a vector u is completely specified by
giving its components written in order. With the notation of
t With a self-reciprocal basis there is no need for superscripts; we therefore
write S as Sii.
§ 24 COMPONENTS OF A VECTOR 49

§ 12, we write a vector u as a number triple, [ul, u2, u31 or


[ul, U2, u31 according as it is referred to the basis ei or et. We
denote the (non-zero) box products of the base vectors by
(5) E = [ele2e3l, 1/E = [ele2e31.
Addition and multiplication by scalars follow the formulas of
§ 12; thus, with covariant components,
(6) X [u1, u2, u31 = [Xu1, Xu2, Xu3]

(7) [u1, u2, u31 + [v1, v2, u31 _ [u1 + v1, u2 + v2, u3 + vJ
A simple expression for the scalar product u v is obtained when
u and v are referred to different, but reciprocal, bases. Thus, if
we compute
J (u'e1 + u2e2 + u3e3) . (vie1 + v2e2 + v3e3)
u . V = (ulel
l + u2e2 + u3e3) - (vlel + v2e2 + v3e3)
we obtain, by virtue of equations (23.1),
(8) u V = u1V1 + u2V2 + u3e3 = u1v1 + u2e2 + u3V3.
We next compute the components of u x v, relative to bases e2
and ei. Using covariant components gives
u x v = (uie' + u2e2 + u3e3) x (v1e1 + v2e2 + v3e3)
U2 U3 u3 u1 u1 U2
e2 x e3 + e3xel + e1xe2
V2 V3 V3 V1 V1 V2

U2 U3 e1 U3 u1 e2 e3

V2 V3 V3 V1 E E
or, more compactly,

(9) u x v = E-1

Using contravariant components, we find, in similar fashion,

(10) uxv=E
50 VECTOR ALGEBRA § 25

From (9) and (10), we see that the components of q = u x v are


Uj uk I
u' 11k 1

(11) q' = E-1 qt = E


Vi vk v' ak

where the indices ijk form a cyclical permutation of 123.


From (9) and (10), we next obtain two expressions for the box
product u x v w:
U1 u2 U3 u1 u2 u3

(12) [uvw] = E-1 V1 V2 V3 =E V1


v2 v3

W1 w2 w3 W1 w2 w3
The first formula of (12) may be written

(13) [uvw][e1e2e3] = v el v . e2 v - e3 f

whereas the second gives the analogous equation obtained by re-


placing ei by ei. Since e1i e2, e3 may be any linearly independent
set of vectors a, b, c, we also have
ua uc
(14) [uvw][abc] = va vb vc
wa wb wc
When the basis ei is self-reciprocal (an orthogonal triple of unit
vectors) e i = ei and E = +1. The two sets of components of u
then coalesce into a single set which we arbitrarily write U. If
the self-reciprocal basis is dextral, we can write el = i, e2 = j,
e3 = k, E = 1; then
U V = U1V1 + U2V2 + u03 ,

U X S7 = [uvw] =

in agreement with our previous results.


25. Vector Equations. A vector is uniquely determined when
its scalar and vector products with two known non-perpendicular
vectors are given. Thus let
(1) ua=a, uxb=c, (ab-/- 0);
§ 26 HOMOGENEOUS COORDINATES 51

the scalar a and the vectors a, b, c are regarded as known, and


b c = 0. We have
ax(uxb) =a -c, or

(2)
ab
By direct substitution, (2) is seen to be a solution of equations (1).
A vector is also determined when its scalar products with three
known non-coplanar vectors are given. For example, if
(3) u - e1 = u1, u - e2 = u2, u - e3 = u3,
and the sets et, e i are reciprocal, we have, from (24.2),
(4) u = ulel + 112e2 + u3e3.
By direct substitution, (4) is seen to be a solution of equations (3).
26. Homogeneous Coordinates. Coordinates of a point, line or
plane are called homogeneous if the entity they determine is not
altered when the coordinates are multiplied by the same scalar.
A coordinate that depends upon the choice of origin 0 bears the
subscript 0; such coordinates are written after those independent
of the origin. In equations r denotes any position vector from 0
to the entity in question, whereas r1i r2 denote position vectors
to points R1, R2 given in advance. In this article we use Latin
letters for vectors, Greek for scalars.
Point Coordinates. If r = OA, we write r = ao/a. The "equa-
tion" of the point A is then
(1) ra = ao,

and its homogeneous coordinates are (a, ao). Note that (Xa, Xao)
determine the same point; hence the coordinates (a, ao) depend
on three independent scalars: "there are o03 points in space."
Plane Coordinates. The equation of a plane through R1 and
perpendicular to the vector a is (r - r1) a = 0; or, on writing
ao=r1-a,
(2) r a = ao.
The homogeneous coordinates of the plane are (a, ao). Since
(Xa, Xao) determine the same plane, the coordinates (a, ao) de-
pend on three independent scalars: "there are oo 3 planes in space."
52 VECTOR ALGEBRA § 26

Line Coordinates. The equation of a line through Rl and par-


allel to the vector a is (r - r1) x a = 0; or, on writing ao = rl x a,
(3) r x a = ao.
The homogeneous (or Plucker) coordinates of the line are (a, ao)
and are connected by the relation,
(4) a ao = 0.
In view of. (4) and the fact that (Xa, Xao) determine the same line,
vile coordinates (a, ao) depend on four independent scalars: "there
are ooa lines in space."
We thus have the homogeneous coordinates:
Point (a, ao), Plane (a, ao), Line (a, ao).

The first coordinate cannot vanish. When the second coordinate


(subscript 0) is zero, the origin 0 is on the point, plane, or line,
respectively.
The distances of the point, plane, or line from the origin are,
respectively,
ao! laol aol
(5)
Iai' Jai ' Ia!
The first result is obvious. If p is the vector from 0 perpendicular
to the plane r a = ao,
p=

p ao a p is the vector from


0 perpendicular to the line r x a = ao,
axa0
p x a = ao, p a= 0; hence p=
from (25.2), and I p I = ( ao I/I a I
If the origin is shifted from 0 to P, we must replace r in (1),
(2), (3) by PO + r to obtain the second coordinate referred to P.
We thus obtain the shift formulas:
(6) ap = ao + P0 a,
(7) ap =
-3
(8) ap=ao+POxa,
in the respective cases.
§ 26 HOMOGENEOUS COORDINATES 53

Two Elements Determine a Third. We have the following cases:


Two distinct points determine a line:
(9) (a, ao), (a, bo) (abo - $ao, ao x bo).
Two non-parallel planes determine a line:
(10) (a, ao), (b, $o) --> (a x b, 00a - aob).
A point and line determine a plane:
(11) (a, ao), (b, bo) -> (abo - aoxb, ao bo),
unless abo - ao x b = 0; then also ao bo = 0, and the point lies
on the line.
A line and plane determine a point:
(12) (a, ao), (b, $o) -* (a b, 0oa - aoxb),
unless a b = 0. If 0oa - ao x b = 0 (and hence a b = 0), the
line lies in the plane.
Proofs in outline.
For (9) : The line AB has the equation,
as as
Cr x( = 0, or r x (abo - gao) = ao x b0.
/ a J
For (10) : The line is parallel to a x b; and
rx(axb) = $oa -aob.
For (11) : Refer the line (b, bo) to the point A(a, ao); then,
from (8),
aoxb
bA=bo - OAxb=bo -
a
The plane through A normal to bA has the equation

(r -aoa- (abo - ao x b) = 0, or r (abo - ao x b) = ao bo.

For (12) : The equations of line and plane, r x a = ao and r b


13o, have the solution (§ 25) :

r=-Qoaa- aob x b
54 VECTOR ALGEBRA §26

Two Lines. If Rl and R2 are points on the lines (a, ao), (b, bo),
the lines are coplanar when and only when the vectors r1 - r2,
a, b are coplanar. Since
(13) (r1

a necessary and sufficient condition that the lines be coplanar is


(14)
If the lines are not coplanar, let e be a unit vector in the direc-
tion a x b of their common normal; then, from (13),
a bo + b ao
(15)
IaxbI

This is the component of R2R1 in the direction of e; its numerical


value gives the shortest distance between the lines.
Equations in Point and Plane Coordinates. If the point (a, so)
lies on the plane (t, To) we have
(16)
on eliminating r from ra = so and r t = To. This may be re-
garded as the equation of the point (a, so) in plane coordinates; or
as the equation of the plane (t, To) in point coordinates.
If the line (p, po) contains the point (a, so) we have
(17) soxP -apo = 0,
on eliminating r from ra = so, r x p = po. This is the equation of
the line (p, po) in point coordinates.
If the line (p, po) lies on the plane (t, TO) we have p t = 0 and
hence
(18) Top - Po x t = 0,
on eliminating r from r x p = po and r t = To. This is the equa-
tion of the line (p, po) in plane coordinates.
Example. If three points (a, ao), (S, bo), (y, co) lie on the line (p, po), we
have, from (17),
ao bo co
-xP = 16
XP = _XP = po;
Y
hence

(i) y)+µ(0 y) =0,


§ 27 LINE VECTORS AND MOMENTS 55

since both vectors in parenthesis are parallel to p. This is a linear relation


between ao/a, bo/)3, co/y in which the sum of the coefficients is zero (§ 6).
If the three planes (a, ao), (b, ,o), (c, yo) pass through the line (p, po), we
have, from (18),
a b c
POX-=POX-=Pox-=P,
ao lio yo
provided ao, 00, yo are not zero; hence
Op

\ao yoi + µ \ljo -Yo -


since both vectors in parenthesis are parallel to po. This is a linear relation
connecting a/ao, b/3o, C/-Yo in which the sum of the coefficients is zero.
A shift in origin does not alter the coefficients in (i). In (ii), however, the
coefficients are changed but their sum still remains zero; if we write v =
-A - IA and shift the origin to P,

X o+ AA v-=
Po -+ -Yo
0 becomes X' p +lz'-+v
op 'YP
-=0,

where X' = Xap/ao, u' = -flP/0o, d = vyp/yo Let the student prove that
x'+A '+v'=0.
27. Line Vectors and Moments. A vector which is restricted to
lie in a definite line is called a line vector. If f is the vector, and
(1) r x f = fo, (f fo = 0)
the equation of its line of action, the line vector is completely
specified by the Pliicker coordinates f, fo.
The vector fo is called the moment of f about the point 0. If
f=AB,
(2) fo = OA x AB
is twice the vector area of the triangle OAB (§ 17); fo remains con-
stant as f is shifted along its line of action.
If the origin is shifted from 0 to P, the moment of f about P
is PA x AB = (PO + OA) x AB; hence

(3) fp = fo + PO x f.
If s is any axis through P with the unit vector e, the component
of fp on s, namely, e fp (15.6), is called the moment of f about the
axis s. We speak of moment about an axis because e fp is inde-
56 VECTOR ALGEBRA § 27

pendent of the position of P on this axis; for, if the axis is given


by the unit line vector (e, eo),

(4)

e fp as the component of twice the vector


area PAB on a directed plane of unit normal e (§ 17). In Fig. 27,
A1B1 is the projection of AB on the plane;
the moment of f about s is numerically
equal to twice the area PA1B1, that is, to
the product of the length A1B1 and the per-
pendicular distance h of A1B1 from P; its
sign is plus or minus according as a turn in
the sense PA1B1 would advance a right-
handed screw in the direction of s or -s.
We repeat these important definitions.
The moment of the line vector f = AB
about a point P is
A
Fro. 27 (5) fp=PAxAB=rxf,
where r is a vector from P to any point on f's line of action. The
moment of f about an axis is the component of r x f on the axis,
where r is a vector from any point on the axis to any point on the
line of action. The moment about a point is a vector; about an
axis, a scalar.
Example. The line of action of the force f = [1, -1, 2] passes through the
point A (2, 4, -1). Find its moment about an axis through the point
0
P(3, -1, 2) and having the direction of the vector [2, - 1, 2].
Since PA = [-1, 5, -3] and e = 3[2, -1, 21 is a unit vector along the axis,
i i k
fp = PA x f = -1 5 -3 = [7, -1, -4]
1 -1 2

is the moment of f about P; and

is the moment of f about the axis.


§ 28 SUMMARY: VECTOR ALGEBRA 57

28. Summary: Vector Algebra. Equal vectors have the same


length and direction. Free vectors are added by the triangle con-
struction. The negative of a vector is the vector reversed. To
subtract a vector, add its negative. The product Au is a vector
A I times as long as u; its direction is the same as u if X > 0, the
reverse if X < 0.
Vectors may be added, subtracted, and multiplied by real num-
bers in conformity with the laws of ordinary algebra.
The n vectors ui are linearly dependent if there exist n real num-
bers Ai, not all zero, such that EAiui = 0. When n = 2, 3, linear
dependence implies that the vectors are collinear or coplanar, re-
spectively; and conversely. In space of three dimensions any four
vectors are linearly dependent.
A point P divides a segment AB in the ratio X = /3/a when
----3 -3
AP = APB; then
-3 --> -3
(a + $)OP = aOA + SOB.
----3
A linear relation2;A1OP1 = 0 connecting n position vectors will
hold for any origin when and only when :Ai = 0. When n =
2, 3, 4, the points Pi are coincident, collinear, or coplanar, re-
spectively.
A set of weighted points miPi has a unique centroid P if 2;mi
$ 0; its defining equation is
--3
lmiP*Pi = 0;
-*
and OP* =
XmiOPi
Emi
If all mi = 1, P* is called the mean center.
The scalar product u v is defined as
Jul Ivlcos(u,v).
Laws:
V. Up U. (v+w) =
uv I
uv
uv=lu vl

is a u v
e a dextral set.
58 VECTOR ALGEBRA § 28

Laws :
UXV = `VxU, Ux(V+w) = uxv+uxw.
If u, v 0, u x v = 0 implies that u l v and conversely.
l

Expansion rule:
ux(vxw) =
Cross multiplication is not associative.
The box product, u x v w or [uvw], is numerically equal to the
volume of a "box" having u, v, w as concurrent edges; its sign is
+ or - according as u, v, w form a dextral or sinistral set. The
value of u x v w is not affected by a change in cyclical order of
the vectors, or by an interchange of dot and cross. If u, v, w 5-4- 0,
[uvw] = 0 implies that u, v, w are coplanar, and conversely.
A set of three vectors ej forms a basis if E = [ele2e3] 0. To
every basis there corresponds a unique basis ei such that ej e'
= S1; such bases are called reciprocal. If the indices i, j, k form
a cyclical permutation of 12 3,
e,xek ei x ek
ei = ei = [ele2e3] ; [ele2e3][e'e2e3] = 1.
[e1e2e3]

Both bases are dextral if E > 0; sinistral, if E < 0.


Given a basis e1, a vector u may be written as Euiei or ruiei;
the numbers ui are contravariant components of u, ui covariant
components.
u+v = E(ui'+vi')ei = E(ui+vt)ei;
Xu = 2:xuiei = 2;xuie`;
uV= uIv1 + u2v2 + u3v3 = u1v1 + u2v2 + u3v3;
e2 e3
el e2 e3 I I
el

U x v = E-1 U2 E ul
V3 v2 v3
Ut
vl
U1 u2 U3 u2 u3
[uvw] = E-1 I V1 V2 V3 E 1vi I

W1 W2 W3 w2
When the basis ej is self-reciprocal (ei = ei), its vectors
w3 form a
mutually orthogonal set of unit vectors; E = 1, if the basis is dex-
tral, E = -1, if sinistral. A dextral self-reciprocal basis is written
PROBLEMS 59

i, j, k. For such a basis ui = ui; these rectangular components are


given by
ui = I u I cos (ei, u), and = (ul)2 + (u2)2 + (u3)2.
I U I2

The two formulas previously given for u + v, Au, u - v, u x v,


u x v w in each case become identical.
The equations of a point, line, and plane, in terms of their
homogeneous coordinates (a, ao), (b, bo), (c, -yo), are
ra=ao, rxb=bo, r - c=yo,
respectively. When the origin is shifted to P,

ap = ao+POa, by = bo+POxb, yP =
The moment of the line vector (f, fo) about the point P is
fp = r x f, where r is any vector from P to its line of action. The
moment of (f, fo) about an axis through P is the component of fp
on this axis; if the axis is given by the unit line vector (e, eo), this
axial moment

PROBLEMS
1. If ABC is any triangle and L, M, N are the mid-points of its sides, show
that, for any choice of 0, a
-- 3 -3 -3 -4
OA + OB + OC = OL + OM + ON.
--3 -3
--- ---, --> -9
2. If OA' = 3 OA, OB' = 2 OB, in what ratio does the point P in which
AB and A'B' intersect divide these segments?
3. Show that the mid-points of the four sides of any quadrilateral (plane
or skew) form the vertices of a parallelogram.
4. P and Q divide the sides CA, CB of the triangle ABC in the ratios
--4 -3
x/(1 - x), y/(1 - y). If PQ = X AB, show that x = y = X.
5. E, F are the mid-points of the sides A B, BC of the parallelogram A BCD.
Show that the lines DE, DF divide the diagonal AC into thirds and that AC
cuts off a third of each line.
6. OAA', OBB', OCC', ODD' are four rays of a pencil of lines through 0
cut by two straight lines ABCD and A'B'C'D'. If C and D divide AB in
the ratios r and s, C' and D' divide A'B' in the ratios r' and s', prove that
r/s = r'/s'.
7. The points A, B, C and A', B', C' lie, respectively, on two intersecting
lines. Show that BC', CB'; CA', AC'; AB', BA' intersect in the collinear
points P, Q, R (Pascal's Theorem).
60 VECTOR ALGEBRA
8. Lines drawn through a point P and the vertices A, B, C, D of a tetra-
hedron cut the planes of the opposite faces at A', B', C', D'. Show that the
sum of the ratios in which these points divide the segments PA, PB, PC, PD
is -1. [Equation (10.2), with e = - (a + 0 + y + S) may be written
as+8b+yc+ad+ep =0, a+i9+y+s+e =0.
From this we conclude, as in § 7, ex. 3, that a' _ (aa + ep)/(a +,E), etc.]
9. The line DE is drawn parallel to the base AB of the triangle ABC and
is included between its sides. If the lines AE, BD meet at P, show that the
line CP bisects AB.
10. The points P, Q, R divide the sides BC, CA, AB of the triangle ABC
in the ratios a/(1 - a), l3/(1 - 0), y/(1 - y). If P, Q, R are collinear,
xp+yq+zr=0, x+y+z=0;
putting p = (1 - a)b + ac, etc. in this equation, we obtain a linear relation
in a, b, a whose coefficients have a zero sum. Show that this implies that the
separate coefficients vanish; hence deduce the Theorem of Menelaus (§ 7, ex. 2):
afy/(1 - a) (1 - p)(1 - y) = -1.
11. Using an argument patterned after that in Problem 10, prove Carnot's
Theorem:
If a plane cuts the sides AB, BC, CD, DA of a skew quadrilateral ABCD
in the points P, Q, R, S, respectively, the product of the ratios in which
P, Q, R, S divide these sides equals 1.
12. Id a plane quadrilateral ABCD, the point P in which the diagonals
AC, BD intersect divides these segments in the ratios 4/3 and 2/3, respec-
tively. In what ratio does the point Q, in which the sides AB, CD meet,
divide these segments? 0

13. If el = O 1, e2 = OE2i e3 = 0 3E
form a basis, the reciprocal set
el, e2, e3 are vectors perpendicular to the planes OE2E3, OE3E1i OE1E2 and
having lengths equal to the reciprocals of the distances of El, E2, E3 from
these planes, respectively. Prove this theorem.
14. Prove that a necessary and sufficient condition that four points
A, B, C, D be coplanar is that
[dbc] + [add] + [abd] - [abc) = 0.
15. Show that the shortest distance from the point A to the line BC is
Iaxb +bxc +cxal/Ib - cI.
16. If the vectors el, e2, e3; el, e2, e3 form reciprocal sets, show that
the vectors e2 x e3, e3 x el, el x e2; e2 x e3, e3 x e1, el x e2 do likewise.
17. Prove the formulas:
(a) [a x b, bxc, c x a] = [abc]2,
(b) (bxc) (a x d) +(c x a) (b x d) +(a x b) (c x d) =0,
(c) (bxc) x (a x d) +(c x a) x (b x d) +(a x b) x (c x d) = -2[abc]d,
(d) (d-a) (b-c)+(d-b) (c-a)+(d-c) (a-b)=0,
(e) (a-d) x (b-c)+(b-d) x (c-a)+(c-d) x (a-b) =2(a x b+b x c+c x a).
PROBLEMS 61

18. Find the shortest distance between the straight lines AB and CD when
(a) A(-2, 4, 3), B(2, -8, 0); C(1, -3, 5), D(4, 1, -7),
(b) A(2, 3, 1), B(0, -1, 2); C(1, 2, 5), D(-3, 1, 0).
19. Show that the lines AB, CD are coplanar, and find the point P in which
they meet:
A(-2, -3, 4), B(2, 3, 0); C(-2, 3, 2), D(2, 0, 1).
--4
[CD is parallel to CP = AP - AC = XAB - AC; find X and then P from
--->
OP = OA + XAB.]
-4
20. If the points P, Q, R divide the sides BC, CA, AB of the triangle ABC
in the same ratio, show that A, B, C and P, Q, R have the same mean center.
21. If the vectors OA, OB, OC lie in a plane and are equal in length, prove
that
OA + OB + OC = OH,
where H is the orthocenter of the triangle ABC. [§ 15, ex. 4.]
22. The power of a point P with respect to a sphere of center C and radius r
is defined as (CF)2 - r2. Prove that the power of P with respect to a sphere
having AB as diameter is PA I'B.
23. Prove that the sum of the n2 powers of n given points, P1, P21 , P,,,
with respect to the n spheres having for diameters the n segments joining the
given points to a variable point P in space, is constant. [Am. Math. Monthly,
vol. 51, p. 96.1
24. The lines (a, ao), (b, bo)'are coplanar; then a bo + b ao = 0 (26.14).
Show that they determine the plane (a x b, ao b) if a x b 0 0; and the point
aoxbo) if 00.
Solve Problem 19 using these results.
25. Show that the three planes (a, ao), (b, $o), (c, yo) meet in the point,
([abc], aobxc +Rocxa +yoaxb),
provided [abc] 0 0.
26. Show that the three points (a, ao), (0, bo), (y, co) determine the plane,
(a bo x co + 0 co x ao + y ao x bo, [aobocol),
provided aboxco +3coxao +yaoxbo 0 0.
27. The points P, Q, R divide the sides BC, CA, AB of the triangle ABC
in the ratio of 1/2. The pairs of lines (AP, BQ), (BQ, CR), (CR, AP) inter-
sect at X, Y, Z, respectively. Show that the area of the triangle XYZ is
1 /7 of the area of ABC. [Twice the vector area of XYZ is y x z + z x a +
xxyl
When P, Q, R divide the sides in the ratio t/1, the area of XYZ is
(1 - t)2/(1 + t + t2) times the area of ABC.
62 VECTOR ALGEBRA
28. Using (26.16), prove that
(a) If four points (a, ao), (0, bo), (y, co), (S, do) lie on a plane, the vectors
ao/a, , do/S are connected by a linear relation in which the sum of the
coefficients is zero;
(b) If four planes (a, ao), (b, 3o), (c, yo), (d, So) pass through a point, the
vectors a/ao, , d/So are connected by a linear relation in which the sum
of the coefficients is zero, provided ao, Qo, 'yo, So 54 0.
29. If a sphere S with a fixed center cuts two concentric spheres S1, S2i
prove that the distance between the planes of intersection SS, and SS2 is
independent of the radius of S. Extend this theorem to cover the case when
S fails to cut one or both of the concentric spheres. [Consider the radical
planes SS1, SS2.1
30. Prove that the radical planes of the three spheres (r - c1)2 = P1 (1 =
1, 2, 3) meet in the line
J
`C1 X C2 + C2 X C3 + C3 X c1, 2c1(e2 - e3 - P2 + A3) + cycl.;,
their radical axis. Consider the case when e1 X C2 + C2 X C3 + C3 X cl = 0.
31. A plane system of forces Fi acting through the points ri is equivalent
to a single force F = 2;Fi. When all the forces Fi are revolved about these
points through an angle B, show that their resultant F also revolves through
B about the point f (the astatic center) given by
F X M - (11ri Fi)F
t= F
2 where M = 'ri x Fi.

When the forces are parallel (Fi = Xie), show that the astatic center is the
centroid of the weighted points Xiri.
32. P* is the centroid of a set of n weighted points miPi for which Tmi = in.
Prove the Theorems of Lagrange:

(a) 1mi(OPi)2 = m(OP*)2 + 2:mi(P*Pi)2.

(b) Vmim1(PPi)2 = in mi(P*Pj)2,


where ij ranges over 2n(n - 1) combinations.

[OPi = OP* + P*Pi; PiP7 = P*Pi - P*Pi.]


33. Deduce the following results from Prob. 32.
(a) If ABCD are the vertices of a square in circuital order, (OA)2 + (OC)2
_ (OB)2 + (OD)2 for any 0. Generalize.
(b) If r is the radius of a sphere circumscribed about a regular tetrahedron
of side a, r2 = 3x2/8.
(c) The mean square of the mutual distances of all the points within a
sphere of radius r is 6r2/5.
34. If P* and Q* are the mean centers of p points Pi and q points Q re-
spectively, prove that
Mean (PA)2 = Mean (P*Pi)2 + Mean (Q*Q1)2 + (P*Q*)2.
CHAPTER II

MOTOR ALGEBRA

29. Dual Vectors. The vector f, bound to the line whose equa-
tion, referred to the origin 0, is
(1) rxf = fo,
is completely determined by the two vectors f, fo, its Plucker co-
ordinates. These obviously satisfy the relation,
(2) f fo = 0.
The vector f does not depend upon 0; but fo, the moment of f
about 0, becomes
(3) fp=fo+POxf,
when the origin is shifted from 0 to P.
We now amalgamate f and fo into a dual vector.
(4) F=f+efo
where a is an algebraic unit having the property e2 = 0. If
f = 1, F is called a unit line vector. The unit line vectors,
A=a+eao,
stand in one-to-one correspondence with the c0 4 lines of space.
A line vector F of length X always may be written F = XA, where
A is a unit line vector. Unit line vectors depend upon four in-
dependent scalars, general line vectors upon five.
Finally, for applications in mechanics, we consider dual vectors
F without the restriction (2). The vectors f, fo then involve six
independent scalars. The dual vector (or the entity it represents)
then is called a motor, provided its resultant vector f is independent
of the choice of 0, while its moment vector fo changes in accord-
ance with (3) when the origin is shifted to P.
Line vectors are thus special motors for which f fo = 0; for
unit line vectors also f f = 1.
63
64 MOTOR ALGEBRA § 30

30. Dual Numbers. In analogy with the complex numbers


x + ix', W. K. Clifford introduced dual numbers x + ex', in which
x, x' are real and e is a unit with the property e2 = 0.
In x + ex', x is called the real part and x' the dual part. We
write
x-{-ex'=y+ey' when x=y,x'=y';
x+ex'=0 when x = 0, x' = 0.
Addition and multiplication of dual numbers are defined by the
equations :
(1) (x+ex')+(y+ey') =x+y+e(x'+y'),
(2) (x + ex') (y + ey') = xy + e(xy' + x'y)
Observe that (2) may be obtained by distributing the product on
the left and putting e2 = 0. From these definitions we see that
addition and multiplication are commutative and associative and
that multiplication is distributive with respect to addition. In
fact, the formal operations are precisely those of ordinary algebra
followed by setting e2 = e3 = . . . = 0.
The negative of x + ex' is defined as -x - ex'.
If A = a + ea', B = b + eb', the difference X = A - B and
quotient Y = A/B satisfy, by definition, the equations B + X
= A, BY = A. We find that
(3) A - B = a - b + e(a' - b');
and that BY = A has the unique solution,
A a a'b - ab'
(4) B=b+ e b2

when and only when b F6 0. Division by a pure dual number eb'


is not defined. The quotient (4) may be remembered by means of
the device (a + ea') (b - eb')/ (b + eb') (b - eb') used in complex
algebra.
A dual number a + ea' in which a 5=1 0 is said to be proper; the
product and quotient of proper dual numbers are also proper.
If the dual product (2) is zero, there are three alternatives:
x=x'=0; y=y'=0; x=y=0.
The last shows that a dual product can vanish when neither factor
is zero; for any two pure dual numbers have zero as their product.
§ 31 MOTORS 65

If the function f(x) has the derivative f'(x), we define its value
for the dual argument X = x + Ex' by writing down its formal
Taylor expansion and setting E2 = E3 = = 0; thus
(5) Ax + Ex') = AX) + Ex'f' (x)
In particular,
(6) sin (x + Ex') = sin x + Ex' cos x,
(7) cos (x + ex') = cos x - ex' sin x.
When x = 0, we have sin Ex' = Ex', cos ex' = 1; consequently (6)
and (7) have the form of the usual addition theorems of the sine
and cosine. Note also that sine X + cost X = 1.
31. Motors. We now can characterize a motor as a dual multiple
of a unit line vector. Thus, on multiplying the unit line vector
A = a + Eao (a a = 1, a ao = 0) by the dual number X + EX',
we obtain the dual vector,
(1) M = m + Emo = (X + EX') (a + Ea.o).
On equating the real and dual parts of both members, we obtain
(2) m = Xa, mo = Xao + A'a.
To show that M is a motor we need only verify that mo transforms
in accordance with (29.3):
(3) MP =mo+POxm.
In view of (2), this result follows from

mp = Nap + A'a = X (ao + PO -a) + X'a


_ (Xao + X'a) + PO x (Xa).
From (3), m mp = m mo; hence the scalars m m,and m mo
are invariants in the sense that they are not altered by a shift of
origin. From (2),
(4)

m 0, we call the invariant,


X' mmo
(5)
mm
66 MOTOR ALGEBRA §31

the pitch of the motor. Choosing A > 0, we have the unique


solution,
A=ImI, A'=,zim1,
Aa=m, Aao=mo-µm.
When m 0 0, M has the dual length,
(6) A + EX, = I m { (1 + eµ)
and its axis is along the line vector,
(7) ImIA=m+e(mo-µm).

The equation of the axis is therefore


(mxmo)Xm
(8) rxm =mo -gym= ;

this shows that the axis passes through the point Q given by
--f mxmo
(9) r=OQ= ;

and, from (3),


(mxmo)Xm
(10) mQ=mo - =µm.

At all points on its axis M has the same moment µm, a fact also
apparent from (7). Only for points P on the axis is the moment mp
parallel to m.
Motors for which m 0 (A 0) are called proper. Proper
motors are screws if m mo 0 0 (A, A' 0 0) ; line vectors if m mo =
0 (A 0 0, A' = 0). Only proper motors have a definite axis.
If m = 0, mo 0 0 (A = 0, A' 0), M = emo is pure dual; then
mo is not altered by a change of origin. In this case M is called
a couple of moment mo. A couple may be regarded as a screw- of
infinite pitch with an axis of given direction but arbitrary position
in space.
Finally if m = 0, mo = 0, the motor M = 0, then A = A' = 0
from (2). Hence M = 0 only when its dual length A + eA' = 0.
Our classification of motors is therefore as follows:
Screw 0;1
} Proper
Line vector
Couple (m=0,
Zero (m=0, m0 =0):A=0,A' =0.
§ 32 MOTOR SUM 67

If two proper motors M, N are connected by a linear relation


with proper coefficients,
(a + ia')M + ($ + ia')N = 0, as 34 0,

either may be expressed as a dual multiple of the other; hence both


M and N are multiples of the same unit line vector and are there-
fore coaxial. Conversely, if M and N are coaxial, they satisfy a
linear relation with proper coefficients.
32. Motor Sum. The sum of the motors M = m + em0,
N = n + eno is defined as the motor,
(1) M + N = m + n + e(mo + no).
That M + N is a motor follows from (29.3) :

mp+nr = mo+no+POx(m+n)
THEOREM 1. The sum of two line vectors is a line vector only when
their axes are coplanar and their vectors have anon-zero sum.
Proof. If M and N are line vectors, M + N is a line vector
when and only when
m+n 0, and (m + n) (mo + no) = 0.
Since m mo = 0, n no = 0, the latter condition reduces to
(2)
which is precisely the condition (26.14) that the axes be coplanar.
THEOREM 2. If two line vectors intersect, their sum is a line vector
through the point of intersection.

Proof. If r1 is the position vector of the point Pl in which the


line vectors M and N intersect, we may take mo = rl x m, no =
r1 x n. Since the axis of M + N has the equation,
r- (m + n) = mo + no = r1 x (m + n),
it passes through the point P1.
THEOREM 3. If the line vectors M and N are parallel and n =
Xm (X F6 -1), the axis of their sum divides any segment from M to
N in the ratio X/1.
68 MOTOR ALGEBRA § 33

Proof. If Pl, P2 are points on the axes of M and N, we may


take mo = rl x m, no = r2 x n. The line vector,
M + N = (1+A)m+E(r,+Ar2)xm;
and the equation of its axis,
r x (1 + A)m = (r1 + Xr2) x m,
is satisfied by the point r = (r1 + Ar2)/(1 + X) which divides
PiP2 in the ratio A/1. The division is internal when m and n have
the same direction (X > 0), external when they have opposite di-
rections (A < 0).
THEOREM 4. The sum of two couples, if net zero, is another
couple.
Proof. If M and N are pure dual, M + N, if not zero, is also
pure dual.
THEOREM 5. If M and N are line vectors such that n = -m and
P1, P2 are points on their respective axes, M + N is a couple of
moment (r1 - r2) x m.
Proof. Since M = m + er1 x m, N = -m + ere x (-m),
M + N = e(rl - r2) x m.
33. Scalar Product. The scalar product of the motors M =
m + emo, N = n + eno is defined as the result of distributing the
product,. (m + emo) + (n + eno), namely,
(1)
This dual number is independent of the choice of origin; for on
computing mp, np from (29.3), we find that
(2)
The definition (1) shows that
(3)
The definition (1) for M M. N differs from that given by R. von Mises
("Motorrechnung, ein neues Hilfsmittel der Mechanik," Z. angew. Math.
Mech., vol. 4, 1924, p. 163) who defines M N as the invariant real scalar (2).
The definition (2) is suggested by applications in mechanics; its consequences
are (a) the familiar rules of vector algebra do not all carry over into motor
algebra, and (b) the elegant generalization of the scalar product of unit vectors,
given in (5), is lost.
§ 33 SCALAR PRODUCT 69

In order to compute the scalar product of two unit line vectors


A = a + ea0i B = b + - ebo, we shall suppose first that they are
not parallel and write
axb = esin <p,
where e is a unit vector. There is, then, a definite unit line vector
E in the direction of e which cuts both A and B at right angles.
If we choose the origin at the point of intersection of A and E,
and let denote the perpendicular distance between A and B,
taken positive if a X b points from A to B, we may write
A = a, exb, E = e.
Then we have

cos cp - ep sin cp.


If we now define the dual angle t between A and B as
(4) 'P + Ecp
a reference to (30.7) shows that
(5)

in complete analogy with a b = cos <p.


If A and B are parallel, let e be any unit vector perpendicular
to both A and B. If the origin is chosen on A, the foregoing com-
putation still applies; since sin <p = 0, we find A B = cos p = f 1.
If A and B are collinear, we choose the origin on their common
line; then A = a, B = b, and A B = a b = -4-1 as before.
Formula (5) covers these cases; in both cos 4) = cos gp = f 1 ac-
cording as a and b have the same or opposite directions.
We note that
=cos(p-ecp'sin(p=0,
when and only when
a
cos,p=0, rp'sin<p=0, or cp=2, =0;
that is, when the line vectors cut at right angles.
t The concept of dual angle is due to Study, Geometrie der Dynamen, Leipzig,
1903, p. 205.
70 MOTOR ALGEBRA § 34

THEOREM. In order that the axes of two proper motors cut at right
angles, it is necessary and sufficient that their scalar product vanish.
Proof. We express the motors in the form,
M = (a + ea )A, N = (0 + E#')B,
where A, B are unit line vectors along their axes and a0 - 0.
Then
M - N = [a$+
and, since the first factor is neither zero nor pure dual, M N = 0
implies A B = 0, and conversely.
We denote the dual part of M N by
(6)
As previously noted, this is von Mises' definition of the scalar
product. In common with M - N, the product M c N is commu-
tative and distributive with respect to addition. The preceding
calculation shows that
AoB = -
hence A o B = 0 implies p' = 0 or sin p = 0, that is, the line
vectors intersect or are parallel, and conversely. The same cri-
terion applies to line vectors aA, OB of arbitrary length. Conse-
quently a necessary and sufficient condition that two line vectors
1V ,JI be coplanar is that M e N = 0. This is the condition (26.14).
Writing the motor M = (a + ea')A gives
(a+ea')2=a2+2eaa'
as the square of its dual length; and M o M = 2aa'. Evidently
M o M = 0 implies that M is a line vector (a' = 0) or a couple
(a = 0).
34. Motor Product. The motor product of the motors M =
m + ano, N = n + eno is defined as the dual vector obtained by
distributing the product (m + emo) x (n + eno), namely,
(1) MxN = mxn + e(mxno +moxn).
When the origin is shifted to P, the dual part of M x N becomes

mxnp + mp x n = m x (no + POxn) + (mo + POxm) xn


= mxno + moxn + POx (mxn),
§ 34 MOTOR PRODUCT 71

in view of the identity,


PO x (m x n) + m x (n x PO) + n x (PO x m) = 0.
This transformation conforms with (29.3) and shows that M x N
is a motor.
The definition (1) shows that
(2) MxN = -NxM, Lx(M+N) =LxM+LxN.
In order to compute the motor product of two unit line vectors
A = a + eao, B = b + ebo, we shall suppose, first, that they are
not parallel. With the same notation and choice of origin as in
§ 33, we have
AxB =a xb + &p ax (exb)
= e(sin (P + ecp' cos (p)
or, in terms of the dual angle 4 = c + eqp ,
(3) AxB = Esin4
in complete analogy with a x b = e sin cp.
When A and B are parallel, let e be any unit vector perpendicular
to both A and B. If the origin is chosen on A, the foregoing com-
putation still applies, and
AxB = e
e + Eeo, an arbitrary line vector in the direction of e.
This conforms with the fact that A x B, being pure dual a couple),
has its axis determined in directionlbut not as tolposition in space.
When A and B are collinear, we choose an origin on their common
line;thenA=a,B bandAxB=axb=0andEisentirely
arbitrary. Formula (3) covers these cases in which sin 4 = f Ev'
and 0, respectively.
We note that
A x B = sin (D E = (sin p + erp cos cp) E = 0,
(or sin 4) = 0) when and only when
sin(p=0, gyp'=0,
that is, when the line vectors are collinear.
THEOREM. In order that two proper motors be coaxial, it is neces-
sary and sufficient that their motor product vanish.
72 MOTOR ALGEBRA § 35

Proof. With the notation of § 33,


M x N = [a/3 + e(af' + a'f )]A x B, (a/3 X 0).
Since the first factor is neither zero nor pure dual,,M x N = 0 im-
plies sin 4) = 0, and conversely.
Finally let us consider the motor product,
(4) M x N = [a$ + E(a3' + a'/3) ] [sin gp + cos go] E,

for any proper motors M, N. If their axes coincide, M x N = 0.


If their axes are parallel, sin <p = 0, the coefficient of E is pure
dual, and the axis of the couple M x N is any line parallel to the
common normal to the axes of M and N. If their axes are non-
0, and M x N is a proper motor whose axis (along
parallel, sin cp
E) is the common normal to the axes of M and N.
35. Dual Triple Product. The dual triple product of three
motors L = I + do, M = m + em0i N = n + eno is defined as
the dual number obtained by distributing the product L x M N;
thus
(1)

This definition shows that


(2)

LxM N = LxM N = 0,
in exact analogy with vector products.
With von Mises' definition of a scalar product, the preceding
triple product becomes
(5) LxMoN = (lxm) - no + (lxmo +loxm) - n;
this is the dual part of L x M N. Equations (2), (3), (4) also
apply to L x M o N.
We now propose to find under what conditions L x M N = C
when L, M, N are proper motors. If the axes of L, M, N are all
parallel, l x m= m x n= n x 1= 0 and L xM N= 0. If their
axes are not all parallel, choose the notation so that 1 x m 0.
Then L x M is a proper motor whose axis cuts the axes of L and M
at right angles. The condition L x M N = 0 then holds when
and only when the axis of N cuts the axis of L x M at right angles.
§ 36 MOTOR IDENTITIES 73

In this case the axes of L, M, N have a common normal (the axis


of L x M). We therefore may state the
THEOREM. The dual triple product of three proper motors will
vanish only when (a) their axes are all parallel, or (b) their axes have
a common normal.
If the proper motors L, M, N satisfy a linear relation with proper
coefficients,

(6) (a+ecx')L+($+E$')M+(y+ey')N=0, aIiy 0,


L x M N = 0, and we can apply the preceding theorem. More
precise information, however, is given by the equations,
MxN NxL LxM
a+ea'=a+e0' = +e1''
which follow from (6). These show that the three motors M x N,
N x L, L x M are all proper or all couples. When proper these
motors are all coaxial; when couples, their moments are all
parallel. Hence from (6) we conclude that either (a) the axes
of L, M, N, no two of which are parallel, have a common normal;
or (b) their axes are all parallel and coplanar. Thus the linear
relation (6) implies that the axes of L, M, N have a common normal.
Conversely, if the axes of L, M, N have a common normal, we can
infer a linear relation (6), provided the case in which two axes are
parallel to each other but not to the third is excluded. We omit
the proof.
36. Motor Identities. All the identities of vector algebra are
still valid when the vectors are replaced by motors. For the motor
products are defined as the results of applying the distributive law
to dual vectors and subsequently equating e2, e3, . . . to zero.
Now the identities apply to the ordinary vectors forming the dual
vectors, so that the results before equating e2, e3, to zero are
true for any value of the scalar e. In particular, the identities
hold when e2, e3, are replaced by zero. Thus we have the
motor identities:
(1) Lx(MxN)
(2) Lx (M N) + Mx (NHL) +Nx (L M) = 0,
(3) (L M) (NxP) =LNMP -LPMN.
74 MOTOR ALGEBRA § 37

The identity (2) has an interesting geometric interpretation


when all nine motors,
L,M,N;MxN,NxL,LXM;Lx(MxN),Mx(NxL),Nx(LxM),
are proper. In this case, the axes of the three last motors have a
common normal (§ 35). This common normal and the axes of the
preceding nine motors form a configuration of ten lines, each one
of which is normal to three others. This is essentially the theorem
of Peterson and Morley. $ When L, M, N are line vectors along
the sides of a plane triangle, this theorem implies the concurrence
of its altitudes. The theorem of Peterson and Morley therefore
may be regarded as a generalization of this result.
37. Reciprocal Sets of Motors. Consider two sets of proper
motors M1i M2, M3; M1, M2, M3 for which ml m2 x m3 0,
m1 m2 x m3 /- 0. The sets are said to be reciprocal when the nine
equations,
(1) Mi Mi = S (i, j = 1, 2, 3),
are satisfied. The indices are merely labels, and S is the Kro-
necker delta.
Let the set Mi be given; we then may compute the reciprocal
set M' uniquely. Since M2 M1 = 0, M3 M1 = 0, the axis of
M1 is the common normal to the axes of M2 and M3; hence
M1 = (X + eX')M2 x M3i and, since MI-M' = 1, A + eX' =
1/M1 M2 X M3; the condition ml m2 x m3 0 0 ensures that
M1 M2 X M3 is not pure dual. Thus we find
MX Mk M'xMk
(2) M2 = M1 M2 x M3 , Mi = M2 x M3
M1
where ijk represent any cyclical permutation of 123. That the
motors (2) satisfy equations (1) is shown by direct substitution.
Equations (2) have the same form as the corresponding equations
in vector algebra. Moreover, in
(M2xM3) . (M2xM3)
M1,M1 =
(M1 M2 x M3) (M1 M2 x M3)
the left member is 1, and, from (30.3) the numerator on the right
is 1; hence
(3) (M1 M2 x M3)
(Ml M2 x M3) = 1.
$ E. A. Weiss, Einfuhrung in die Linien-geometrie and Kinematik, Leipzig
1935, p. 85.
§ 38 STATICS 75

The axes of two reciprocal sets taken in the order


M1M3M2M'M3M2 form a skew hexagon with six right angles.
Since
(4) M1 x M1 + M2 x M2
+ M3 x M3 = 0,
the common normals of its opposite sides themselves admit a com-
mon normal. The six sides, the three common normals of the
opposite sides, and their common normal, thus form a Peterson-
Morley configuration (§ 36).
If we put Mi = Mi in (1), we find that the only self-reciprocal
motor sets are formed by three mutually orthogonal unit line vec-
tors through a point.
Any motor M may be expressed linearly with dual coefficients
in terms of M1, M2, M3, provided m1 m2 x m3 X 0. To show
that the representation is possible, we note that
(5) M = (a + ea')M1 + ($ + ea')M2 + (y + ey')M3
is equivalent to the two vector equations,
(6) m = amt + /3m2 + 'Ym3,
(7) mo = amlo + /3m20 + 'Ym30 + a'm1 + /3'm2 + 7'm3.
Since m1 m2 x m3 ; 0, the vectors ml, m2, m3 have a reciprocal
set m1, m2, m3. From (6),
a = m- m1, /3 = m-m2, y = m-m3;
with a, 0, y known, we may determine a', 0', y' in the same way
from (7). The actual computation, however, can be effected most
simply from (5) by using the reciprocal set m1, m2, m3. Thus we
find
(8) M = M M1M1 +M M2M2 +M M3M3
38. Statics. The statics of rigid bodies may be developed in-
dependently of dynamics from four fundamental principles: t
Principle I (Vector Addition of Forces). Two forces acting on
the same particle may be replaced by a single force, acting on the
particle, equal to their vector sum.
Principle II (Transmissibility of a Force). A force acting on a
rigid body may be shifted along its line of action so as to act on
any particle of that line.
f Sec Brand, L., Vectorial Mechanics, New York, 1930, Chapter II.
76 MOTOR ALGEBRA § 38

Principle III (Static Equilibrium). If the forces acting on a


rigid body, initially at rest, can be reduced to zero by means of
principles I and II, the body will remain at rest.
Principle IV (Action and Reaction). This need not concern us
here.
Principle II states in effect that a force acting on a rigid body is
a line vector. Such a force is determined by the vector f giving
its magnitude and direction and by its moment fo about 0 (§ 27).
Thus the dual vector,
(1) F=f+efo
represents a force f acting along the line whose equation is r x f = fo.
If two forces P, Q act on a particle with position vector r,
P=p+Erxp, Q=q+E1xq;
principle I asserts that both forces may be replaced by a single
force, their resultant, whose vector and moment are p + q,
r x (p + q) ; that is, the intersecting line vectors P and Q are
equivalent to the line vector P + Q, their motor sum (§ 32).
Two parallel forces P, Q also have a resultant P + Q provided
p, + q 0. To see this, we need only introduce a pair of opposed
forces F, -F (equivalent to zero) acting along any line cutting the
lines of P and Q. We may now apply principle I to find the re-
sultants of P + F and Q - F. These coplanar forces intersect,
since
(p+f)x(q-f) =fx(p+q) 96 0,
and therefore have the resultant,
(P + F) + (Q - F) = P + Q.
If Q = XP (X 96 -1), the line of P + Q divides all segments from
P to Q in the ratio X/1 (§ 32, theorem 3).
If p + q = 0, po + qo 0, the parallel forces P, Q are equal
in magnitude and opposite in direction and are said to form a
couple of moment po + qo.
In any case, when a system of forces Fa acting on a rigid body
is changed into an equivalent system G; by the application of
principles I and II, the force-sum and moment-sum remain un-
altered :
21,; = 2;g1, lfto = 2;gio.
§ 38 STATICS 77

For each application of principle I leaves these sums unaltered;


and shifting a force in accordance with principle II does not alter
its vector or moment.
Now it can be shown that any system of forces Fi acting on a
rigid body can be reduced by means of principles I and II to two
forces, * say P= p + epo, Q = q + eqo. Hence, if EF1 = m +
emo, we have
p+q=m, po+qo=mo.
If m 0, m mo 0, the system Fi is equivalent to a screw
M = m + emo (§ 31) which may be expressed in many ways as
two non-coplanar forces. A screw cannot be reduced to a single
force. The screw M may be regarded as a force m acting through
the origin and a couple of moment mo.
If m 5,4- 0, m mo = 0, the forces P and Q are coplanar; for
m mo = 0 implies p qo + q po- = 0 (26.14). The forces P, Q
now may be reduced to the single force M = m + ano, their
resultant.
If m = 0, mo 0, the system F1 is equivalent to the couple
P, Q of moment mo.
Finally, if m = 0, mo = 0, the system F, reduces to zero, and,
in accordance with principle III, the rigid body is in equilibrium.
In brief, the system of forces Ft acting on a rigid body is equivalent
to the motor M = EFz j this may be a screw, a single force, a couple,
or zero; only in the last case is the rigid body in equilibrium.
The moment of a force F = f + do about an axis along the unit
line vector A = a + eao was defined (27.4) as a fp, P being a
point on the axis of A. The moment of a system of forces, equiva-
lent to the motor M = m + emo, about this axis is therefore

Now
a

and, since OP * a = ao,


(2)
We call this the moment of the motor M about the axis A. This
moment is independent of the choice of P on the axis.
* Brand, L., Vectorial Mechanics, p. 143.
78 MOTOR ALGEBRA § 39

39. Null System. A line is called a null line with respect to a


motor M if the moment of M about this line is zero. In order that
a line, given by the unit line vector A = a + eao, be a null line
with respect to M, it is necessary and sufficient that
(1)

If L = XA, L o M = XA o M; hence the axis of any line vector L


will be a null line when and only when L o M = 0. The totality
of such lines constitutes the null system of the motor M.
A motor M can be replaced in many ways by two line vectors
F, G. When M is a screw (m mo F6 0), we shall prove that the
line of one may be chosen at pleasure, provided it is not a null
line or parallel to the axis of M; after this choice F and G are
uniquely determined.
Write F = aA, G = oB as multiples of unit line vectors, and
choose A as any line not in the null system. Now M = F + G is
equivalent to
m = as + ob, mo = aao + obo
Since a ao = b bo = 0,

(2) a =

thus F = aA is known, and G = M - F. Since A is not a null


line, a mo + m ao 7 0; and, since A is not parallel to m,
o 7 0, and G is actually a line vector.
All null lines through a point P are perpendicular to mp and
therefore lie in a plane p, the null plane of P. Moreover all the
null lines on any plane p pass through a point P, the null point
of p, where mp is perpendicular to p.
Analytic Proof. Consider a null line through the points (a, ao),
(o, bo); the coordinates are (abo - aoo, ao x bo) from (26.9). Since
it is a null line,
(3) m ao X bo + mo (abo - aoo) = 0.
This may be written
(4) (amo - ao x m) bo = ao moo,
§ 39 NULL SYSTEM 79

hence, if the point (a, ao) is fixed and ($, bo) varies over all null
lines through (a, ao), (,l, bo) always will lie in the plane whose co-
ordinates are (amo - ao x m, ao mo).
Next, let the null line be common to the planes (a, ao), (b, Rio);
its coordinates are (a x b, a00 - aob) from (26.10). Since it is a
null line,
(5) m (a(3o - aob) + mo a x b = 0.
This may be written
(6) (mao - mo x a) b = m a $o;
hence, if the plane (a, ao) is fixed and (b,,30) turns about all null
lines in (a, ao), (b, Qo) always will pass through the point whose
coordinates are (m a, mao - mo x a).
Equations (3) and (5) are not altered by an interchange of the
two points or two planes; hence the same is true of (4) and (6).
Therefore:
If the null plane of point A passes through B, the null plane of B
passes through A.
If the null point of plane a lies on b, the null point of b lies on a.
From the preceding results:
(7) Point (a, ao) - null plane (amo - ao x m, ao mo) ;
(8) Plane (b, Ro) - null point (m b, mao - mo x b).
Indeed, if we solve
b=amo - aoxm,
for (a, ao), we obtain the null point in (8). Note also that ao b
= a0o; let the student interpret this relation in the light of (26.16).
When M is a line vector, the null lines are lines that intersect
its axis. Then (7) gives the plane through the point and M (26.11);
and (8) 'the point on the plane and M (26.12).
If M is a screw (m mo F!5 0), not only are points associated
with planes, but also lines with lines in general. For if A is any
unit line vector not in the null system of M or parallel to its axis,
we can determine uniquely a unit line vector B, its conjugate, such
that
(9) M = aA + (3B,
80 MOTOR ALGEBRA § 40

where a = m mo/A o M from (2). Any line vector C that cuts two
conjugate lines is a null line; for
MoC = aAoC+$B°C = 0.
In view of this property we may characterize conjugate lines as
follows: The conjugate of a line A is the line B: (i) which is common
to the null planes of all points on A, or (ii), which contains the null
points of all planes through A. With this point of view, null lines
are self-conjugate. Along a line parallel to the axis of M, the
moment of M is constant; since the null planes of its points are
all parallel, we may say that its conjugate is at infinity.
40. Summary: Motor Algebra. The number X = x + ex' is
called dual if x and x' are real and a is a unit with the property
E2 = 0. If x 0, the dual number is proper. Operations with
dual numbers are carried out as in ordinary algebra and then
e2, E3 , . set equal to zero. Division, however, is only defined for
.

proper dual numbers. The function f(x + ex') is defined by means


of. its formal Taylor expansion; only the first two terms, f (x) +
Ex' f'(x), appear since E2 = e3 = . . 0.
A line vector with the Pliicker coordinates a, ao (a ao = 0) is
written as a dual vector, A = a + eao; for a unit line vector
I a I = 1. When the origin is shifted to P, ao becomes

ap = ao + PO- a.
A motor M = m + emo is a dual multiple of a unit line vector,
M = (X + EA')A = (X + EA') (a + eao)
now m mo = 0 only when A' = 0. A shift of origin to P changes
the moment vector as before:
mp=mo+P0xm,
but m mp = m mo is invariant.
Motors for which X F 0 are said to be proper. Proper motors
are screwf when X' 0 0, line vectors when A' = 0. Proper motors
have a definite axis given by the unit line vector A. When X = 0,
A' 0 0 we have a pure dual motor emo; it is called a couple of
moment mo. The moment of a couple is a free vector, determined
in direction but not in position.
When X = A' = 0 (X + EX' = 0), the motor reduces to zero.
§ 40 SUMMARY: MOTOR ALGEBRA S1

The sum of the motors M = m + em0, N = n + en0 is defined


as
M + N = m + n + e(mo + no).
If M and N are line vectors, M + N is a line vector only when
m no+n mo=0.
The scalar and motor products, M N and M x N are obtained
by distributing the products as in ordinary vector analysis and
then setting e2 = 0:
M N = m - n+
MxN = mxn+ e(mxno+moxn).
If A and B are line vectors making an angle p and at a normal
distance apart (reckoned positive if a x b points from A to B),
the dual angle between A and B is defined as F + &p'. For
unit line vectors,
A B= cos A x B= sin -tE
where E is a unit line vector in the direction a x b and cutting
A, B at right angles. These results are perfect analogues of
If M and N are non-parallel proper motors, the common normal
to their axes is the axis of M x N. When M and N are parallel,
M x N is a couple whose moment is perpendicular to their axes.
If M and N are proper motors, M N = 0 implies that their
axes cut at right angles; M x N = 0 that their axes coincide; and
conversely.
For three motors, the dual number L x M N is obtained by dis-
tributing the product (1 + el0) x (m + emo) (n + eno) :

If L, M, N are proper, L x M N = 0 implies that their axes are


all parallel or have a common normal, and conversely.
Two sets of three motors Mti, M' whose triple products are
proper are said to be reciprocal when M; M' = S (i, j = 1, 2, 3).
All formulas for reciprocal vectors have exact analogues for recip-
rocal motors.
The forces acting on a rigid body are line vectors. A system F;
of such forces is equivalent to their motor sum M = SFj, which
82 MOTOR ALGEBRA

may be a screw, a single-force, a couple, or zero; only when M = 0


is the body in equilibrium.
All the rules of calculation in "real" vector algebra have exact
analogues in motor algebra.

PROBLEMS
1. If A is a unit line vector, show that the screw M = (A + ea')A may be
expressed as the sum of a line vector and a couple whose moment is parallel
to the line vector, namely,
M = aA + eµm,
where A= X'/X is the pitch of the screw.
2. Prove that the axes of the motors M, N and M + N either are parallel
or have a common normal.
3. Express the screw M = 2i + e(i + j + k)
(a) As the dual multiple of a unit line vector.
(b) As the sum of a line vector and a couple whose moment is parallel to
the line vector.
(c) Find the equation of the axis of M.
-- ---,
4. In the tetrahedron OABC, el = OA, e2 = OB, e3 = OC. If the forces
Pet, Qe2, Re3; P'(e3 - e2), Q'(e1 - e3), R'(e2 - el)
acting along its six edges are equivalent to the motor m + emo, show that
they are uniquely determined by the six equations:

P1 = mo e1 ... P + Q' - R' = m , e1,


[eie2e31

5. Show that the n forces P1Q1, P2Q2, , P. Q are equivalent to the


motor,
n(q* -p*) +expi gi,
where p*, q* give the mean centers of the points pi and q;, respectively. [Cf.
(9.4)].
6. In order that the axes of the motors, M = m + emo, N = a + eao, be_
coplanar, it is necessary and sufficient that

where µ and v denote the pitch of M and N.


7. If three line vectors, a + eao, b + ebo, c + eco, are parallel to a plane,
as +,6b + -yc = 0 (§ 5). Prove that any line vector d + edo meeting them
is parallel to a fixed plane normal to aao + Rbo + yco.
8. The forces Xi, Yj, -Zk act through the points (0, 0, 0), (a, b, c), (0, b, 0),
respectively. Show that they reduce to a single force if
X/a + Y/b + Z/c = 0.
PROBLEMS 83

9. If ABCD are the vertices of a skew quadrilateral and PQRS are the mid-
points of its sides taken in order, show that the motor equivalent to the forces
AB, BC, CD, DA is also equivalent to four times the couple formed by PQ
and RS.
10. If the moment of the motor M about each of the six edges of a tetra-
hedron is zero, show that M = 0. [Cf. (38.2)].
11. If the origin 0 is taken on the axis of a screw M of pitch µ show that
the equation of the null plane of the point Ti is

If the axis of M is chosen as z-axis, this equation becomes xyl - yxl =


A(z - zi).
CHAPTER III
VECTOR FUNCTIONS OF ONE VARIABLE

41. Derivative of a Vector. Let u(t) denote a vector function


of a scalar variable t over the interval a < t < b; that is, when t
is given, u(t) is uniquely determined.* The function u(t) is said
to be continuous for the value t if u(t + At) -+ u(t) as At -* 0.
To obtain a clear idea of the way in which u varies with t, we
may regard u(t) = OP as a position vector issuing from the origin.
Then as. t varies, the point P traces a certain curve r in space.
For the values t and t + At, let u(t) = OP, u(t + At) = OQ; the
change in u for the increment At is
--4 -* --4
Du=u(t+At) -u(t) =OQ-OP=PQ.
The average change per unit of t is Au/At. When At > 0, Au/At
is a vector (PA' in Fig. 41) in
'R R * the direction of Au and 1/At times
as long; but if At < 0, Au/At has
the direction of -Au (then PA'
in Fig. 41 must be reversed). If
PA' approaches a limiting vector
PA as At -+ 0, we call PA the
derivative of u(t) with respect to t
and denote it by du/dt or u'(t).
The equation defining the de-
Fia. 41 rivative is therefore
du = lim u(t + At) - u(t) Du
(1) = lim
dt w -o At 'U - o At
* The word function implies a single-valued function.
84
§ 41 DERIVATIVE OF A VECTOR 85

If the limit du/dt exists for the value t, u(t) is said to be differen-
tiable at t. Then we may write Au/.t = du/dt + h where h --> 0
as At -* 0; hence
At (du
0u = -dt + h) --). 0, as At --> 0,

and u(t + At) ---> u(t). Thus if u(t) is differentiable at t, it is also


continuous there.
If u(t) is differentiable at P, Q describes the arc QP of the curve
r as At -* 0, and the limiting direction of the chord PQ, and hence
of the vector PA', is along the tangent at P. The limiting vector
PA = du/dt is therefore tangent to I' at P. Since Du/At has the
same direction as Au when At > 0, the opposite when At < 0, du/dt
is a vector tangent t o r in the direction of increasing t. We re-
0
state this important result as follows:
If the vector u(t) = OP varies with t, so that P describes the curve
r when 0 is held fast, the derivative du/dt, for any value of t, is a
vector tangent to r at P in the direction of increasing t.
Example 1. If u = OP is a variable vector of constant direction, P will
move on a straight line when 0 is held fast; hence du/dt, being tangent to this
line, will be parallel to u.
Example 2. If u = OP is a variable vector of constant length, P will describe
a curve r on the surface of a sphere when 0 is held fast; hence du/dt, being
tangent to r at P, will be perpendicular to the radius OP of the sphere. In
brief : If j u I is constant, du/dt is perpendicular to u.
If u is a constant vector, that is, constant in both length and
direction,
Du = 0, Ot = 0, and dt = 0.

The derivative of a constant vector is zero.


When u is a function of a scalar variable s, and s in turn a
function of t, a change of At in t will produce a change As in s
and therefore a change Au in u. On passing to the limit At --* 0
in the identity,
Du Du As du du ds
At
-,
As At
we get -
dt
= ---
ds dt
the familiar "chain" rule.
86 VECTOR FUNCTIONS OF ONE VARIABLE § 42

The higher derivatives of u(t) are defined as in the calculus:

u"(t) = du' , U"'(t) = du" , etc.


dt dt

42. Derivatives of Sums and Products. Let u(t) and v(t) be


two differentiable vector functions of a scalar t. When t changes
by an amount At, let Au, Av, and O(u + v) denote the vectorial
changes in u, v, and u + v. Then
u + v + O(u+v) = u+ Du+v+ Ov,
O(u + v) = Du + AV,
A(u + v) Du AV

At At At

and passing to the limit At - 0 gives


d du dv
v)
(1) dt (u + dt + dt

Consequently, the derivative of the sum of two vectors is equal to the


sum of their, derivatives. This result may be generalized to the sum
of any number of vectors.
Consider next the product f(t)u(t) of a differentiable scalar and
a vector function. When t changes by an amount At, let Af, Au,
and A(fu) denote the increments of f, u, and fu, respectively.
Then, since multiplication of vectors by scalars is distributive (§ 4),
fu + o(fu) = (f + Af)(u + Au) = fu + f Du + Af U + Of Au,
O(fu)=fAu+Afu+OfAu,
AU-u+Of
A(otu)
=
f -;
passing to the limit At - 0, and noting that Af -* 0, we have

(2) dt (f u) = f dt + dt u.

This is formally the same as the rule for differentiating a product


of scalar functions.
§ 42 DERIVATIVES OF SUMS AND PRODUCTS 87

Important special cases of (2) arise when either f or u is constant:


d du d df
-(cu) =cdt,
-(fc) =atc.
If the components of a vector u(t) are ui(t) when referred to a
constant basis ei,
du dug
(3) u(t) = 2;ui(t) e,, ei.
dt dt
The components of the derivative of a vector are the derivatives of its
components.
If we pass now to the products u v and u x v, where u and v
are vector functions of t, the same type of argument used in prov-
ing (2) shows that
d dv du
(4) -(u.v)=u-at+at-V,
d dv du
(5) (u xv - U x - - x v.
dt dt + dt
The proofs depend essentially upon the distributive laws for the
dots and cross product. In (5) the order of the factors must be
preserved.
If 1 is an axis with unit vector e, the orthogonal component of u
on 1 is e u (15.6) ; hence
d du du
comp1 u = e comps
dt -
dt = dt
THEOREM 1. A necessary and sufficient condition that a proper
vector u be of constant length is that
du
(6) u =0.
dt
Proof. Since u 12 = u u, we have, from (4),
I

d du
u12=2u'dt
-1
If I u I is constant, the condition follows: conversely, the condition
implies that I u I is constant.
88 VECTOR FUNCTIONS OF ONE VARIABLE § 43

THEOREM 2. A necessary and sufficient condition that a proper


vector always remain parallel to a fixed line is that
du
(7) uX = 0.
dt
Proof. Let u = u(t)e where e is a unit vector; then

uX -
du
dt
= ue x - +e u -de\J = u2e X de
Cdu
dt dt
-- dt

If e is constant, de/dt = 0, and the condition follows. Conversely,


since u 0 0, the condition implies that
de de
ex-=0.
dt
Also
dt

from theorem 1. These equations are contradictory unless de/dt


= 0; that is, e is constant.
43. Space Curves. Consider a space curve whose parametric
equations are
(1) x = x(t), y = y(t), z = z(t),
where x(t), y(t), z(t) are analytic functions of the real variable t
defined in a certain interval T of t values. To avoid having the
curve degenerate into a point, we explicitly exclude the case in
which all three functions are constants. We also restrict the in-
terval T so that there is just one value of t corresponding to each
point P of the curve. Then equations (1) set up a one-to-one
correspondence between the points of the curve and the values of
t in the interval T. The requirement that the functions x(t),
y(t), z(t) be analytic in T ensures the continuity of the functions
and their derivatives of all orders and also guarantees a Taylor
expansion for each function about any point of T. A curve which
admits a representation (1) with functions thus restricted is called
an analytic space curve. Moreover, if all three derivatives dx/dt,
dy/dt, dz/dt do not vanish simultaneously for any value of t in the
interval T, the curve is said to be regular, and the parameter t
is said to be a regular parameter.
bet us make a change of parameter,
t = P(u),
where jp(u) is an analytic function of u in a certain interval U; we
§ 43 SPACE CURVES 89

assume also that t just covers T as u ranges over U. Then the


equations,
x = x[,P(u)l,
y = y[O(u)l, z = z[co(u)]
constitute a new parametric representation of the curve. Writing
dt/du = <p'(n), we have
dx dx dy dy dz dz
lp
du = at (u)' du = at (u)' du =
a regular parameter, u is a regular parameter when and
only when (u) 0 in U. When this condition is fulfilled, the
implicit-function theorem ensures the existence of the inverse func-
tion u = 4'(t) which is also single valued and analytic in T. Thus
the points of the curve not only correspond one to one to the t
values in T, but also to the u values in U:
P -* t -> fi(t) = U, u --> (P(u) = t --+ P.
Let equations (1) define a regular analytic space curve. The
position vector of the point P(t) is
(2) r = i x(t) + j y(t) + k z(t) = r(t);
and, if we write t, z for dr/dt, dx/dt,
(3) t = it(t) + j y(t) + k i(t) F6 0.
If Po(t = to) is a fixed point on the curve, the length of the arc s
from Po to P is defined as
(4) s= fV
to
dt,

an analytic function of t which is positive or negative, according


as t > to or t < to; moreover
(5) ds/dt = VT-t > 0,
The inverse function t = <p(s) is single valued and analytic, gnd
dt/ds > 0. Putting t = <p(s) in (2), we obtain r = r(s), in which
the arc s is a regular parameter.
The integrand in (4) is unity only when t = 1; hence we have
the
THEOREM. For the curve r = r(t) the parameter t is the length of
arc measured from a fixed point when and only when dr/dt is a unit
vector.
90 VECTOR FUNCTIONS OF ONE VARIABLE § 44

44. Unit Tangent Vector. Let r = r(t) be a regular analytic


space curve on which s = arc POP is reckoned positive in the
direction of increasing t. Then (Fig. 44a)
dr Ar PQ
-- = lim - = lim
ds A -o As Q -. P arc PQ
is a unit vector (§ 43), and hence
chord PQ
Em
Q -. P arc PQ
The tangent line to the curve at P is defined as the limiting posi-
tion of the secant PQ as Q approaches P. Hence from its definition

FIG. 44a FIG. 44b

we conclude that dr/ds is a unit vector T tangent to the curve at


P and pointing in the direction of increasing arcs;
dr
(1) = T.
ds

An important special case arises when r is a unit vector revolving


in a plane. If we imagine this vector R always drawn from the
same initial point 0, its end point will describe a circle of unit
radius (Pig. 44b) and s = 0, where 0 denotes the angle in radians
between a fixed line OA and R. If P is a unit vector perpendicular
to R in the direction of increasing angles, (1) now becomes
dR/dO = P. If k is a unit vector normal to the plane and pointing
in the direction a right-handed screw advances when revolved in
the positive sense of 0, P = k x R, and
dR
(2) = k x R = P.
dB
§ 44 UNIT TANGENT VECTOR 91

In Fig. 44b, k points upward from the paper. Moreover, since


dP/d6 = k x (dR/d6), we have
dP
(3) -=kxP=
d6
-R.
We state these results in the
THEOREM. The derivative of a unit vector revolving in a plane,
with respect to the angle that it makes with a fixed direction, is another
unit vector perpendicular to the first in the direction of increasing
angles.
If P(x, y) is a variable point on a plane curve, r = xi + yj,
dr dx dy
T= + ds ,
as = as'
and i T = dx/ds, j T = dy/ds; hence

(4) ds = cos (i, T), ds = sin 01 T).

If (r, 0) are the polar coordinates of P, we write r = rR, where


the polar distance r and the unit vector R are functions of 0. Now
dr dR d9 dr d6
T = -- R+r --- = - R+r--P,
ds d6 ds ds ds
and R T = dr/ds, P T = r d6/ds; hence
dr do
cos (R, T), r = sin (R, T).
(5)
-= --
Example 1. If 0 is measured counterclockwise from the x axis,
R=icos0+jsin0, P=icos(0+ 117r)
From (2), the corresponding components of dR/d9 and P must be equal; hence

cos 0 = cos (0 + 4r) sin B, d sin B = sin (0 + fir) = cos 0.


dB

Example 2. If rl, r2 are the distances of a point P on an ellipse from the foci,
rl + r2 = const. On differentiating this equation with respect to a, we have,
from (5),
drl dr2
ds
+ ds
= (RI + R2) T = 0.
92 VECTOR FUNCTIONS OF ONE VARIABLE § 45

Since RI + R2 is perpendicular to T, the normal to the ellipse at P has the


direction of RI + R2. The normal therefore bisects the angle between the focal
radii.
45. Frenet's Formulas. Let r = r(s) be a space curve, s =
are POP, and T = dr/ds the unit tangent vector at P. Since T is
of constant length, dT/ds, if not zero, must be perpendicular to T.
A directed line through P in the direction of dT/ds is called the
principal normal of the curve at P. Let N denote a unit vector
in the direction of the principal normal; then we may write
(1) dT
= KN,
ds
where K is a non-negative scalar called the curvature of the curve
at P.
Now B = T x N is a third unit vector perpendicular to both T
and N. Thus at each point of the curve we have a right-handed
set of orthogonal unit vectors, T, N, B, such that
TxN = B, NxB = T, BxT = N.
As P traverses the curve, we speak of the moving trihedral TNB.
A directed line through P in the direction of B is called the bi-
normal to the curve at P.
Since B is a unit vector, dB/ds, if not zero, must be perpendicular
to B. Differentiating B T x N, we have
dB dT dN dN
ds dsxN+Txds = Txds
in view of (1). Hence dB/ds is perpendicular to T as well as B
and therefore must be parallel to
-B x T = N. We therefore may write

dB
__ (2) = -TN,
7-7- ds

I/ where T is a scalar called the torsion


of the curve at P. The minus sign
Torsion positive is introduced in (2) so that, when 7-
Fla. 45 is positive, dB/ds has the direction
of - N; then, as P moves along the
curve in the positive direction, B revolves about T in the same sense
as a right-handed screw advancing in the direction of T (Fig. 45).
§ 45 FRENET'S FORMULAS 93

We may now compute dN/ds from N = by use of (1) and


(2) ; thus
dN dB dT
(3) -xT+Bx-= -rNxT+KBxN.
ds ds ds
Collecting (1), (2), and (3), we have the set of equations:
dT
ds

dN
(4)
ds

dB
ds

known as Frenet's Formulas, which are fundamental in the theory


of space curves.
If we write (3) in the form,
dN
= (rT + KB) x N,
ds
and introduce the Darboux vector,
(5) S = rT + KB,
we have
SxT = KN, SxB = -TN.
Hence Frenet's Formulas may be put in the symmetric form:
dT dN dB
(6) = S x T, - =SxN SxB.
ds ds ds
Since S N = 0, we may write the Darboux vector in the form:
dN
(7) S = Nx(SxN) = .

Nxd8
From (5), we see that the curvature K and torsion r are the
components of the Darboux vector on the binormal and tangent,
respectively. From (4), it is clear that both K and r have the
dimensions of the reciprocal of length; hence
p=1/K, a=1/r
94 VECTOR FUNCTIONS OF ONE VARIABLE § 45

have the dimensions of length and are called the radius of curva-
ture and the radius of torsion, respectively.
Since N, by definition, has the same direction as dT/ds, the
curvature K is never negative. If K vanishes identically,
dT dr
ds =0,
T=ds=a, r=as+b,
where a and b are vector constants. The curve is then a straight
line. Conversely, for a straight line, T is constant and K = 0.
The only curves of zero curvature are straight lines. For a straight
line the preceding definition fails to determine N. We therefore
agree to give N any fixed direction normal to T, and as before
define B = T x N. Since B is also constant, dB/ds = 0 and T = 0.
For a straight line the Darboux vector is zero.
The torsion may be positive or negative. As P traverses the
curve in the positive direction, the trihedral TNB will revolve about
T as a right-handed or left-handed screw, according as r is posi-
tive or negative. The sign of r is independent of the choice of
positive direction along the curve; for, if we reverse the positive
direction, we must replace
dT dN ' dB dT dN dB
s, N, By
by
-8) -T, N,
,
- B,
T, ds , ds , ds ds , ds ds '
and equations (4) maintain their form with unaltered K and r.
If r vanishes identically, dB/ds = 0, and B is a constant vector;
hence, from
dr
BT=B 0, B (r - ro) = 0,
ds

that is, the curve lies in a plane normal to B. Conversely, for a


plane curve, T and N always lie in a fixed plane, while B is a unit
vector normal to that plane; hence dB/ds = 0 at all points where
N and B are defined (K* 0) and -r = 0. The only curves of zero
torsion are plane.
Example. Parallel Curves. Two curves r and rl are called parallel if a
plane normal to one at any point is also normal to the other. The common
normal plane cuts r and rl in corresponding points P and PI; then PP, = rl - r
lies in the plane of N and B and may be written XN + µB, where X, µ are
scalars. Thus we have
(i) rl=r+XN+pB;
§ 46 CURVATURE AND TORSION 95

on differentiating (i) with respect to s and using Frenet's Formulas, we have


ds1 _
(ii) Ti- dX
=T+dN +X(-KT +rB)+dsduB -uTN
ds
_ (1 - AK)T + (A' - µr)N + (A' + Xr)B.
Since T1 and T are normal to the same plane, we may choose the positive sense
on t1 so that T1 = T; hence, from (ii), we conclude that

^ = 1 - AK, ds = Jlr, = -Ar.


ds
Moreover
dX du 1 d
2ds()2+u2) =0,

and A2 + µ2 is constant. Consequently, the distance PP1 between corresponding


points of parallel curves is always the same.
On differentiating T1 = T with respect to s, we have also

(iv) KjN1 1 = KN; hence ds =


d3 Kl ,

and N1 = N, B1 = B. Comparison with (iii) gives


K/KJ = 1 - AK or pl = p - A.
Finally, on differentiating B1 = B with respect to s, we have
dsi r
(v) -r1N1 = -rN; hence, if r 0 0, - =
7-1

From (iv) and (v), K/r = K1/7-1 at corresponding points.


46. Curvature and Torsion. By use of Frenet's Formulas, the
curvature and torsion are readily computed from the parametric
equations of the curve. Thus, on differentiating r = r(t) three
times and denoting t derivatives by dots, we have
_drds
$T,
ds dt =

f=ST+ dT- SZ

ds

= ST + S2KN,
dT dN
+(2saK+a2K)N+S3K
ds CAS

= ST + SSKN + (2SSK + S2K)N + S3K(-KT + TB)


= (S - S3K2)T + (3MK + S2K)N + S3KTB.
96 VECTOR FUNCTIONS OF ONE VARIABLE § 46

Hence
t x y= S3KB, t x It y= S6K2T,
and, since t = IsI 0,
xr
(1), (2) K= I T=
It t I 3 Itt x I2

If the positive direction on the curve is that of increasing t,


ds/dt > 0; and the preceding equations show that
T has the direction of t,
B has the direction of t x i',
and, since N = B x T,
N has the direction of (f x Y) x t.
If the parametric equations of a plane curve are x = x(t),
y = y(t), we have
r = xi + yj, t= + yJ, r = xi + yl,
and, from (1), its curvature is

(3) K = (x2 -
I xy + yx I
y2)

If the curve has the Cartesian equation y = y(x), we can regard


x as parameter: x = t, y = y(t). Then (3) becomes
K =
(4)
(1 + y,2)I'
where the primes indicate differentiation with respect to x.
If the curve has the polar equation r = r(9), we may write
r = rR, where R is a unit radial vector. Then regarding 0 as the
parameter t, we have, from (44.2) and (44.3),
t = rR + rP, f = (r - r)R + 2rP;
hence, from (1),
K_ Ire+2r2-rrI
(5)
(r2+t2)§
Example. Find the vectors T, N, B and the curvature and torsion of the
twisted cubic
x=2t, y=t2, z=t3/3
at the point where t = 1.
1 47 FUNDAMENTAL THEOREM 97

Since r(t) _ [2t, t2, 0/31, we have


f = [2, 2t, t2], 1' = [0, 2, 2t], 7 = [0, 0, 2];
hence, when t = 1,
f = [2, 2, 1], i = [0, 2, 2], f = [0, 0, 2],
t x 1' = [2, -4,41, fxi' f = 8.
Since T, B are unit vectors in the directions of t, t x t, and N = B x T,
T = 13 [2,2,11, N=1t2,-1,-2],
3 B=1[1,-2,21.
3

Moreover, from (1) and (2), K = $, r=$


47. Fundamental Theorem. Two curves for which the curvature
and torsion are the same functions of the arc are congruent.
Proof. Let the curves have the equations r = r1(s), r = r2(s).
Bring the origins of are and the trihedrals T1N1B1, T2N2B2 at these
points into coincidence. At the points P1i P2 of the curves, corre-
sponding to the same value of s, consider the function,
(1) f(s) =
By Frenet's Formulas,
d K1N1 T2 K2N2 T1

ds
= +
(-K1T1 + 71B1) N2 + (-K2T2 -- T2B2) . N1
T1N1 - B2 - T2N2 B1;
and, since K1 = K2, T1 = T2 for the same value of s,
-- = 0, f (s) = C, a const.
ds
But at the point s = 0, the trihedrals coincide, and
C=f(0)=1+1+1=3.
Hence for any value of s, f(s) = 3, an equation that implies that
each scalar product in (1) equals one; consequently,
T1 = T2, N1 = N2, B1 = B2.
From the first of these equations,
dr1 dr2
' rl = r2 + a;
ds ds
and a, the vector constant of integration, is zero, since r1 = r2
when s = 0. Therefore r1(s) = r2(s), and the curves coincide.
98 VECTOR FUNCTIONS OF ONE VARIABLE § 48

48. Osculating Plane. The osculating plane of a curve at a


point PI is defined as the limiting plane through three of its points
P1, P2, P3 as P2 and P3 approach P1.
Let the points P1, P2, P3 of the curve r = r(s) correspond to the
are values 81, s2i 83 (sI < s2 < s3). Then, if n is the unit normal
to the plane P1P2P3, the function,
f(s) = [r(s) - rl] n
vanishes when s = s1i 82, s3. Hence from Rolle's Theorem,
f(s) = r'(s) n
vanishes twice, and
f"(s) = r"(s) n
vanishes once in the interval s1 < s < 83. Consequently, as P2
and P3 approach P1, n approaches a limiting vector n1i such that
r'(sl) n1 = T1 n1 = 0,
r"(sl) n1 = K1N1 nl = 0.
If ICI 76 0, n1 is perpendicular to both T1 and N1i that is, parallel
to B1. The osculating plane at a point of a curve is the plane of the
tangent and principal normal at this point. For a plane curve
T and 'N lie in the plane of the curve-the osculating plane at all
of i points.
F r the curve r = r(t), B is parallel to t x 'r (§ 46); hence the
equation of the osculating plane at any point r is
(1)

her e R is the position vector to any point of the osculating plane.


The osculating planes of a curve form a one-parameter family.
The osculating planes at the points P1 and P2 intersect in a straight
line; and, as P2 approaches P1, the limiting position of this line
is called the characteristic of the osculating plane at P1. To find
the characteristics of the osculating planes to the curve r = r(s),
we must adjoin to their equation (R - r) B = 0, the equation ob-
tained by differentiating it with respect to s, namely,
or
if r 0 0. The two equations,
(R-r)B=0, (R-r)N=0,
§ 49 CENTER OF CURVATURE 99

represent planes through the point r of the curve perpendicular


to B and 11, respectively; together, they represent a line through
the point r parallel to N x B = T. The characteristics of the oscu-
lating planes of a skew curve are tangents to the curve.
Example. For the twisted cubic
x=2t, y=t z=t3/3
of the example in § 46, t x 'r = 2[1, -2, 2] at the point (2, 1, for which
t = 1. Hence, from (1), the equation of the osculating plane to3)the curve at
this point is
(x-2)-2(y-1) }2(z-3)=0.
49. Center of Curvature. The center of curvature of a curve
at the point P1 is defined as the center of the limiting circle through
three of its points P1, P2, P3, as P2 and P3 approach P1.
Since the limiting plane of P1, P2, P3 is the osculating plane
through P1, the limiting circle lies in this plane. If c and a denote
the position vector of the center and the radius of the circle through
the points P1, P2, P3 of the curve r = r(s), the function,
f(s) = [r(s) - c] [r(s) - c] - a2
vanishes for s = 81, 82i 83 (s1 < 82 < s3). Hence
f(s)
vanishes twice, and
f"(s)
vanishes once in the interval s1 < s < s3.
As P2 and P3 approach P1, c and a approach limits c1, a1, such
that
(r1 - c1) (rl - c1) - ai = 0,
(r1 - c1) Tl = 0,
(r1 - c1) Nl + pi = 0.
Since r1 - cl lies in the osculating plane, its components on
T1, N1, B1 are 0, -pl, 0, respectively, and r1 - cl = -p1R1. Thus
the center and radius of the limiting circle are
cl = r1 + p1 N1 and a1 = p1.
The center of curvature at any point P of the curve has the posi-
tion vector,
(1) c=r+pN.
100 VECTOR FUNCTIONS OF ONE VARIABLE § 50

This is a point on the principal normal at a distance p from P in


its positive direction.
Example. Curves of Constant Curvature. Let r = r(s) be a twisted curve t
of constant curvature K. The locus t1 of its centers of curvature has the equa-
tion,
(2) rl=r+pN.
Differentiating with respect to s, we have
ds1 1
T1 = T + P(rB - KT) = B.
ds K

Choose the positive direction on rl so that Ti = B; then


dal
(3) rK
(8

Differentiating T1 = B with respect to s now gives


K1N1ds1_
ds -7,N, or K1N1 = - KN;

hence N1 = -N and K1 = K. Thus r1 is also a curve of constant curvature.


Since (2) may be written
(4) r = r1 + p1N1,
r is also the locus of the centers of curvature of r1. Consequently from (3)
we also have d8/ds1 = r,/K1; hence rr1 = KK1. The two curves thus have the
following relations:
(5) T1 = B, Ni = -N, B1 = T; K1 = K, 771 = KK1.

Twisted curves of constant curvature may be associated in pairs, each curve


being the locus of the centers of curvature of the other.

50. Plane Curves. For plane curves the torsion is zero and
Frenet's Formulas reduce to
dT dN
(1) = KN = -KT.
ds ds
Since dT/ds is always directed towards the concave side of a plane
curve, the same is true of N. The osculating plane at any point P
is the plane of the curve; and, if K 0 0, the center of curvature P1
is given by
(2) rl = r + pN.

At points of inflection dT/ds = 0, K = 0, and N ceases to be de-


fined. As we pass through a point of inflection, N abruptly re-
verses its direction, and hence B = T x N does the same.
150 PLANE CURVES 101

To remedy this discontinuous behavior of N and B at points of


inflection, the following convention often is adopted in the differ-
ential geometry of plane curves. Take B once and for all as a
fixed unit vector normal to the plane of the curve, and define
N = B x T. As before T, N, B form a right-handed set of orthog-
onal unit vectors. The curvature K is defined by (1) and is posi-

Fia. 50a FIG. 50b

tive or negative, according as N has the direction of dT/ds or the


opposite. Equations (1) still hold good; for
dN dT
- =BxdB =BXKN=-KT.
Let 4, be the angle from a fixed line in the plane to the tangent
at P, taken positive in the sense determined by B (Fig. 50a).
Then, from (44.2),
dT dT do d4, d#
ds = dt' ds
= BxT- = N-
ds ds
and hence, from (1),
(3) K=a,
d4,
p=dds .
The locus of the centers of curvature of a plane curve is called
its evolute.Let P and P1 be corresponding points of a curve r and
its evolute r1 (Fig. 50b); and s = AP and sl = A1P1 denote cor-
responding arcs. On differentiating the equation (2) of r1 with
respect to s, we have
ds1 dN dp dp
=T+pds+dsN=-N.
T1
102 VECTOR FUNCTIONS OF ONE VARIABLE §50

Choose the positive direction on F1 so that T1 = N; then


ds1 dp
and s1 = p + const.
ds ds'

Hence the tangent to F1 is normal to r; and, since As1 = Op, an


arc of F1 is equal to the difference in the radii of curvature of r
to its end points. These properties show that a curve may be
traced by a taut string unwound from its evolute; the string is
always tangent to F1 and its free portion equal to p. From this
point of view, r is called the involute of Fl.
From T1 = N, we have '1 = 4, + 7r/2; hence, from (3),
ds1
p1=d4,1 =
dp
d
d2s
4,2.
d
Example 1. The only plane curves of constant non-zero curvature are circles.
For from
dN dr dN
ds
= -KT, or
ds
= -p ds
,

we have, on integration,
r - c = -pN, or Ir-cl2 =p2
This is the equation of a circle of radius p and center c.

FIG. 50c

Example 2. An involute r1 of a plane curve r is generated by the point P1


of a taut string unwound from r (Fig. 50c). If t is the curve r - r(e), the
involute is given by
r1 =P r - sT.
§ 50 PLANE CURVES 103

Hence, on differentiation with respect to 8,


dsl
Tlds =T - T - sKN = -8rcN.
If we choose the positive direction on r, so that Ti = -N,
dsl d4, dsl
(i) = sx = s - , or = s.
d8 dp

The intrinsic equation of a plane curve is the relation connecting s and ',
say s = s(4,). For the are sl along its involute we have, from (i),

sl = f provided sl = 0, when ¢ = VO.


0

The intrinsic equation of a circle of radius r is s = ry& when >G is measured


from the tangent at the origin of axes. For its involute we have sl = rp2/2,
provided sl = 0 when' = 0.
The intrinsic equation of a catenary of parameter c is s = c tan when a
is measured from the vertex (V, - 0). t The are sl, measured along the involute
from the vertex of the catenary, is

sl = c f tan k dy, = clog see ¢.


0
1k

Example 3. Envelopes. Consider the plane vector function of two variables,


r = f(u, v). The one-parameter family of curves u = c (constant) is given by
rl = f(u, v), u = const.
If this family has a curve envelope given by v = p(u), namely,
(ii) R = f[u,,p(u)],
the vectors,
dR u_ of of drl _ of
and
au + av `P (u) dv av

are parallel at the points of contact; that is,


of of
- x -- = 0.
au av
This condition must be fulfilled if the envelope exists. If (iii) leads to a rela-
tion v = p(u) (which does not make of/au oraf/av zero), this relation gives the
envelope (ii).
t See Brand, Vectorial Mechanics, equation (97.3). In Fig. 97b (p. 202), the
string PA will unwrap into the position PL; as P moves along the catenary,
the locus of L is the involute of the catenary. The line LQ is tangent to the
involute (Tj = -N); and, since the x-axis intercepts a constant segment
LQ = c on this tangent, the involute of the catenary is a tractrix, a curve
characterized by this property.
104 VECTOR FUNCTIONS OF ONE VARIABLE § 50

Consider, for example, the one-parameter family of normals to the plane


curve r = r(s), namely,
rl = r(s) + v N(s) (8 = const.)
If they have an envelope,
or,xari = (T - v,cT)xN = (1 -vK)B = 0.
as av

Hence the envelope is given by v = 1/K = p, or


R = r(s) +'N(s),
namely the locus of the centers of curvature of the curve.
Example 4. Plane Caustic by Reflection. Light issuing from a point 0 is
reflected from a mirror r (Fig. 50d).
TR _ The curve enveloped by the reflected
rays is called a caustic. If R and e are
the unit vectors along the incident and
reflected rays, the reflected rays are the
one-parameter family of lines,
ri = r(s) + ve(s).
To find their envelope, we form the
equation :
arl arl _ (T de)
+ v dsJ x e =
(iv) 0.
Fia. 50d as x 49V

Denote the angles (i, R) = 0, (R, T) - y; then (T, e) = y by the. law of


reflection, and
4,=(i,T) -0+y, p=(i,e)=0+2y-24, -0.
Now
de
.a de d0\
=kxe(2'--)=kxe(2K-
dy& sin yl
r /
from (44.5). Substitution in (iv) now gives

ksin y-kv(2K- sinr y =0;


whence
1 1 2
(v)
v r p sin y

This equation determines v and the caustic curve.


Letting r --, co, we obtain a beam of parallel rays; then v - 1p sin y.
When r is a circular are of radius p with center on Ox, the caustic has a cusp
on Ox at a distance vi from r given by 1/vl + 1/r - 2/p.
§ 51 HELICES 105

Example 5. The tractrix is a plane curve for which the segment of any
tangent between the point of contact and a fixed line is constant. In Fig. 50e
the fixed line is the x-axis, the constant length c, and, if r = OP, rl = OP1,

r1=r+cT.
Hence if we differentiate with respect to s,
ds,
i = T + cKN;

F iu. 50e

on multiplying by j we have
sink,
or
ds
d = c tan V,.
Ifs =Owhen# -0, this gives
s = c log sec

for the intrinsic equation of the tractrix (cf. ex. 2).

51. Helices. A helix is a twisted curve whose tangent makes a


constant angle with a fixed direction. If e is a unit vector in the
fixed direction, and a is the constant angle, the defining equation
of a helix is
(1) (0<a<Zr)
106 VECTOR FUNCTIONS OF ONE VARIABLE § 51

Differentiating (1) twice with respect to s gives


dN
ea 0,
=0;
hence the Darboux vector 6 = N x (dR/ds) is parallel to e. Con-
versely, if S for a curve is always parallel to a fixed vector e,
dT
e-N=0, a--=0,
ds
and e - T=cosa,

where cos a is a constant of integration; the curve is therefore a


helix. The only twisted curves whose Darboux vector has a fixed
direction are helices. Thus helices are characterized by the prop-
erty,
dS
(2) SX-=0 (§ 42, theorem 2).
ds
The fixed direction of
S = TT+KB
is along the axis of the helix. Now
dS dr dK
(3) B
ds ds T + ds

for T dT/ds + K dB/ds = 0 (45.4); hence


dS d- d-\ d (K-
sx- = KT - T -- a = -'r2 N--
ds C ds A ds TJ

Thus 6 x (dS/ds) = 0 implies that the ratio K/r is constant, and


conversely. Helices are the only twisted curves for which the ratio
of curvature to torsion is constant.
If we put T = dr/ds in (1) and integrate, we obtain
e - r=scosa+c
for the component of r in the direction of e. Hence, if r1 denotes
the projection of r on a plane perpendicular to e,
(4) r = r1 + (s cos a + c)e.
Every helix therefore lies on a cylinder with generators parallel
to e. The plane curve r1 traced by r1 is a normal section of the
cylinder.
§ 51 HELICES 107

If s = POP is the arc along the helix, and s1 = POP1 the arc
along the normal section r1 through P0, we have, on differentiating
(4) twice with respect to s,
ds1
(5) T=T1-+cosae
ds

dsl 2
(6) KN = K1N1
C ds
From (5),
ds1 2 =
(T - e cos a) - (T - e cos a)
ds
= 1 -2cos2a+C082a sing a;

and, if we choose the positive direction on r1 so that s1 increases


with s,
ds1
(7) - = sin a.
ds

Remembering that K and K1 are essentially positive, we now have,


from (6), N = N1i and

(8) K = K1 sin2 a.

On differentiating e = T cos a+B sin a we find K/r =tan a; hence

(9) r = K1 sin a cos a.

Finally, on integrating (7), we obtain

(10) s1 = s sin a,
the constant of integration being zero, since both s and s1 are
measured from the-same point Po.
The only twisted curve of constant curvature and torsion is the
circular helix. For, since K/r is constant, such curves are helices;
and, from (8), K1 is constant; that is, the normal section is a
circle.
The circular helix is the only twisted curve for which the Darboux
vector is constant. For from (3) we see that 6 is constant when
and only when K and r are constant.
108 VECTOR FUNCTIONS OF ONE VARIABLE § 52

Example. The curve,


x = a cos t, y = a sin t, z = bt,
is a circular helix; for the curve lies on the cylinder x2 + y2 = a2, and, since
t=[-asint,acost,b],
cos (k, T) = k t/I t I = b/1/a2.+ b2 (const.).
Moreover
[-a cos t, -a sin t, 0],
T = [a sin t, -a cost, 0],
and, from (46.1) and (46.2),

s
_ ItXrI _ a
r
b
II I3 a2 +b2' ItxiI2 a2+ b2

We now find the Darboux vector S = k/1/a2 + b2; it is constant in magni-


tude and direction.
52. Kinematics of a Particle. To define the position P of a
particle moving along a curve r, choose a point PO of r from which
to measure arcs, and take a definite direction along r as positive.
Then, if the arc s = POP is given as a function of the time, s = f (t),
the motion of the particle is determined; for its position is given
at every instant. The speed v of the particle at the instant t then
is defined as
ds
(1) v=
dt

Thus the speed will be positive or negative, according as s is in-


creasing or decreasing at the instant in question.
The speed measures the instantaneous rate at which the par-
ticle is moving along its path, but gives no information about its
instantaneous direction of motion. We therefore introduce a
vector quantity, the velocity, which gives the rate at which the
particle is changing its position in both magnitude and direction.
Let 0 be a definite point of a reference frame a, say the origin of
a system of rectangular axes fixed in a rigid body. Then if r = OP
denotes the position vector of P, the velocity v of P, relative to ,
is defined as the time derivative of its position vector:

(2) v=
§ 52 KINEMATICS OF A PARTICLE 109

As P moves along r, r may be regarded as a function of the


arc s; hence
dr dr ds ds
T-
dt = ds dt = t

where T is the unit tangent vector to r at P in the direction of


increasing arcs. Thus the velocity and speed are connected by
the relation,
(3) V = VT.

The velocity of P is represented by a vector tangent to the path at P


in the direction of instantaneous motion and of length numerically
equal to the speed.
Finally we define the acceleration a of the particle as the time
derivative of its velocity :
dv d2r
(4) a
- dt dt2
From (3), we find
dv dT
a=-T+V t;

or, since T may be regarded as a function of s,


dT dT ds v
= (KN)V=-N
dt ds dt p

dv v2
(5) a= T+ N.
dt p

The components of the acceleration in the positive direction of


the tangent, principal normal, and binormal, are therefore
dv v2
(6) ae=- t, an- , P
as = 0.

The acceleration of a particle P is a vector lying in the plane to


the tangent and principal normal to the path at P. The tangential
component is the time derivative of the speed, the normal component
the square of the speed divided by the radius of curvature at P.
110 VECTOR FUNCTIONS OF ONE VARIABLE § 53

The acceleration will be purely tangential when the motion is


rectilinear (p = x) ; it will be purely normal when speed is con-
stant (dv/dt = 0).
The velocity and acceleration vectors are regarded as localized
at the moving particle.
With rectangular axes, r = xi + yj + zk, and
dr dx dy dz
(7 ) v
dt dt 1 + dt
l+ dt
k,

dv d2x d2y , d2z


(8) a 1 + dt2 J + k.
dt dt2 dt2

The rectangular components of v and a are the first and second


time derivatives of x, y, z.
Example 1. Circular Motion. In the case of motion in a circle of radius r.
de dv _ d20
s=rO, v=rdt, dt=rdt2,
the angle 0 being expressed in radians. On writing w = dB/dt for the angular
speed, we have
dw
v = rw, at = r , an = rw2.

These results may also be deduced directly from the equation r = rR of the
circle if we make use of (44.2) and (44.3).
Example 2. Uniformly Accelerated Motion. If a particle has a constant
acceleration a, and r = TO, v - vo when t - 0, we have, on integrating dv/dt =
a twice;
v = at + vo, r = jat2 + vot + TO.
The path is the result of superposing the displacement "'ate, due to the accelera-
tion, upon TO + vot due to rectilinear motion at constant velocity vo. It is
easily shown to be a parabola having its axis parallel to a. At the vertex of
the parabola, v is perpendicular to a; the condition v a = 0 gives t =
-a vo/a a for the time of passing the vertex.
53. Relative Velocity. In § 52, we have seen how to find the
velocity v of a particle P relative to any given reference frame .
If ' is a second reference frame, in motion with respect to ,
how is the velocity v' of P relative to a' related to v?
At any instant t let P coincide with the point Q of a'. At a
later.instant t1 = t + At, let P and Q have the positions P1, Q1;
§ 53 RELATIVE VELOCITY 111

here Q is a fixed point of a', and its motion is due to the motion
of ' relative to a. Then
PP1 is the displacement of P relative to ,

QQ1 is the displacement of Q relative to a,


Q1P1 is the displacement of P relative to ',
and
PP1 = PQ1 + Q1P1 = QQ1 + Q1P1
If we divide this equation by At and pass to the limit At -* 0, we
obtain
(1) VP = VQ + VP, -,-

where VQ, the velocity of the point Q of R' relative to a, is called


the transfer velocity of P.
If we regard the frame as "fixed" and velocities referred to it
as "absolute," while velocities referred to a' are "relative," we
may state (1) as follows:
The absolute velocity of a particle is equal to the sum of its transfer
and relative velocities.
In many applications all points of the frame ' have the same
velocity relative to a; then vQ is the velocity of translation of ',
and we may write
(2) Vp = Vtr + VP-
Example 1. Wind Triangle. An airplane p has the velocity v relative to
the ground (the earth e), v' relative to the air; and the air (the wind w) has
the velocity V relative to the ground; then
v = V + v', from (2). In Fig. 53a,
v = ep, V = ew, v' = up;
thus vectors from e and w represent velocities
relative to the earth and wind, respectively.
The magnitudes of v and v' give the ground w
speed (ep) and air speed (wp) of the plane. The
directions of v and v', given as angles a and B'
measured from the north around through the Fla. 53a
east (clockwise), determine the track and
heading of the plane. The plane is pointed along its heading but travels
over the ground along its track. The angle (v', v) from heading to track
is the drift angle.
112 VECTOR FUNCTIONS OF ONE VARIABLE § 53

Example 2. Interception. A plane p is flying over the track PX with the


ground speed ep (Fig. 53b). As plane p passes the point P, a plane q departs
from Q to intercept plane p. If the air speed of plane q is given, over what
track shall q fly in order to intercept p?
Solution. In order that plane q may intercept plane p, the velocity of q
relative to p must have the direction QP.
Draw the vectors ep and ew, giving the velocities of plane p and the wind
w relative to earth e. With to as center describe a circle having the known air
speed of plane q as radius. If a ray drawn through point p in the direction of

,,q

Fia. 53b
-4
QP cuts the circle at point q, plane q will intercept plane p on flying with the
ground speed eq over the track QY parallel to eq; for pq, the velocity of plane
q relative to plane p, has the direction QP. Interception occurs at I after a
flying time of QI/eq (or PI/ep) hours.
Interception is impossible if the air-speed circle of plane q fails to cut the
ray. If the circle cuts the ray in just one point, as in Fig. 53b, plane q can
intercept p on only one track. But if the circle cuts the ray in two points,
say ql and q2, plane q can intercept p along two different tracks, parallel to
eql and eq2, respectively.
Example 3. Plane Returning to a Carrier. An airplane p leaves a carrier
a at 0 and patrols along the track OY while the carrier follows the course
OX with constant speed of v miles per hour (Fig. 53c). If the fuel in the tank
allows the plane T hours of flying time at a given air speed, at what point B
must the plane turn in order to rejoin the carrier at A, T hours after its depar-
ture? Find also the time ti, the heading, and the ground speed vl on the leg
out; and the time t2, the heading, and the ground speed v2 on the leg back.
Solution. Let the vectors es and ew give the velocities of carrier and wind.
With to as center describe a circle having the known air speed of p as radius-
the air-speed circle. If this circle cuts the ray through e in the direction OY
at pl, epl is the velocity of plane p on the leg out (ground speed vl = epl).
§ 53 RELATIVE VELOCITY 113

The course of plane p relative to the carrier s is out and back along the same
straight line. Now sp1 is the velocity of p relative to s on the leg out; hence
the velocity of p relative to s on the leg back is a vector sp2 whose direction is
opposed to sp1. Thus the point p2 is at the intersection of the air-speed circle
with the line pis prolonged, and ep2 is the velocity of p on the leg back (ground
speed V2 = ep2).
p2

Fro. 53c

After T hours the carrier will travel the course OA = T es. The return track
of p is along a line BA parallel to ep2. Thus p travels t1 = OB/ep1 hours on
the track OB, t2 = BA/ep2 on the track BA, rejoining the carrier after
t1 + t2 = T hours.
On the two legs the plane has the speeds ul = spli u2 = sp2i relative to the
carrier. At any time carrier and plane lie on a line parallel to p18p2. When the
plane is at B, the farthest point out, the carrier is at C(BC 11 pisp2). The dis-
tance r = CB is called the radius of action of the plane. Relative to the carrier
the plane travels the course CB out, BC back; hence

-+_=T, r= T
r

u1
r UIU2
U1 + u2
U2

t1 = r =T u1 U2+ u2 t2=r=T u1
.
U1 u2 u1 + u2
These times agree with those previously given; for

t1=-=-, CB OB - = BA
t2 = BC _-
spl epl 8P2 ep2

Example 4. Alternative Airport. Let a plane p depart from the airport 0


along the track OY (Fig. 53c). If a landing at an airport Y is rendered danger-
114 VECTOR FUNCTIONS OF ONE VARIABLE §54

ous by local bad weather, the plane may be directed to land at an alternative
airport A. If the fuel supply allows T hours flying time to the plane, how far
may the plane fly on the track OY in order still to reach the port A by changing
its course?
This problem is reduced to the one preceding if the airport A is regarded
as a carrier traveling from 0 to A with the uniform velocity OA/T. Lay off
es = OA/T and proceed precisely as in the carrier problem. The line CB,
drawn parallel to pisp2, fixes the farthest point B at which the plane can change
its course and still reach the airport A. The time t1 when the turning point is
reached is given in ex. 3.
54. Kinematics of a Rigid Body. We shall now investigate the
velocity distribution of the particles of a rigid body moving in any
manner.
Consider first a rigid body having a fixed line or axis; its motion
is then a rotation about this axis. The position of the body at

Fia. 54a Fia. 54b

any instant may be specified by the angle 0 between an axial plane


o fixed in our frame of reference and an axial plane p fixed in
the body (Fig. 54a). By choosing a positive direction on the axis
(unit vector e), we fix the positive sense of 0 by the right-handed
screw convention. Then the angular speed w of the body at any
instant is defined as
de
(1) dt

Thus w is positive or negative, according as 0 is increasing or de-


creasing at the instant in question.
§ 54 KINEMATICS OF A RIGID BODY 115

The velocity distribution in the revolving body may be simply


expressed if we define the angular velocity as the vector,
do
w = -- e.
dt

Note that w always is related to the instantaneous sense of rota-


tion by the rule of the right-hand screw.
Choose an origin 0 on the axis and let r = OP be the position
--
vector of any particle of the body. Then (Fig. 54b)

r=OQ+QP=ze+pR,
where R is a unit vector perpendicular to the axis and revolving
with the body. Since ze and p are constant during the motion of
P, the velocity of P is
dr dR do do (do \
V p e J x (ze + pR)
dt = do dt = p dt e x R = --
that is,
(2) v = wxr.
The velocity of any particle of a body revolving about a fixed axis is
equal to the vector product of the angular velocity and the position
vector of the particle referred to any origin on the axis.
Let us next consider a rigid body having one fixed point 0. Let
i be a unit vector fixed in the body, j a unit vector in the direction
of di/dt (perpendicular to i) and k = i x j. Then we may write
di dk di dj dj
dt
=aj, -=-xj+ix-=ix-
dt dt dt dt

Hence dk/dt (perpendicular to k) is also perpendicular to i and


therefore parallel to k x i = j. Thus we have
dk dj d
j3j, (k x i) = /3j x i + k x aj,
dt = dt dt
or, on collecting results
di dj dk
aj, = (ak - #i) x j, = Rj.
-= dt dt
116 VECTOR FUNCTIONS OF ONE VARIABLE 54

If we now write
w=ak - ii,
these equations assume the same form:
di dj dk
wxjf =
dt = wxl' dt = dt

Now the position vector r of any particle P in the body may be


Now
referred to the orthogonal triple i, j, k fixed in the body; thus

r=OP=xi+yj+zk,
where x, y, z remain constant during the motion. Hence the ve-
locity of P is given by
dr di dj dk
=xd6+ydt+zdt = wx(xi+yj+zk),
dt

(3) V = co x OP.

Thus, at any instant, the velocity distribution in a rigid body with


one point 0 fixed is the same as if it were revolving about an axis
through 0 with angular velocity co. The line through 0 in the
direction of co is called the instantaneous axis of rotation, and co is
called, as before, the angular velocity. Now, however, co may change
in direction as well as in magnitude. With a fixed axis of rotation,
do d
w= e= (oe),
dt dt

so that co may be regarded as the time derivative of the vector


angle Be. But with a variable axis of rotation co no longer can be
expressed as a time derivative.
Finally let us consider the general motion of a free rigid body.
If, at any instant, all points of the body have the same velocity v,
the motion is said to be an instantaneous translation. When the
velocity distribution is given by v = co -A P, the motion is said
to be an instantaneous rotation about an axis through A in the
direction of co. We now shall show that, in the most general
motion of a rigid body, the velocities may be regarded as com-
pounded of an instantaneous translation and rotation.
§ 54 KINEMATICS OF A RIGID BODY 117

Let A be any point of the rigid body, and denote its velocity
relative to by VA. Consider a second reference frame ' having
a translation of velocity VA relative to . Then the motion of the
body relative to a' is an instantaneous rotation about an axis
through A, since A has zero velocity relative to '. The velocity
of any particle P of the body is therefore
vy = wxAP,
relative to a', and, consequently,
(4) VP = V.4 + wxAP,
relative to R. Moreover, for any other point Q of the body,
(4)' VQ = vA + w x AQ;
and, on subtracting this from (4), we get
VP = VQ + 0) x QP.
The content of these equations is stated in the following
THEOREM 1. If A is any point of a free rigid body, the velocities
of its points are the same as if they were compounded of an instan-
taneous translation VA and an instantaneous rotation w about an
axis through A; and w is the same for any choice of A.
Thus the instantaneous velocity distribution of a rigid body is
determined by two vectors, its angular velocity w and the velocity
VA of any point A of the body. These may be combined into the
velocity motor,

(5) V = Co + EVA;
for, from (4),
(6) VP= VA +PA xw,
in accordance with (31.3).
If w VA 0 0, the motor Visa screw. Since w VP = CO VA,
the velocities of all particles of the body have the same projection
on Co. The axis of V is called the instantaneous axis of velocity;
its equation, from (31.8), is
(wxv4)xw
(7) rxw = (origin A),
ww
118 VECTOR FUNCTIONS OF ONE VARIABLE § 54

a line parallel to co and passing through the point Q given by


AQ = Co x VA/0) - uo; and, from (4)',
_ w.VA
(8) VQ = VA + w x AQ = -co = projw vA.
w-w

All points of the instantaneous axis have the velocity VQ; for, if R
is another of its points,

VR = VA -I- w x (AQ -{- QR) = vQ since w x QR = 0.

Moreover vQ is characterized by being parallel to co; and, since all


particle velocities have the same projection on Co, their least nu-
merical value at any instant is j vQ I. Referred to a point Q on
the instantaneous axis, the velocity motor becomes
(9) V = to -f- evQ.

This form, in which co and vQ are parallel, symbolizes


THEOREM 2. At any instant, the particle velocities of a rigid body
may be represented by a screw motion-a rotation about an instan-
taneous axis combined with a velocity of translation along this axis.
If Co 0, VA = 0, we have vQ = 0. The motion is then a pure
rotation of angular velocity w about the instantaneous axis. V be-
comes a line vector along this axis.
If w = 0, Co VA 0, we see from (4) that all points of the
body have the same velocity. The instantaneous motion is then a
pure translation, and V = EVA is a pure dual motor.
Finally let us consider the case of plane motion of a rigid body.
If co 0, the plane of the motion is perpendicular to w and vQ = 0;
the motion is then an instantaneous rotation about an axis. The
point Q where the instantaneous axis cuts the (reference) plane of
motion is called the instantaneous center; and, as previously,
-- 9 wxVA
(10) AQ =
w - w

If the velocities of the points A, B of the body are known and


VA, VB are not parallel, Q is at the intersection of the lines AQ, BQ
drawn perpendicular to VA and VB, respectively.
§ 54 KINEMATICS OF A RIGID BODY 119

Example. Rolling Curve. Let the curve r roll without slipping over a
fixed curve r1 (Fig. 54c). The points A and Al were originally in contact;
and, if I is the instantaneous point of
contact, let s = arc Al, sl = arc All.
The conditions for pure rolling are
(1) S = Si, T = Tj at I. Vt=V
If I is regarded as a moving particle, its ve- Al
locity relative to fixed and moving frames
attached to r2 and r is the same; for Fia. 54c
dsl ds
VI = Ti = T = Vj.
dt dt

Hence, if I coincides with the fixed point Q of r, we have, from (53.1),


vl=VQ+Vj, VQ=0.
The instantaneous center of the rolling curve is at its point of contact with the
fixed curve.
The speed of rolling,
_ ds _ ds1
v
dt dt
The angular speed is co = de/dt, where 0 is the angle between the tangents at
A and A1. We now differentiate the equation Ti = T with respect to t. Since
T is a function of s and B, we have
dT1 dsl OT ds 8T de
dsl dt = as dt + ae dt '
KgN1v = KNv + Nw,
or, since Nl = N,
1 1 w
(iii)
P1 P v

Here N is ir/2 in advance of T and p and P1 are positive when N points to the
respective centers of curvature (in Fig. 54c, P1 < 0, p > 0 and w < 0).
r

(0>0;p1>0,p<0 6)>O;p1<0,p<0
Fla. 54d
120 VECTOR FUNCTIONS OF ONE VARIABLE § 55

If r and ri are circles of radius r and rl (Fig. 54d),


w _ 1 1) = I + I
for external contact,
V ri ` r) r ri

for internal contact.


V ri \ r/ r ri
When ri -+ co, ri becomes a straight line and v = wr.
55. Composition of Velocities. Let a rigid body have the ve-
locity motor,
V' = w' + EVA,
relative to a frame a'; and let the frame a' have the motor,
V = CO + EVB,
relative to a "fixed" frame , where B is the point of a' which
coincides for the instant with the point A of the body. What is
the motion of the body relative to a?
The velocity of any point P of the body relative to a' is

(1) Vp = VA + W' X AP.


At the instant in question let P coincide with the point Q of a'.
Then the velocity of Q relative to is
(2) vQ=vB+WXBQ.
Now the velocities of P and A relative to are (53.1)
Vp = VQ + Vp, VA = VB + VA.

On adding (1) and (2) and observing that AP = BQ at the instant


considered, we get
VP = VA + (w+W')XAP.
Thus the motion of the body relative to a is compounded of the
velocity of translation VA and the angular velocity w + w' about
an axis through A ; that is, the motion is given by the motor,
V+V'=w+w'+evp.
THEOREM. If the motion of a rigid body relative to a' is given
by the motor V' = w' + evA, and the motion of a' relative to by
V = w + eVB, the motion of the body relative to a- is given by the
§ 56 RATE OF CHANGE OF A VECTOR 121

motor sum V + V' provided A and B coincide at the instant in


question.
Two translations, ev' and evB thus compound into a translation
e(vA + VB).
If V' = co' + evA and V = w + evB represent pure rotations,
these motors are line vectors; hence the motion of the body rela-
tive to a will be a pure rotation when and only when w + w' 0
and the axes of rotation are coplanar (§ 32, theorem 1). If the
axes of rotation intersect at A, the motors referred to A are
V'=w', V=w; and V+V'=w+w'.
Angular velocities about intersecting axes may be compounded by
vector addition.
If V and V' represent pure rotations about parallel axes and
w + w' = 0, V + V' = e(vB + vi); the motion of the body rela-
tive to is then a pure translation of velocity vB + V.
56. Rate of Change of a Vector. Referred to a fixed origin 0
in the frame a, the vector u = PQ = OQ - OP; hence
du
(1) = VQ - VP1
dt
where vp, vQ are velocities relative to .

Similarly, if a' is a second frame in motion with respect to ,

the rate of change of u relative to a' is


d'u
(2) = ve - vp,
dt
where v p', vQ are velocities relative to '.
Let the motion of a' relative to be given by the motor
co + EVA, A being a fixed point of a'; and, at the instant in ques-
tion, let P and Q coincide with the points R and S of a'. Then,
from (53.1) and (54.4),
Vp = oP+OR =VP+oA + WXAR,
VQ = VQ+Vs =VQ+vA+wxAS;
and, on subtraction,
du d'u -> -
+ w X (AS - AR).
dt = dt
122
- --
VECTOR FUNCTIONS OF ONE VARIABLE

But, since AS - AR = RS = PQ = u,
--> --->
§ 56

du du
(3) d = + 0).U;
dt
in particular
du
(4) =wxu if u is fixed in a'.
dt
When u = w, the angular velocity of the frame a', we have,
from (3),
dw d'w
(5)
dt dt

The vector dw/dt is denoted by a and called the angular accelera-


tion vector of a.
Example. Kinematic Interpretation of Frenet's Formulas. At any point
P of a space curve, the trihedral TNB may be used as a frame of reference.
If P moves along the curve with unit speed, ds/dt = 1 and s = t if t = 0
at the origin of arcs. Then the arc s may be interpreted as the time, and
dr/ds = T is the velocity of P.
If w is the angular velocity of TNB referred to a "fixed" frame a, we have,
from (4),
dT dN dB
, x T, N, x B.
ds = ds = p1 x ds = "
A comparison with Frenet's Formulas (44.6) shows that 0) = S, the Darboux
vector. The Darboux vector of a space curve is the angular-velocity vector of its
moving trihedral TNB.
Since the vertex P of the trihedral has the velocity T, its motion is represented
completely by the velocity motor,
V =S+eT=rT+KB +Er.
From (54.7) and (54.8), we see that the axis of this motor passes through the
point Q given by
---> 8xT K ST r
PQ = = ,,2+,2N and vQ = S-S S = K2 + '28.
S 8

In general, the trihedral TNB has, at every instant, a screw motion: a combina-
tion of the angular velocity S about an axis through Q and the velocity vQ
along this axis. The velocity of translation vanishes when and only when the
curve is plane (r = 0); then S = KB, PQ = pN. For plane curves TNB has, at
every instant, a pure rotation of angular velocity KB about the center of
curvature.
§ 57 THEOREM OF CORIOLIS 123

57. Theorem of Coriolis. Let v and a denote the velocity and


acceleration of a particle P, relative to a frame , while v' and
a' denote these vectors relative to a frame ' in motion with re-
spect to . Then if 0 and A are origins fixed in and a', we have
dr
r= OP, v= -- ,
dt
a= dv-; dt

d'r' d'v'
r' = AP v'= a
dt , dt
If w is the angular velocity of ' with respect to , we have, on
differentiating,
-- 3
r = OA + r',
twice with respect to the time, and, on making use of (56.3),
dr'
V =VA+
dt
= VA +wxr'+v',
dr' dv'
a=aA+axr'+wx+ / dw
J
dt dt
(a dt
a A+ a x r' + w x ((o x r' + V) + w x v'+ a'
a4+axr'+wx(wxr')+2wxv'+a'.
The velocity and acceleration of the point Q of the frame j' with
which P momentarily coincides are called the transfer velocity and
transfer acceleration of P. To find them, put v' = 0, a' = 0 in the
foregoing equations; thus
VQ = VA + wxr',
aQ = aA + ax r' +wx
Our equations now read
(1) V=vQ+V',
(2) a=aQ+2wxv'+a'.
Equation (1) restates the theorem on the composition of velocities
already proved in § 55. Equation (2) shows that an analogous
theorem for the composition of accelerations is not in general true;
124 VECTOR FUNCTIONS OF ONE VARIABLE § 57

we have, in fact, the additional term 2w x v', known as the Coriolis


acceleration. If we regard the frame as "fixed" and rates of
change referred to it as "absolute," while the corresponding rates
referred to ' are "relative," we may state (2) as the
THEOREM OF CORIOLIS. The absolute acceleration of a particle is
equal to the sum of its transfer acceleration, Coriolis acceleration, and
relative acceleration.
The Coriolis acceleration, 2w x v', vanishes in three cases only:
(a) co = 0; the motion of ' relative to is a translation.
(b) v' = 0; the particle is at rest relative to '.
(c) v' is parallel to co.
A particle of mass m, acted on by forces of vector sum F, has
the equation of motion F = ma. When the
motion is referred to a rotating frame
the equation of motion becomes
N P(Q) (3) ma' = F - maQ - 2mw x v'.
If we regard -maQ and -2mw x v' as ficti-
U
V tious forces, ma' equals a sum of forces just
as in the case of a fixed frame. The term
O(A)
Fio. 57a -2mw x v' is called the Coriolis force. When
' has the constant angular velocity w about
a fixed axis through A, a = 0, aA = 0, and aQ = w x (w x r').
Taking 0 at A and the z-axis along w (Fig. 57a), we have

-maQ = mw2 (r' - k r' k) = mw2 (OP - ON) = mw2 NP;


this vector perpendicular to the axis and directed outward is called
the centrifugal force on the particle.
Example. Particle Falling from Rest. Refer the particle, originally at 0,
to the revolving frame Oxyz attached to the earth: Oz points to the zenith
(along a plumb line), Ox to the south, Oy to the east (Fig. 57b). The earth
revolves from west to east about the axis SN and its angular velocity at north
latitude a is
_ 2r
(k sin X - i cos a) radians/see.
24 X 602
When the particle is at rest relative to the earth, the force acting upon it
is its local weight mg; hence in (3) F - maQ = mg. Therefore the equation
of motion becomes
(i) a'=g-2o,xv',
§ 57 THEOREM OF CORIOLIS 125

where a' and v' are the acceleration and velocity relative to the earth. We
shall integrate (i) by successive approximations under the initial conditions
= 0, v' = 0 when t = 0 and with g regarded as constant.
1. If we neglect the term in w,
a' = g, v' = gt, =zgt2.

The displacement r' is along the plumb line.

Fla. 57b

2. With v' - gt, (i) gives


a' = g - 2tw x g,
v'=gt-12wxg,
' = zr gt2 - 3 taw x g.

Since g = -gk, the second term gives an eastward deflection,


- 3 taw x g = 3 t3wg cos Xj.

3. With the last value of v',


a' = g - 2tw x g + 2t2w x (w x g),
v' = gt - t2w x g + 3 taw x (w x g),

= 2gt2 - 3t3w x g + Gt4w x (w x


g).
126 VECTOR FUNCTIONS OF ONE VARIABLE § 58

The last term is a deflection in the meridian or xz-plane. Since w x g = -wg


cos X j,
wng
w x (w X g) = cos x (i sin X + k cos X),
r= 1 9 sin X cos X i +
s tow"
1 taw
3 9 cos X J-
' 1 t2 - 114w2
(29 6 9 cos'- X) k.

The first and second terms give deflections to the south and east; the third
shows that the particle in t seconds falls through a distance
-z = 2gt2 - I t4w2g cost X.

58. Derivative of a Motor. Let the motor M= m + EmA be


referred to the frame a' in which A is a fixed point; and let '
itself have the motion V= o + EV A with respect to the frame .
If m and mA are functions of the real variable t, the derivative of
M relative to a' is
d'M dm d'mA
(1) e
dt dt + dt
In order to compute dM/dt relative to a we must remember
that mA depends not only upon t but also upon the point A, which
is in motion relative to a. If A moves to B in the interval At,
Om.4 mB(t + At) - m4(t)
At 'At

MA (t + At) + AB X m(t + At) - m.4 (t)


(31.3)
At

mA (t + At) - mA (t)
At
+ OB
-- At- OA X m(t + At),
and hence, if rA ° OA,
AMA dmA drA
xm.
c l-.o At dt - + dt
Relative to the frame a, we therefore define
dM dm dmA drA
(2) m)
dt dt dt + dt
To verify that dM/dt is a motor, we must show that
dmP drp dmA dr,4 -) dm
dt + dt x m = dt -
+ dt
x m + PA X
dt
§ 58 DERIVATIVE OF A MOTOR 127

This equation, in fact, follows from


mp = mA + (rA - rp) x m
on differentiation.
If t denotes time,
drA
VA;
dt =
and
dm d'm dmA d'mA
wxm+
dt = dt ' dt = dt '
from (56.3). Substituting these results in (2) gives
dM d'M
(3) = (0-M+ E((axmA + VA xm)
dt + dt
or, in view of (34.1),
dM d'M
(4) V X M +
dt dt
This is the motor analogue of (56.3).
Making use of the definition (2), we may verify that the follow-
ing rules of differentiation are valid :
d dM dN
( 5)
dt ( M + N) dt + dt '
d dM dA
(6) (XM) = X M,
dt dt + dt
d dN dM
( 7) (M -N) = M '
dt dt +N dt '

( 8)
d
dt
(M x N) = Mx
dN
dt +
-N
dM
dt
.

Example 1. If the motion of a rigid body is given by the velocity motor


V = w + EVA (A a fixed point of the body), its acceleration motor is

dV = a+e(aA +VAXW).

The axis of this motor is called the instantaneous axis of acceleration; its equa-
tion may be written from (31.8).
128 VECTOR FUNCTIONS OF ONE VARIABLE § 59

Example 2. Let F = f + EfA be a force acting on a rigid body whose velocity


motor is V = w + EvA. Then
dF
(i) f) .
dt + E ( dt + VA x
If F acts at the point P of the body, fp = 0 and
dF
(ii) = df + Evp x f, referred to P.

Hence, referred to A,

(iii) d do+E(vPxf+APxdt/

let the reader show that (i) and (iii) are consistent.
Example 3. The moving trihedral TNB at the point P of a space curve
r = r(s) may be regarded as a rigid body having the velocity motor,
V=S+Evp=8+ET,
as the point P traverses the curve with unit speed. The three line vectors
T = T, N = N, B = B through P are fixed in the trihedral; hence, from (4),
B =VxB.
ds =VxT, dN =VxN,
These are Frenet's Formulas in motor form. Written out in full they become

T = KN, dN = -7-T + KB + EB, d- _ -rN - EN;


for example
dB
- =(S+ET)xB=-dB- EN=-TN-EN.

59. Summary: Vector Derivatives. The derivative du/dt of a


vector function u(t) is defined as the limit of Au/At as At approaches
zero. If u = OP is drawn from a fixed origin and P describes the
curve C as t varies, du/dt is a vector tangent to C at P in the direc-
tion of increasing t. If I u I is constant, C is a circle and du/dt is
perpendicular to u.
The derivative of a constant vector is zero. The derivatives of
the sum u + v and the products fu, u v, u x v are found by the
familiar rules of calculus; but for u x v the order of the factors must
be preserved.
§ 59 SUMMARY: VECTOR DERIVATIVES 129

If s is the are along a curve r = r(s), dr/ds = T, a unit vector


tangent to the curve the direction of increasing s. The unit prin-
cipal normal ft to the curve has the direction of dT/ds; and the
unit binormal B = T X N; then [TNB] = 1. The vectors of the
moving trihedral TNB change conformably to Frenet's Formulas:
dT
Kft = SXT,
ds
dN
-KT + TB = 6 X ft,
ds

dB
-rft = SXB;
ds

K is the curvature, r the torsion of the curve; and the Darboux vector
o = TT + KB is the angular velocity of the moving trihedral as its
vertex traverses the curve with unit speed (ds/dt = 1). For plane
curves r = 0.
If t denotes time, a particle P traveling along the curve r = r(t)
has the velocity and acceleration,
dr dv d2r
v a
dt ' - dt dt2

If v = ds/dt is the speed, and p = 1/K the radius of curvature of the


path,
dv v2
V = VT, a = -- T+ - ft.
dt p

If A is any point of a free rigid body, the velocities of its points


are given by the velocity motor,
V = w + EvA ;
here co, the angular velocity vector, is the same for any choice of A.
Since V is a motor, for any point P of the body,
Vp = VA + PA Xw = vA +(0 XAP.
If, at any instant, co 0, the body has a screw motion about the
axis of V, the instantaneous axis of velocity; this reduces to a pure
rotation if V is a line vector (co VA = 0). If co = 0 the motion is
an instantaneous translation of velocity VA.
130 VECTOR FUNCTIONS OF ONE VARIABLE
If the frame a' has the angular velocity relative to , the rates
of change of a vector u relative to these frames are connected by
du d'u
w .U.
dt dt +
A particle P has the velocity and acceleration v', a' relative to
frame a-', v, a relative to the frame a; then, if the motion of a'
relative to a is given by the motor V = co + EvQ, where Q is the
point of ' coinciding at the instant with P,
v = VQ +v', a = aQ +2wxv' + a'.
The term 2w X v' is the acceleration of Coriolis. When a' is in trans-
lation relative to , the velocity equation may be written v =
vl, + V'.
The derivative of the motor M = m(t) + emA(t) is defined as
dM dm dmA drA
at-_ Xm
at = at + + at

If t denotes time, and d'M/dt refers to a frame a' having the


motion V = co + evA, relative to a frame , then
dM d'M
dt V X M + dt

PROBLEMS
1. If r and x are the distances of a point on a parabola from the focus and
directrix, r - x = 0. Show that (R - i) - T = 0, and interpret the equation.
2. Prove that the tangent to a hyperbola bisects the angle between the
focal radii to the point of tangency. [Cf. § 44, ex. 2.1
3. An equiangular spiral cuts all vectors from its pole 0 at the same angle a
(R T = cos a). If (r, 0) are the polar coordinates of a variable point P on
the spiral, show that
(a) ds/dr = sec a, s - so = (r - ro) sec a;
1 dr
(b) = cot a, log r = (9 - Bo) cot a;
rd0 ro

(c) p = ds/dB = r/sin a;


and that the center of curvature is the point where the perpendicular to OP
at 0 cuts the normal at P.
PROBLEMS 13 1

4. If r = r(s) is a plane curve. r, show that


(a) r1 = r + eN (c cont.) is a parallel curve r1 (§ 45, ex.l at it ii: l dis-
tance c from r;
(b) Si = s - c, provided Si = s = 0, when ,y = 0;
(C) P1 = P - C.
5. A curve r = r(s) has the property that the locus r1 = r + rT (e const.)
is a straight line. Prove that the curve is plane, in fact, a tractrix. [Cf. § 50,
ex. 5.)
6. An involute of a curve r, r = r(s), is a curve r1 which cuts the tangents
of r at right angles. Prove that r has the one-parameter family of involutes,
r1 = r + (c - s)T (c = const.);
and, if we take T1 = N, ds1/ds = (c - s)K.
7. Show that the involute of the circle,
x = a cos t, y = a sin t,
obtained by unwrapping a string from the point t = 0 is
x1 = a(cost +tsint), y1 = a(sint - tonst);
and that s1 = Zat2 gives the arc along the involute.
8. The cylinders x2 + y2 = a2, y2 + z2 = a2 intersect in two ellipses, one
of which is
x = a cos t, y = a sin t, z = a cos t.
Show that its radius of curvature is
P = a(l + sine t)I/''2.
9. The equation of a cycloid is
x = a(t - sin t), y = all - cos t).
Prove that
t 1
(a) T = [sin , cos , ds/dt = 2a sin
2 2J 2

(b) >y = (i, T) = 2 - 2 , P = -4a sin 2 (44.3);

(c) The equation of its evolute is


x1 = a(t + sin t), y1 = -a(1 - cost).
10. Find the vectors T, N, B and the curvature and torsion of the twisted
cubic,
x = 3t, y = 312, z = 20,
at the points where t = 0 and t = 1. Write the equations for the normal and
osculating planes to the curve at the point t = 1.
11. Show that curvature and torsion of the curve,
x = a(3t - t3), y = 3at2, z = a(3t + t3),
are
K = r = 1/3a(1 + t2)2.
132 VECTOR FUNCTIONS OF ONE VARIABLE
12. Find the envelope of the family of straight lines in the xy-plane,
r = p(O)R + XP,
where R and P = k x R are the unit vectors of § 44, 0 = angle (i, R), and p
is the perpendicular distance from the origin to the line. Show that the
curve,
r, = pR + p'P
is the envelope, and that
Ti = T, dsl/de = p + p".
13. Find the envelope of the family of lines for which the segment included
between the x-axis and y-axis is of constant length c. (In Problem 11 put
p = c sins cos 0.1
Show that the envelope has the parametric equations,
x = c sin3 0, y = C C083 0;

and that the entire length of the curve is 6c.


14. Show that the curvature and torsion of the curve,
x = et, y = e-t, z= /2 t,
are
Sc = -T + a-t)2.
15. Verify that the curve,
x = a sin2 t, y = a sin t cos t, z = a cos t,
lies on a sphere. Show that the curve has a double point at (a, 0, 0) for the
parameter values t = ±r/2, and that the tangents to the curve at this point
are perpendicular.
16. If a curve r = r(s) lies on a sphere (r - c) (r - c) = a2, show that

r - C = -pN - TI--B.
dp
ds
Hence, prove that
d (ldP)
pr+ ds r ds
0

is a necessary and sufficient condition that a twisted curve (r d 0) lie on a


sphere.
17. The points on two curves r and r, are in one-to-one correspondence.
If T = T, at corresponding points, prove that
N = N1, B = B1, dsl/ds = K/Kt = r/T,.
When both curves cut the rays from 0 at the same angle, show that r, = cr;
and that s, = as, if both arcs are measured from the same ray.
18. A curve r is called a Bertrand curve if its principal normals are principal
normals of another curve r1; then r1 is also a Bertrand curve, and
rl = r + XN, Ni = eN, where e = 1 or -1.
PROBLEMS 133

Show, in turn, that


(a) X = c (const.), Ti d = (1 - cK)T + cTB;
11

(b) If = angle (B, Bl), taken positive relative to N,


B1 = B cos (P + T sin p, ET1 = T cos rp - B sin rp;
(c) = const., and
ds1 ds1
-ET1 K sin p - r COS gyp, Kl COS cp + T sin gyp;
ds = ds
ds1 1 - CK -Cr
(d) Eds- = cos p
_
sin ;

(e) C2TT1 = Sin2 (P, K1 = -E(Ti cot (p + 1/c);


since K1 = 0, the last equation determines E.
'(f) As to the Darboux vectors, Si ds1/ds = S.
19. The characteristics of a one-parameter family of planes whose homo-
geneous coordinates (§ 26) are (a(t), ao(t)), are the limiting lines of inter-
section of the planes corresponding to the parameter values t and t + h as
h -- 0. The line of intersection of the two planes has the homogeneous coor-
dinates (26.10),
{a(t) x a(t + h), ao(t + h)a(t) - ao(t)a(t + h) [;
and, if we divide both by h and pass to the limit h -+ 0, show that we obtain
(a x a', apa - aoa')
as the coordinates of the characteristics.
20. From the result of Problem 19 show that the three families of planes
associated with the twisted (r 0 0) curve r = r(s), namely,
(T, r T), normal planes,
(N, r N), rectifying planes,
(B, r B), osculating planes,
have as characteristics the respective families of lines:
[B, (r + pN) x B], parallel to B through the centers of curvature;
(8, r x 8), parallel to S through points of the curve;
(T, r x T), tangent to the curve (§ 48).
21. If a particle P of mass m is subject to a central force F = mf (r)R, where
r = OP and R is the unit vector along OP, its equation of motion is
z
dv
m dt2 = F or = f (r)R.

Prove that r x v = h, a vector constant; hence, show that the motion is plane
and that r = OP sweeps out area at a constant rate (Law of Areas).
134 VECTOR FUNCTIONS OF ONE VARIABLE
22. When mf(r) = --ymM/r2 in Problem 21 the particle P is attracted
towards a mass M at 0 according to the law of inverse squares:
dv
(a)
dt
= - -k R (k
Show, in turn, that
(b) rxv = h
(c) h = r2 R x dR ,

=kdR,

(d) dt xh
(e) v x h = k(R + e),
where a is a constant vector. From (b) and (e),
r x v h = h2, r vxh = kr(1 +acose)
where 0 = angle (e, r); thus obtain the equation of the orbit,
hz/k
(f)
1 + e Cos 0

a conic section of eccentricity a referred to a focus as pole. When e < 1,


prove that the orbital ellipse is described in the periodic time,
7, - area of ellipse _ 2a

and, hence,
h/2 k a'
(g) T2/a3 = 4,r2/-yM (Kepler's Third Law).
(See Brand's Vectorial Mechanics, § 177.)
23. If the plane Ax + By + Cz = 1, fixed in space, is referred to rectan-
gular axes rotating with the angular velocity w = [wl, w2, w3], show that
rdA dB dC]
[A, B, C] x [WI, (02Y w3].
l dt ' dt ' dt =
24. A particle P has the cylindrical coordinates p, p, z (cf. § 89, ex. 1). If
P moves in the pz-plane so that dp/dt and dz/dt are constant while the plane
itself revolves about the z-axis with the constant angular speed dip/dt = w,
find the acceleration of P,
(a) By direct calculation.
(b) By use of the Theorem of Coriolis.
25. A real, everywhere convex, closed plane curve with a unique tangent at
each point is called an oval. It can be shown that an oval has just two tangents
parallel to every direction in the plane. The distance between these tangents
is the width of the curve in the direction of the perpendicular. Prove that
the perimeter of an oval of constant width b is 7rb (Barbier's Theorem).
CHAPTER IV

LINEAR VECTOR FUNCTIONS

60. Vector Functions of a Vector. A vector v is said to be a


function of a vector r if v is determined when r is given; and we
write v = f(r). Since r is determined by its components, f(r) is a
function of two or three scalar variables according as r varies in a
plane or in space. A vector function may be given by a formula,
as f(r) _ (a X r) x r; or it may be defined geometrically. Thus if
r = OP varies over the points P of a given surface, and v = OQ
is the vector perpendicular on the tangent plane to the surface at
P, v = f(r).
A vector function f(r) is said to be continuous for r = ro if
(1) lim f(r) = f(ro).
r - ro

This means that, when the components of r approach those of ro


in any manner, the components of f (r) approach those of f (ro).
A vector function is said to be linear when
(2) f(r + s) = f(r) + f(s),

(3) f (Xr) = Af (r),


for arbitrary r, s, X. For example, linear vector functions are de-
fined by the formulas kr, a X r, a b r, in which k, a, b are constant.
It can be shown that, when a continuous vector function satisfies
the relation (2), it also satisfies (3) and is therefore linear.
Since we assume that (2) holds when r = s = 0, f(0) = 2f(0);
hence, for any linear vector function,
(4) f(0) = 0.

A linear vector function is completely determined when f(a,), f(a2),


f(a3) are given for any three non-coplanar vectors al, a2, a3. For, if
we express r in terms of al, a2, a3 as a basis,
r = xlal + x2a2 + x3a3,
135
136 LINEAR VECTOR FUNCTIONS § 61

we have, from (2) and (3),


(5) f(r) = x'f(a,) + x2f(a2) + x3f(a3)

Let a', a2, a3 denote the set reciprocal to a,, a2, a3 (§ 23); then,
if we write
f(ai) = b1, xt = r at,
f(r) may be written in either of the forms:
(6) f(r) = r (albs + a2b2 + a3b3),
(7) f(r) = (b,al + b2a2 + b3a3) r.
These formulas represent the most general linear vector function.
61. Dyadics. In linear vector functions of the form,
f(r) = alb, r+a2b2.r+...+anbn.r,
we now regard f (r) as the scalar product of r and the operator,
(1) = alb, + a2b2 + ... + anbn.
Assuming the distributive law for such products, we now write
(2) f (r) = ' r.
Following Willard Gibbs, we call the operator 4) a dyadic and each
of its terms aib1 a dyad. The vectors ai are called antecedents, the
vectors bi consequents.
While a b is a scalar and a x b a vector, the dyad ab represents
a new mathematical entity. Gibbs regarded ab as a new species
of product, the "indeterminate product." We shall find indeed that
this product conforms to the distributive and associative laws, but
is not commutative (in general ab 0 ba).
Since r follows. in (2), we call r a postfactor. If we use r as a
prefactor, we get, in general, a different linear vector function:
g(r) = r 4).

We proceed to develop an algebra for dyadics, laying down for


this purpose definitions for equality, addition, and multiplication
of dyadics.
Definition of Equality. We write 4) = NY when
(3) 4> r = 'Y r for every vector r.
§ 61 DYADICS 137

If s is an arbitrary vector, we have from (3)


s - (4) - r) = s - or r
Since s 1 and s ' are two vectors that yield the same scalar
product with every vector r, these vectors must be equal; thus
(4) s iD = s - T for every vector s.

Conversely, from (4) we may deduce (3). Therefore 4) = T when


either (3) or (4) is fulfilled.
The Zero Dyadic. We write = 0 when
(5) 4) r = 0 for every r.
The preceding argument shows that 4) = 0 also when
(6) s S. 4) = 0 for every s.
Definition of Addition. The sum 4) +' of two dyadics is de-
fined by the property,
(7) (4) +'Y) r= for every r.
If is given by (1), and
' = c1d1 + C2d2 + ... + Cmd,n,
4) +' = a1b1 + ... + anbn + c1d1 + ... + c..d,,,..
In the sense of this definition, 1 (or ') is the sum of its dyad
terms, thus justifying our notation. The order of these dyads is
immaterial; but the order of the vectors in each dyad must not
be altered, for, in general, ab r ba r and hence ab 54 ba.
The Distributive Laws,
(8), (9) a(b + c) = ab + ac, (a + b)c = ac + bc,
are valid. The proof follows at once from the definition of equality;
thus from
a(b + c) r = (ab + ac) r,
we deduce (8). We may now perform expansions as in ordinary
algebra, if the order of the vectors is not altered; for example:
(a+b)(c+d) =ac+ad+bc+bd.
138 LINEAR VECTOR FUNCTIONS § 62

If X is any scalar,
(10) (Xa)b = a(Xb),

and we shall write simply Xab for either member.


From (60.6) or (60.7) we conclude that any dyadic D can be
reduced to the sum of three dyads. To effect this reduction on 4)
as given by (1), express each antecedent ai in terms of the basis
ei, e2, e3,
ai = ale, + aie2 + a;e3,

expand by the distributive law, and collect the terms which have
the same antecedent: thus
4) _ (a1 ei + a2e2 + a3e3)bi = elf, + e2f2 + e3f3,
where f;
We also may reduce 4) to the sum of three dyads by expressing
each consequent bi in terms of the basis el, e2, e3, and then ex-
panding and collecting terms which have the same consequent;
then 4) assumes the form,
4) = giei + 92e2 + 93e3-

Thus it is always possible to express any dyadic so that its ante-


cedents or its consequents are any three non-coplanar vectors chosen
at pleasure.
If a,, a2, a3 are non-coplanar, a dyadic 4) is completely deter-
mined by giving their transforms (§ 60). If
4) - a, =b,, 4> - a2 = b2, (D - a3= b3,
and the set a', a2, a3 is reciprocal to al, a2, a3, we have explicitly
(11) 4) = bla' + b2a2 + b3a3.
For a physical example of a dyadic the reader may turn to
§ 116, where the stress dyadic is introduced. The name tensor, now
used in a much more general sense, originally was ajTplied to this
dyadic.
62. Affine Point Transformation. If we draw the position vec-
tors,
r=OP, r'=4 r=OQ,
from a common origin 0, the dyadic 4) defines a certain transfor-
mation of the points of space: to each point P corresponds a defi-
§_63 COMPLETE AND SINGULAR DYADICS 139

nite point Q. If, when P ranges over all space, Q does likewise,
this transformation is called affine; the dyadic 4) then is called
complete.
Important properties of an affine point transformation follow at
once from the equations,
4) (a+b) =
which characterize a linear vector function. Since - 0 = 0, the
transformation leaves the origin invariant. Lines and planes are
transformed into lines and planes. Thus, for variable x, the
Line r = a + xb - Line r' = a' + xb';
and, for variable x, y, the
Plane r = a + xb + yc - Plane r' = a' + xb' + yc'.
The transformed equations always represent lines and planes when
4) is complete; for we shall show in § 70 that b 5-4 0 implies b' F6 0,
and b x c 5x-I 0 implies b' x c' 54 0.
63. Complete and Singular Dyadics. Given an arbitrary basis,
e1, e2, e3, we can express any dyadic in the form,
(1) 4) = glee + 92e2 + g3e3
If we express r in terms of the reciprocal basis,
r = xle' + x2e2 + x3e3,
then, by virtue of the equations, et e' = St,
4) r = x1g1 + x282 + x383
When g1, g2, g3 are non-coplanar, r' = 4) r assumes all possible
vector values as r ranges over the whole of space. If we put
---3 --a
r = OP, r' = OQ, the dyadic defines an affine transformation
r' _ ( r of space into itself; 3-dimensional P-space goes into
3-dimensional Q-space. A dyadic having this property is said to
be complete. A complete dyadic cannot be reduced to a sum of
less than three dyads; if, for example, we could reduce 4) to the
sum of two dyads, ab + cd, all vectors r would transform into
.vectors r' = a b r + c d r parallel to the plane of a and c.
140 LINEAR VECTOR FUNCTIONS § 63

If, however, g1, g2, g3, are coplanar, but not collinear, we can ex-
press each gi in terms of two non-parallel vectors f1, f2, and reduce
c to the sum of two dyads:
(2) 4, = f1h1 + f2h2.

This dyadic transforms all vectors r into vectors r' = 1 - r in the


plane of f1 and f2; 3-dimensional P-space goes into 2-dimensional
Q-space. A dyadic having this property is said to be planar. A
planar dyadic cannot be reduced to a single dyad ab; for then all
vectors r would transform into vectors a b . r parallel to a.
If 91, g2, g3 are collinear, we can replace each gi by a multiple
of a single vector f and reduce 4' to a single dyad:
(3) 4 = fh.
This dyadic transforms all vectors r into vectors r' = fi - r parallel
to f; 3-dimensional P-space goes into 1-dimensional Q-space. Such
a dyadic is called linear.
Finally, if g1, g2, g3 are all zero, 4' = 0.
Planar, linear, and zero dyadics collectively are called singular.
The point transformation,
OQ = (D -OP,
corresponding to a singular dyadic 4' reduces 3-dimensional P-space
to a 2-, 1-, or 0-dimensional Q-space.
This discussion shows that, when - is reduced to the form (1)
in which the consequents are non-coplanar, then 4, is complete,
planar, linear, or zero, according as the antecedents are non-
coplanar, coplanar but not collinear, collinear, or zero.
In particular, we have the
THEOREM. A necessary and sufficient condition that a dyadic
al + bm + cn be complete is that the antecedents a, b, c and conse-
quents 1, m, n be two sets of non-coplanar vectors.
As a corollary,
(4) 4) - r = 0 implies r = 0 when - is complete.
For, if - = al + bm + cn,
al-r+bm-r+cn-r=0, 1 - r = m - r = n - r = 0,
and hence r = 0.
§ 64 CONJUGATE DYADICS 141

64. Conjugate Dyadics. The dyadics,


4) = alb, + a2b2 + ... + anon,
,D, = b1a1 + b2a2 + + bn.an,

are said to be conjugates of each other. In general 4) and (D, are


different dyadics; but evidently
(1) 4) r=r
define the same linear vector function. If two dyadics are equal,
their conjugates are equal; for 4, =' implies that
4)

r, and hence 4), by definition.


A dyadic is called :
Symmetric if 4), = 4),
Antisymmetric if 4 = - 4).
The importance of these special types of dyadics is due to the
THEOREM. Every dyadic can be expressed in just one way as the
sum of a symmetric and an antisymmetric dyadic.
Proof. For any dyadic 4), Ave have the identity,
4)+4), 4)
(1) = + = + 52;
2 2
since
4rc = = *I St, = = -Q,
2 2

41 and Sl are, respectively, symmetric and antisymmetric. More-


over 4) can be so expressed in only one way. For, if
+Q_N,,+9,
41

gave two such decompositions, we have, on taking conjugates,


q/ - SZ =
and hence _ V, Sl = S2'.
142 LINEAR VECTOR FUNCTIONS § 65

65. Product of Dyadics. If the transformations corresponding


to two linear vector functions,
v=4) - r,
are applied in succession, their resultant,

is a third linear vector function; for w1 + w2 corresponds to


r1 + r2, and Xw to Xr. This function is written
w=
and 'I' ( is called the product of the dyadics ' and 4), taken in
this order. The defining equation for the product 4) is there-
fore
(1) (* for every r.
From (1) we find that the distributive and associative laws hold
for the products of dyadics:
(2) 4) ('I' + S2) -{- 4 S2,

(3) ((F +') +


(4)
Proofs. For every vector r,
(I' + l)] r = 4 [( + 0) r]
r+l r]
r)+ (St r)
r+ r;
[( +'F) St] r = ((F +'I') (St r)
r)+ r)

r= [(`I' ' ) ' r]

r)
§ 65 PRODUCT OF DYADICS 143

Equations t2), (3), (4) follow from these results by the definition
of equality. In the proofs of (2) and (3), definition (1) justifies
the first and last steps; in the proof of (3), (1) is applied in every
step.
In order to compute a product explicitly, we first find the
product of two dyads. By definition,

for every r, and hence


(5) (ab) (cd) = (b c) ad.
The product of ab and cd, in this order, is the scalar b c times
the dyad ad. Similarly,
(6) (cd) (ab) = (d a) cb,
which in general differs from (5).
Making use of the distributive law, we now may form the prod-
uct of any two dyadics,
n m
4) = E aibi, ' = E c;d;,
i=1 i=1
by expanding into nm dyads.
n m
(7) =E (bi c,)aidr
i=1 j=1

The conjugate of 4) ' is


m n

(8) (I' ) = Ej=1 i=1


71 (c; bi)d;ai

The conjugate of the product of two dyadics is the product of their


conjugates taken in reverse order.
Making use of (8), we now find that
r (4) - *) = (4) 'P ) , r = ( * ,-( 1 ) , )-r
hence
(9) r ((D 'Y) = (r (D) 4, for every r,
an associative law analogous to (1) but with r as prefactor.
If el, e2, e3 form a basis and e', e2, e3 the reciprocal basis, we
can express any two dyadics in the form,
4) = fie, + f2e2 + f3e3, ' = elg1 + e2g2 + e3g3
144 LINEAR VECTOR FUNCTIONS § 66

In the product 4 - ', six dyads vanish, and we find


(10) 4) `I' = figs + f2g2 + f393-
If 4) and' are complete, fl, f2, f3 and g1, g2, g3 are non-coplanar
sets; then (- 4, is also complete. But if (F or 4, is singular, one
of these sets must be coplanar, and 4) 'Y is likewise singular.
Therefore the product of two dyadics is complete when and only when
both dyadic factors are complete.
If 4) is complete, it may be "canceled" from equations such as
4.4,=O,4, (F=0, to giveT =O. Thus, if
are non-coplanar in (10), and hence g1 = g2 = g3 = 0. We also
may "cancel" - in (F = 4) St; for this is equivalent to
(4, - 9) = 0, and hence 4, - SZ = 0.
66. Idemfactor and Reciprocal. The unit dyadic or idemfactor
I is defined by the equation,
(1) I r=r for every r.
The idemfactor is unique (§ 60). For any basis e1, e2, e3, we have
I e; = e I e' = e'; hence, from (60.7),
(2) I = e1e' + e2e2 + e3e3 = e'e1 + e2e2 + e3e3;
in particular the self-reciprocal basis i, j, k gives
(3) I = ii + jj + kk.
Evidently I is symmetric and complete.
THEOREM. In order that a, b, c and u, v, w form reciprocal sets,
it is necessary and sufficient that
(4) au+bv+cw=I.
When the sets are reciprocal, we have just proved (4). Con-
versely, if (4) holds, a, b, c are non-coplanar, since I is complete;
let a', b', c' denote the reciprocal set. Using these vectors as pre-
factors on (4) gives u = a', v = b', w = c'.
Multiplying a dyadic by I leaves it unaltered; for, from
r (I.4)) _
we conclude that
(5) 4I=I (F = 4.
§ 66 IDEMFACTOR AND RECIPROCAL 145

If 4 I = I, 4) and ' are both complete, since I is complete


(§ 65). From
T or (4,

4) = I. Two dyadics 4), 4, are said to be recip-


rocals of each other when
(6) 4) 41 _I (D =I,
and we write 4, = 4)-1, 4) = T-1
A complete dyadic 4) has a unique reciprocal (D-1. If
(7) 4) = elf, + e2f2 + e3f3, (D-1 = flee + f2e2 + f3e3,

for their products in either order give I:


4) . 4>-I
= e,e' + e2e2 + e3e3, 4)-1 4) = f If, + f'f2 + f3f3.
Since dyadic multiplication is associative
(4) F) (T-1 4,-I) _ 4) ('F -1) 4,-1 = 4) (D-1 = I;

(8) (4) . *)-I = 'F-I .,I>-,.


Making use of (8), we have
(4) . * . Q) -1 = 2-1 , (4) q') -1 =
ggeneral,
and, in the reciprocal of the product of n dyadics is the
product of their reciprocals taken in reverse order.
For positive integral n we define
4,n
= (P .
4)-n = (4)-i)n = (4)n)-I;

4)°=I.
With these definitions,
(12) 4)m . 4,n = Pm+n, (4)m) n. _ 4)mn,

for all integral exponents. But owing to the non-commutativity


of the factors in a dyadic product, (4>. ')n is not in general equal
,0 cpn . *n.
146 LINEAR VECTOR FUNCTIONS § 67

By means of reciprocal dyadics, we readily may solve certain


vector and dyadic equations. Thus, if (F is complete, the equations,
(F r=v, r 4 ) =v, 4 )-4 1=2, (F 2,
have the respective unique solutions:
r=4)-1 v,
67. The Dyadic (1) x v. If (F = Maibi, we define the dyadics:
(1) - x v = Ea1bi x v, v x 4) = Tv x aibi.
From these definitions we have the relations:
(2) (4) x v) r = 4) (vxr), r (v x (F) = (rxv) (F;

(3) (r-(P)-v,
When 4, = I, these give
(4)
(Ixv) = r = vxr.
Since these equations hold for any r,
(6) Ixv = vxI,
(7) (I x v), _ -Ixv.
Thus I x v is antisymmetric and planar; it transforms all vectors f
into vectors perpendicular to v.
Since
aibi (I x v) = aibi x v, (I x v) aibi = v x aibi,
the definitions (1) give
(8) (P (Ixv) _ xv, (Ixv) - 4) = vx(F.
Consequently, the operations v x and x v on vectors or dyadics
may be replaced by dyadic products (I x v) and (I x v).
For any vectors u, v, r we have
[I x (u x v)] r = (u x v) x r = (vu - uv) - r,
and hence
(9) 1 x (u x v) = vu - uv.
§ 68 FIRST SCALAR AND VECTOR IN\'AItIANT 147

From this identity we arrive at the general form of any anti-


symmetric dyadic.
THEOREM. If the dyadic f = 2;aibi is antisymmetric, and the
vector co = tai x bi,
(10) l = -21xw.
Proof. The conjugate of f is -St = 2;biai; hence
-2SZ = 2; (biai - aibi) = EI x (ai x bi) = I x w.
Corollary.Every antisymmetric dyadic is planar.
68. First Scalar and Vector Invariant. We may express a dyadic
(P = 2;aibi in various forms by substituting vector sums for ai, bi,
and expanding and collecting terms by applying the distributive
law. For all these forms there are certain functions of the vectors,
in terms of which is expressed, which remain the same. These
functions, which may be scalar, vector, or dyadic, are called in-
variants of the dyadic.
Each step of the process in changing = 2;aibi from one form
to another may be paralleled by the same step in transforming the
scalar and the vector,
(1) Cpl = tai bi,
(2) 4) = 1dai x bi,
obtained by placing a dot or cross between the vectors of each dyad
of (D. Each application of the distributive law in the transforma-
tion of 4) is also valid in the corresponding transformations of 01
and 4; and, just as 1 is not altered by these changes, the same is
true of the scalar cpj and vector 4. These quantities are therefore
invariants of 4 with respect to the transformations in question.
They are called, respectively, the scalar (or first scalar invariant)
and vector of the dyadic.
For example, if P = ij + jk + kk,
c1 1, 4)=ixj+jxk+kxk=k+i.
For the idemfactor I = ii + fl + kk, the scalar is 3 and the vec-
tor 0; and we obtain these same values if I is expressed in terms
of an arbitrary basis: I = ele' + e2e2 + e3e3. Again, for the
dyadic,
S2 = IxV = iixv+jjxv+kkxv,
148 LINEAR VECTOR FUNCTIONS § 69

we have co, = 0, and


w=ix(ixv)+Jx(Jxv)+kx(kxv) -2v;
thus S2 = -ZI x w, as in (67.10).
The scalar or vector of the sum of two dyadics is the sum of
their scalars or vectors: thus, if
(3) 4) =`I'+0; 'Pi=4i+Wi, 4=4+w.
It is principally to this property that these invariants owe their
importance.
In (64.1), we have expressed any dyadic 4) as the sum of a sym-
metric and antisymmetric dyadic:
(4)

The antisymmetric part S2 = (4) - 4) has the vector,


z

From the theorem of § 67 we now have


(5) SZ = -2Ix4.
Hence, from (4),
24=4) +F,,-Ix4,
(6) +Ix
From (6), we have the
THEOREM. A necessary and sufficient condition that a dyadic be
symmetric is that its vector invariant vanish.
69. Further Invariants. We may obtain further invariants of
the dyadic c = 1aibi by processes that are distributive with re-
spect to addition. The most important of these are the dyadic,
(1) 4)2 = 2;aixajbixbj,
i.l
called by Gibbs the second of I ; its scalar invariant,
(2) (P2 = 2 (ai x a3) (bi x bj);
i,)

and the scalar,


(3) (P3 = * (aixaj .ak)(bixbj .bk).
i.i.k
§ 69 FURTHER INVARIANTS 149

In (1) the summation is taken over all permutations i, j. When


i = j, the dyad vanishes; when j - i, the permutations i, j and
j, i give the same dyad, so that each dyad occurs twice in the
final sum; this doubling is avoided by the factor 2.
In (3) the summation is taken over all permutations i, j, k.
When two subscripts are the same, the term vanishes; when
i, j, k all differ, the 3! = 6 permutations of these subscripts give
the same term, so that each term occurs six times in the final sum;
this multiplication of terms is avoided by the factor s i
Let CF be reduced to the three-term form:
(4) CF = al+bm+cn.
The invariants considered thus far are now
(5)
(6) 4 = axl+bxm+cxn,
(7) 42 = bxcmxn+cxanxl+axblxm,
(8) IP2 = (b x c) (m x n) + (c x a) (n x 1) + (a x b) (l x m),
(9) (p3 = (a x b c) (1 x m n).
The numbers 01, P2, 03 often are called the first, second, and
third scalars of 4).
If CF is singular, the antecedents a, b, c or the consequents
1, m, n in (4) will be coplanar, and p3 = 0. Conversely, if = 0,
a, b, c or 1, m, n are coplanar sets and 4) is singular. Therefore a
dyadic is singular when and only when its third scalar is zero.
If V3 = 0, 4) must be planar, linear, or zero. * When CF is linear
it can be reduced to the form al and cF2 = 0. Conversely, if
4'2 = 0, 4) will be linear or zero; for, if we choose a non-coplanar
set a, b, c as antecedents in (4), b x c, c x a, a x b are also non-
coplanar, and `1'2 = 0 implies that
mxn = nxl = lxm = 0;
then 1, m, n are parallel or zero.
Therefore we may state the
THEOREM. Necessary and sufficient conditions that a dyadic CF be
Complete 'P3 3-' 0,
Planar are that <P3 = 0, CF2 0,
Linear CF2 = 0, (1) 0.
150 LINEAR, VECTOR FUNCTIONS § 69

We next compute the invariants of the dyadic _ 'D2. From


(7),
*2 = (c x a) x (a x b) (n x 1) x (1 x m) + cyclical terms
= [abc] [lmn] (al + bm + cn) ;
hence, from (9),
(10) *2 = t03-1-
We now may compute the three scalars of 4,:
2;
(11) V1 = P2, 412 = '3'P1, 4'3 =
the first follows from (8), the second from (10), and the third from
(b x c) (c x a) x (a x b) = [abc]2, (m x n) (n x 1) x (1 x m) = [lmn]2.
Finally, the vector invariant of ' is
4 = (bxc) x (mxn) + (cxa) x (nx1) + (axb) x (lxm).
If we express 4r in terms of a, b, c, the term in a is
ab lxm = al - (bxm+cxn) = al 4
from (6) ; hence
(12) 4r = (al + bm + cn)
It can be shown that all scalar invariants of 4) may be expressed
in terms of the six scalars,
(13) 'PI,'02,'3,'
This property is expressed by saying that these six scalars form a
complete system of invariants. When 4) is symmetric, 4 _ 4, = 0,
and the last three scalars vanish.
The third scalar of the product of two dyadics is equal to the product
of their third scalars. For any two dyadics 4), ' can be put in the
form,
4) = f1e1 + f2e2 + f3e3, = e'g1 + e2g2 + e3g3,
where e1, e2, e3 and e1, e2, e3 are reciprocal sets (§ 65); hence
[f1f2f3][e1e2e3][e'e2e3][919293]
'P43 = = [f1f2f3][g1g2g3],
which is the third scalar of
4' ' = f1g1 + f2g2 + f3g3
§ 70 SECOND AND ADJOINT DYADIC 1.51

Example. For the idemfactor I = ii + jj + kk, the second 12 = I, the vector


is zero, and the three scalars are 3, 3, and 1. From
(i) I = eiel + e2e2 + e3e3,
we have, for 12,

(ii) I = e2xe3e2xe3 +e3xe1e3xel +elxe2e1xe2,


and the vector,
elxel + e2xe 2 + e3 x e3 = 0.
The third scalar gives
[ele2e3][ele2e3] = 1.

From (ii) we have, for example,

I I. e3 = el x e2[e'e2e3], e3 = el x e2/[ele2e3]

Thus (i) and (ii) give a handy compendium of the properties of reciprocal
vector sets.

70. Second and Adjoint Dyadic. The second of the dyadic


4) = Eaibi, namely,
ialai
(1) (D2 = x a; bi x b;,
i.;
has the property,
(2) (4 u) x (4) v) = -4)2 (u x v).

Proof. We have
v)
i ;

= Mai x a; (bi u) (b; v)


i;

= Ma; x ai (b, . u) (bi v),


i;
on interchanging i and j. The left member also equals half the
sum of the two last expressions :

ij
-u)
or, with regard to (20.1),

"M(aixa;)(bixb;) (uxv) = 4)2 (uxv).


i;
152 LINEAR VECTOR FUNCTIONS § 70

The conjugate of 4'2 is called the adjoint of 4) and written 4'a.


The adjoint satisfies the important relation,
(3) 4)=gC3I.
Proof. Write (D = al + bm + cn; then
(4) 4)a = mxnbxc+nx1cxa+lxmaxb.
Choose for the antecedents a, b, c of 4) a non-coplanar set. (§ 61).
Then, if a', b', c' denote the reciprocal set, we have, by direct
multiplication,
D 4)a = (1 mxn)(abxc +b cxa + caxb)
_ (1 m x n) (a b x c) (aa' + bb' + cc')
=(P3I.
If we choose the consequents 1, m, n of 4) as a non-coplanar set
(then a, b, c may or may not be coplanar), let 1', m', n' denote the
reciprocal set. Then

(a b x c) x n) (1'l + m'm + n'n)


= (P31-

If 4) is complete, P3 0; then (3) shows that


(5) 4'-1 = 4'a/1P3
From (3) we also may show that, if u, v, w are any three vectors,
(6) [4) u, 4) v, 4) w] = ,P3[uvw].
Proof. Since 4'a is the conjugate of 4>2i we have, from (2),
(U V) 4)a,

(U V)

4)a 4) by P3I, we obtain (6).


We now can deduce important properties of the affine trans.
formation (§ 62) :
(7) r' r ((P3 $ 0).
The vector area u x v (§ 17) is transformed into
(8) u' x v' = 4'2 (u x v).
§ 71 INVARIANT DIRECTIONS 153

From (4), the third scalars of ad and P2 equal [abc]2[lmn]2 =


,p3 0. Since 4) and c2 are complete, r 7-4 0 and u X v 0 imply
r' 0, u' x v' 0 (thus filling a gap in § 62). An affine trans-
formation invariably changes lines into lines, planes into planes.
Moreover (7) transforms any parallelepiped [uvw] into another
[u'v'w'] whose volume is p3[uvw] according to (6). As any volume
can be regarded as the limit of a sum of parallelepiped elements,
the affine transformation alters all volumes in the constant ratio of
93/1.
71. Invariant Directions. We next seek those vectors r which
are transformed by 4) into scalar multiples of r, say
(1) - r = Ar.
If we write this equation,
It - r = 0 where '1 = 4) - Al,
it is clear that the multiplier A must make the dyadic T singular;
its third scalar 43 is then zero (§ 69). Conversely, if A is a root
of the equation 03 = 0, A is a multiplier of 4); for, when 4, is planar
or linear, 4, - r = 0 for all vectors r normal to the plane of the con-
sequents of T.
Let us write
4) = al + bm + cn, I = aa' + bb' + cc',
where a, b, c are non-coplanar and a', b', c' the reciprocal set;
then
4, = a(l - Xa') + b(m - Xb') + c(n - Ac'),
4'3 = [abc] [(1 - Xa')(m - Xb')(n - Xc')].
The second box product in 43 gives, on expansion,
[lmn] - A { [a'mn] + [b'nl] + [c'lm] }

+ A2 { [b'c'l] + [c'a'm] + [a'b'n] l - A3[a'b'c'].


Substituting for the primed vectors gives

.I bxc ... ' a .. .


a= [abc] ' b x c = [abc] '
154 LINEAR VECTOR FUNCTIONS §71

and then multiplying the entire expansion by [abc] yields


L'3 = [abc][1mn]
- A{(bxc) (mxn) + (cxa) (nx1) + (axb) (lxm)}
-A3,
or, on making use of (69.5), (69.8), (69.9),

(2) 43 = <P3 - A<P2 + A2,P1 - A3.

Denote the right member of (2) by f (A) ; then the cubic equation,

(3) 'P3 = f(A) = 0,


is called the characteristic equation of 4,. Since 4) - AI is singular
when f (X) = 0, the coefficients of f (X) depend only upon the nature
of the dyadic 4) and not upon the particular form in which it is
expressed. We thus have an independent proof of the invariance
of <Pl, 'P2, 'P3
The three roots A1, A2i A3 of (3) are called the multipliers or
characteristic numbers of -4). From the relations between the roots
and coefficients of an algebraic equation, we have
(4) IP1 = Al + A2 + A3, <P2 = X2X3 + A3A1 + A1A2, P3 = A1A2A3.

The cubic (3) may have three real roots (not necessarily dis-
tinct) or one real root and two conjugate complex roots. To find
the invariant direction corresponding to a root A1i we consider in
turn the three cases in which 4) - X1I is singular.
1. If 4 - X11 is planar, let 4' - A1I = ch + dk; then the post-
factor r1 = h x k reduces the right member to zero and gives the
invariant direction for A1. If 4) - X1I has more than two dyads,
the cross product of any two non-parallel consequents will be nor-
mal to their plane and give a vector parallel to r1.
If Al is real, r1 is a real vector. But if Al is complex, say Al =
a + i(3, then r1 = a + ib is also complex. Then, on equating the
real and imaginary parts of
(5) 4) (a + Zb) _ (a + ia) (a + Zb),
we have
(6) (D a = as - $b, 4) b = #a + ab.
§ 71 INVARIANT DIRECTIONS 155

In this case we know a second multiplier,


X2= a-i$, with I2=a-2b,
as invariant direction. For, from (6),
4) (a - ib) = (a - i(3) (a - ib).
From (5), we conclude that a and b are not parallel; for, if b = ka,
4) a = (a + i(3)a, which is impossible unless # = 0.
2. If 4) - X1I is linear, it may be reduced to a single dyad ch.
Then, if r1 is any vector in the plane perpendicular to h, 4, r1
= X1r1. In this case there is a whole plane of invariant directions
corresponding to X1.
3. If (D - X1I is zero, 4 r1 = X1r1 for any vector r1. All direc-
tions are invariant with the multiplier X1.
If we write r' _ 4) r, we have, from (70.2),
(7) U" V' = (D2 (Uxv).
Now 4) transforms all vectors in the plane of u, v into vectors of
the plane u', V. These planes will be the same if u' x v' = x u x v,
that is, if u x v is an invariant direction of 42. Hence the invariant
planes of 4) are normal to the invariant directions of (P2.
Example 1. = ii + j(i + 2j) + k(j + 2k). Then
fit = i(4i - 2j + k) + j(2j - k) + 2kk,
pp1= 1+2+2=5, I'2=4+2+2=8, S3=4.
The characteristic equation,
f(T)=4-8a+5X2-X3=(1-X)(2-A)2=0,
has the roots, X1 = 1, X2 = X3 = 2.
For the root, X1 = 1,
- XiI = ji + jj + kj + kk = j(i + j) + k(j + k),
and the corresponding invariant direction is
r1= (i+j)x(j+k) =i-j+k.
For the double root, X2 = X3 = 2,
'P -1\2I= -ii +ji+kj = (-i+j)i+kj,
and r2 = i x j = k.
Example 2. 4) = ij + jk + Id. Then
'Pi=0, 92=0, IP3=1.
156 LINEAR VECTOR FUNCTIONS § 72

The characteristic equation, 1 - X3 = 0, has the roots,

X1 = 1, X2 =
+
= w, X3 =
-1 -2 = w2.
2
For X1 = 1,
4, -all=i(j-i)+j(kj)+k(i-k)
is planar- its consequents are all normal to
r1=(j-i)x(k-j)=i+j+k,
which is the corresponding invariant direction.
For X2 = w, the consequents of
4 -X21=i(j-wi)+j(k-wj)+k(i-wk)
are all normal to
r2 = (j-wi)x(k-wj) =i+wj+w2k.
For X3 = w2, the consequents of
4' -X3I=i(j-w2i)+j(k-w2j)+k(i-w2k)
are all normal to
r3 = (j - w2i) x (k - w2j) = i + w2j + wk.
72. Symmetric Dyadics. The characteristic numbers of a sym-
metric dyadic are all real.
Proof. If X1 = a + i(3 is a complex multiplier and rl = a + 2b
the corresponding invariant direction of the symmetric dyadic 4),
we have (§ 71)
,D(a+ib) = (a+ii)(a+ib);
c a = as -(3b, 4 ) b = 0a+ab.
Since cD is symmetric, b (D a=a 4) b; hence
or =0.
But a and b are real vectors, and a a + b b is positive; hence
(3 = 0, and X1 is real.
The invariant directions corresponding to two distinct character-
istic numbers of a symmetric dyadic are perpendicular.
Proof. From the equations,
4 rl = X1rl, `k r2 = X2r2 (X1 0 X2),
and
r2,
we deduce
(X1 - X2)rl r2 = 0, rl r2 = 0.
§ 72 SYMMETRIC DYADICS 1,57

Consider now the following cases.


1. If f (X) = 0 has three distinct roots, the corresponding in-
variant directions are mutually perpendicular and may be denoted
by i, j, k. Then
4, - i = X1i, 4 'j = X2j, 4) - k = X3k,
and, from (61.11),
(1) 4) = X1ii + X2JJ + X3kk.
2. If two roots of f (X) = 0 are equal, let X1 5-' X2 = X3. Then,
if i, j are the perpendicular invariant directions corresponding to
Xl and X2, we may write
4, - i = X1i, (D 'j = X2j, 4, - k = v,
v being the (unknown) transform of k; then, from (61.11),
4) = X1ii + X2jj + vk.
Now the symmetry of 4) requires that vk = kv, and hence v must
be a multiple of k, say yk. Thus
4' = X1ii + X2jj +ykk, <PI = X1 +X2+y;
but, since Cpl = X1 + 2X2 by hypothesis, -y = 1\2, and
(2) 'D = X1ii + X2jj + X2kk.
Corresponding to the double root X2 = X3, we have a whole plane of
invariant directions.
3. If all three roots of f(X) = 0 are equal, X1 = X2 = X3, let i
denote an invariant direction corresponding to X1, and write
4) i = Xli, 4> j = u, 4 ) k = v;
then, from (61.11),
= X1ii + uj + vk.
Now the symmetry of 4) requires that uj + vk = ju + kv, and
this in turn shows that u and v have the form,
u = aj + yk, v = yj + $k.
Thus
4) = X1ii + (ai + yk)j + (yj + $k)k,
(Pi = X1 + a + 0, o3 = Xl (a$ - y2);
158 LINEAR VECTOR FUNCTIONS § 72

but, since <pl = 3X1, X03 = X1 by hypothesis,


a + /3 = 2X1, a(3 - y2 = ai
Elimination of X1 gives the relation,
(a + 0)2 - 4(a/3 - y2) = (a - (3)2 + 4y2 = 0;
and, since neither (a - 0)2 nor 4y2 can be negative, their sum can
vanish only if a-i3=0,y=0. Hence a = 0 = X1, and 4> re-
duces to
(3) 4, = X1ii + A1jj + X1kk = XII.
In the case of a triple root all directions are invariant.
Evidently all cases are included in (1); by making two or three
of the roots equal, we obtain (2) and (3).
THEOREM. Every symmetric dyadic may be reduced to the form,
(4) 4, = aii + $jj + 7kk,
in which a, 0, y are the real multipliers and i, j, k corresponding in-
variant directions.
The reciprocal of 4) is
(15)
=-ii+-jj+-kk;
1 1 1

a a y
for .4 4)-1 = I. Moreover,
(6) 4'2 = f7ii + yajj + a$kk
and, by direct multiplication,
(7) 4)" = a"n + a"jj + y"kk.
If 4) is symmetric and 4)" = 0, then 4) = 0. Moreover, if
r 4 r = 0 for every r, - = 0; for, if we choose r = i, j, k in
turn, a=/3=y=0.
Example 1. For the symmetric dyadic,
(D = i(j +k) +j(k +i) +k(i +j),
w1 = 0, v2 = -3, 93 = 2, and the characteristic equation,
2+3X-X3=(2-a)(1+x)2=0,
has the roots, X1 = 2, X2 = X3 = -1.
§ 72 SYMMETRIC DYADICS 159

For X1 = 2,
-XiI=i(-2i+j+k)+j(-2j+k+i)+k(-2k +i+j),
The consequents are all normal to
(-2i+j+k)x(-2j+k+i) =3(i+j+k);
hence r1 = i + j + k.
For a2=X3= -1,
-X2I= (i + j + k) (i + j + k)
is linear. Hence any vector in the plane perpendicular to i + j + k is an
invariant direction.
If we choose the unit vectors,

i' = (i + j + k)/V'3, j' = (i - j)/V'2,


as invariant directions for the roots 2 and -1, respectively, then the unit
vector,
k' =i'xj' = (i+j -2k)/V,
gives a second invariant direction for -1. Therefore
4) =2i'i'-j'j'-kk',
as we may readily verify.
Example 2. Inertia Dyadic. Let 0 be any point of a rigid body and s an
axis through 0 in the direction of the unit vector e. Then, if dm is an element
of mass at P, at a distance p from the axis s, the moment of inertia of the body
about s is defined as the integral f p2 dm over the body. If r = OP,

p2 =r2- =e (r2I
hence, if we introduce the dyadic,

(8) K = I fr2dm - frr dm,


known as the inertia dyadic of the body for the point 0,

(9) e- fp2dm.
For example,
i.
f(r2-x2)dm= f(y2 +z2)dm
is the moment of inertia about the x-axis. Thus K effects a synthesis of the
moments of inertia of a body about all axes through 0.
This dyadic also has the property that, for any pair of perpendicular unit
vectors el, e2,
- el . K . e2 = f (el r) (r . e2) dm
160 LINEAR VECTOR FUNCTIONS § 73

is the product of inertia for the corresponding axes; thus

j=f j) dm = fxydm.
The inertia dyadic K is evidently symmetric. Hence we always can find
three mutually perpendicular axes x, y, z through 0 such that
(10) K=Ali+Bjj+Ckk.
These axes are called the principal axes of inertia at 0, and A, B, C are the
moments of inertia of the body about these principal axes. The principal
axes are characterized by the property that the product of inertia for any pair
is zero.
The ellipsoid,

(11) r K . r = 1, or Axe + Bye + Cz2 = 1,


is called the ellipsoid of inertia at 0; its principal axes are the principal axes of
inertia at O. It has the property that the moment of inertia about any axis s
through 0 and cutting the ellipsoid at P is 1/(OP)2. For, if OP = r = re,

73. The Hamilton-Cayley Equation. The identity (70.3) ap-


plied to the dyadic ' = - AI gives
(1) `y 'Ya = f(X)I.
The form of ' given in § 71 shows that we may write
(2) 'a = A + BA + CA2,
where A, B, C are dyadics independent of A; hence, from (1),
V1A2
(3) (4 - XI) (A + BA + CA2) = (-P3 - V2A + - A3)I.
Since (3) is an identity in A, the dyadic coefficients of like powers
of A in the two members must be equal, hence
4) A = (P3I,
-V217
(4)

-C= -I.
If we multiply these equations in order by I, 4), 42, (3 and add,
the first members camel, and we get
(5) te3I - (P24) + (P1 cy - (1,3 = 0.
§ 73 THE HAMILTON-CAYLEY EQUATION 161

Every dyadic 4) satisfies this cubic equation, the Hamilton-C,ayley


Equation; it evidently is formed by replacing X by 4) in the char-
acteristic equation,
(6) f(X) = V3 - cP2X + P1
X2
- X3 = 0,
and inserting I in the constant term. If the X1, X2, X3 are the char-
acteristic numbers of 4),
AX) = (X1 - X) N - X) (X3 - X)
hence (5) also may be written
(7) (4) - X1I) . (1 - X2I) (4 - X31) = 0,
in which the dyadic factors are commutative.
From equations (4), we find
C=I, B=4) -q'1I, A=4)2-914) +92I;
hence, from (2),
(8)
q,. = 4)2
(,P1 - X)4) + (V2 - (PI X + X2)I.
When X = X1, a characteristic number,
91 - X1 = X2 + X3, 'P2 - 'P1X1 + X1 = X2X3,
and (8) becomes
(9) (4) - a1I)a = (4) - X21) (4) - X3I).
Although every dyadic 4) satisfies the cubic (5), 4) will satisfy
an equation of lower degree when `"a = (4) - XI)a vanishes for a
characteristic number X j. The equation of lowest degree satisfied
by 4) is called its minimum equation. *
If 4, = 4) - XI = 0 for a characteristic number X1 (then also
Ta = 0), the minimum equation is linear, namely,
(10) 4) -X1I=0.
When 4) = X11, 4, = (A1 - X)I, and 03 = f(A) = (X1 - X)3; the
Hamilton-Cayley Equation is therefore
(11) (4) - X11)3 = 0.
* For its formation and properties see Maeduflee, C. C., An Introduction to
Abstract Algebra, New York, 1940, pp. 224-6. This treatment for n X n
matrices also applies to dyadics, regarded as 3 X 3 matrices.
162 LINEAR VECTOR FUNCTIONS § 74

If 4, does not vanish for any characteristic number, but *a, = 0


when X = X1, we see from (9) that the minimum equation is the
quadratic,
(12) (4, - X21) (,b - X31) = 0.
When Fd = 0, 4'2 = 0, and, consequently, 4, is linear (theorem,
§ 69) ; hence we may write 4) = AiI + uv. If we take u = i,
v = ai + /3j ± 'Yk, I = ii + jj + kk, we have
-A)I+aIi+flij+yik,
and the determinant of 'F's matrix is
¢3 = f(A) = (A1 - X)2(X1 + a - A).
Thus the characteristic numbers of b are X1, A2 = A1, A3 = Al + a,
and its Hamilton-Cayley Equation is
(13) ((D - A1I)2 . (4, - A3I) = 0.
From 'I = A1I + uv, we see that all vectors perpendicular to v
have the multiplier A1i whereas vectors parallel to u have the
multiplier Xi + u v = Al + a. When u v = a = 0, (13) re-
duces to (11).
We note that, in every case, the minimum equation and the
Hamilton-Cayley Equation have the same linear factors and differ
only in their degree of multiplicity.
74. Normal Form of the General Dyadic. Every complete dy-
adic transforms at least one set of mutually orthogonal directions
(its principal directions) into another set of the same kind.
To find the principal directions of the complete dyadic -1,, con-
sider the dyadic (D. The latter is complete (§ 65) and sym-
metric; for, from (65.8),
(c. 4)). =
We therefore may write (§ 72)
'
(1) (DC = A1u + A2jj + A3kk, (Ai 0).
Consequently,
i.cc.4).j = j.4)a.(D .k = k.4)a.4) .i = 0,
or
(-t i) =0.
§ 74 NORMAL FORM OF THE GENERAL DYADIC 163

The vectors 4 i, 4) j, 4 k are thus mutually orthogonal, and


hence i, j, k give a set of principal directions of 4). If we write
(2) 4 i = ai', 4, j = #j', 4, . k = yk',
where i', j', k' is a second dextral set of unit vectors, we have,
from (63.11),
(3) 4) = ai'i + (3j'j + yk'k.
Moreover, we always can arrange so that a, fl, y have the same
sign. If, for example, a and 0 have one sign, -y the opposite, we
can replace i', j' by -i', -j', and the set -i', -j', k' still will
be dextral.
From (3) and
,Pc = aii' + 3jj' + ykk',
we have, by direct multiplication,
(4) 4rD' 4) = a2ii + a2jj + y2kk,
(5) = a2i'i' + 132j'j' + y2k'k'.
These symmetric dyadics, which in general are different, have the
same multipliers, evidently all positive.
We have therefore proved the
THEOREM. If (P is complete, any three invariant directions of
,I)c 4) that are mutually orthogonal are principal directions of (D, and
conversely. The principal directions of 4) transform into invariant
directions of 4' 4,. Any complete dyadic 4) can be reduced to the
normal form (3) in which the scalars a, (3, y are square roots of the
multipliers of (D, 4) (or 4) (D,) having the same sign.
Example. Homogeneous Strain. In distinction to the ideal rigid body, the
particles of a deformable body are capable of displacements relative to one
another. The totality of such relative displacements is said to constitute its
state of strain.
Suppose that a particle at P moves to P' under the strain; then r' = OP'
is a continuous function of r = OP. The simplest type of strain occurs when
r' is a constant linear vector function of r,
(6) r' = 4' r (4 complete);
the strain is then said to be homogeneous. We have seen in § 62 and § 70 that
a homogeneous strain transforms lines into lines and planes into planes; and
164 LINEAR VECTOR FUNCTIONS § 75

evidently parallelism is preserved. Moreover all volumes are altered in the


constant ratio of (P3/1.
Since 4' is complete, r = 4-1 r'. The particles originally on a sphere
about 0 are displaced so as to lie upon an ellipsoid; for r r = a2 transforms
into

When 4' is reduced to the form (3),

4'-1 +jj'+ kk',


1

12

a i'i' + j'j' + 2
ti
k'k' ,

and the foregoing strain ellipsoid (with a = 1) has the equation,


x'2 ,2 z,2

cr2
+ -2 + = 1.
r2

The principal directions i, j, k of 4) are called the principal axes of strain;


they transform into ad', /3j' 7k', the principal semiaxes of the strain ellipsoid.
75. Rotations and Reflections. In order that a dyadic 4 trans-
form all vectors so that their lengths are unchanged, it is necessary and
sufficient that its inverse be equal to its conjugate:
(1)
-1 = b'.
Proof. If, for any vector r,
(2) (4) r)
then
r 4)c cb r=r I r, r -I) r= 0,
and, since I 4) - I is symmetric, it must be zero:
4), 4' = I, or 4), _ -1.
Conversely, if (1) is fulfilled,
(1
A dyadic that preserves the lengths of vectors also preserves the
angles between them: for, by virtue of (1),
(4) r)(4) - s) =r I I

and, since lengths are unaltered,


cos (c r, 4) s) = cos (r, s).
§ 75 ROTATIONS AND REFLECTIONS 165

Condition (1), although sufficient to ensure preservation of angles,


is by no means necessary. Thus the dyadic XI preserves angles
but multiplies all lengths by X.
Since fi preserves lengths and angles, any orthogonal set of unit
vectors is transformed into another such set. Thus there are two
possible cases: the dextral set i, j, k is transformed into another
dextral set i.', j', k', or into a sinistral set i', j', -k'. Hence we
have two types of length-preserving dyadics:
(3) fi = i'i + j'j ± k'k.
Their third scalar is V3 = =L1. Moreover,
fit = ±i'i ± j'j + k'k = -fi.
This shows that the vector invariant of fit is f+; but, from
(69.12), this vector is fi Equating these values, we obtain
(4) fi + _+;
the vector invariant of fi gives an invariant direction of multiplier,
'P3 = f 1. If we choose k in this direction, (3) becomes
(5) fi = Vi + j'j + p3kk, IP3 = f 1.
When V3 = 1, fi r transforms i, j, k into i', j', k. But a rota-
tion 0 about k as axis, through an angle X such that i, j revolve
into i', j', transforms i, j, k in the same way; and, since a is linear
vector function, fi = e (§ 60). Thus
(6) 0=i'i+j'j+kk
is a rotation about the axis k through an angle X determined by
its scalar and vector invariants:
(7) 01 = 1 + 2 cos X, 0 = -2 sin X k.
When P3 = -1,
fi = i'i + j'j - kk = (i'i + j'j + kk) (ii + jj - kk).
The dyadic in the first factor is the rotation e. The dyadic
ii + jj - kk in the second factor transforms i, j, k into i, j, -k.
But a reflection E in the plane of i, j transforms i, j, k in the same
way; and, since E is a linear vector function,
(8) E = ii + jj - kk = I - 2kk.
Thus, when P3 = -1, fi = 0 E.
166 LINEAR VECTOR FUNCTIONS 76

THEOREM. A dyadic 4) that preserves lengths is a rotation e when


(p3 = 1, and a rotation O followed by a reflection in the plane per-
pendicular to its axis when c03 = -1. In the latter case n E will
reduce to Z, a pure reflection, when O = I.
76. Basic Dyads. If we express both antecedents and conse-
quents of a dyadic 4) in terms of a given basis e1, e2, e3, we obtain
upon expansion 3 X 3 = 9 types of basic dyads eie; (i, j = 1, 2, 3).
On collecting terms we may write 4:
(p12e1e2 P13ele3
(1) 4) _ (p11e1e, + + +
P23e2e3
c'21e2e,
+ cp22e2e2 + +
032 (p33e3e3.
w e3e, + e3e2 +

The nine coefficients 9 are called the contravariant components of


4) relative to the basis ei (cf. § 23). If we drop the nine basic
dyads e;e; in (1), we can represent 4) by the 3 X 3 matrix,
11 1p12 13
(P V
= p21 (p22 (p23

031 032 (p33

a skeleton of numbers arranged in a definite order, which stands


for the full expression (1). This is analogous to the use of a
number triple (u', u2, u3), or 1 X 3 matrix, to represent the vector
a = ule, + u2e2 + u3e3.
If we express the vectors of 4) in terms of the reciprocal basis
e we write
(2) 4)

The nine numbers pij are called the covariant components of (b


relative to the basis ei. Just as before, we may represent 4) by a
matrix of the components cpjj. This corresponds to the use of the
number triple (ul, u2, u3) to represent the vector u = u,e1 +
u2e2 + u3e3.
But with the same basis e1, we can represent 4) in two other
ways. First, we may express the antecedents of 4) in terms of ei,
the consequents in terms of e'; the basic dyads are then eie', and
(3) c _ Z2;oz j eie'.
3 77 NONION FORM 167

Or we may express the antecedents in terms of ei and consequents


in terms of e;, so that the basic dyads are eie;; then
(4) = XMv 5eie;.
If we represent 1 by
The components gyp';, cpti' are called mixed.
matrices of mixed components, we must indicate the order of the
subscripts as shown in the preceding notation; for, in general,
e. p, i However, if we use the full notations (3) or (4), the
components can be written gyp;; for the order of the indices then is
shown by the base vectors.
77. Nonion Form. When the self-reciprocal orthogonal set
i, j, k is used as a basis, all four representations of 4) given in § 76
become the same. Since upper and lower indices no longer are
needed, we write the orthogonal components of D arbitrarily as
Vij. When no basis is indicated, the components of the matrix
(<pij) shall be regarded as orthogonal, and 4) itself is said to be in
its nonion form:
'P11 'P12 'P13

(1) _ 'P11ii + 'P121j + + 'P33kk = rP21 'P22 923

'P31 'P32 'P33

The conjugate of I corresponds to the transpose of this matrix:


'11 V21 'P31

(2) 'P12 'P22 'P32

'P13 'P23 'P33

Hence c is symmetric when 'Pij = 'Pji, antisymmetric when 'Pij =


c° ('i1 = 0).
The first scalar and vector invariant of P are readily computed:
(3) 'P1 = 911 + V22 + 'P33,
(4) = ('P23 - 'P32)1 + ('P31 - VP13)i + ('P12 - (p21)k
In order to compute 'P2 and <P3i write 4 in three-term form:
(5) = i(<P11i + 'P12j + (P13k) +

j (921i + 'P22i + <P23k) +

k((P311 + 'P32j + p033k)


168 LINEAR VECTOR FUNCTIONS § 77

Then, since [ijk] = 1, <P3 is the box product of three consequents,


namely the determinant of matrix ():
'P11 'P12 'P13

(6) <P3= 'P21 'P22 'P23

'P31 'P32 'P33

Thus is singular when the determinant of its matrix vanishes.


As to 4)2i the terms with antecedent i = j X k have as consequents,
(11i + (p1) + (D13k,
('P211 + 'P223 + 'P23k) x (cP311 + <P32i + 'P33k) =

where 'ij denotes the cofactor t of 9i j in the determinant (6) ;


hence
,11 ,12 ,13

4)21 ,22 X23


(7) 2
X31 ,32 ,33

(D11 X22
(8) 'P2 = + + X33,

and (% is the transpose of (7). Moreover, since (D-1 = 4)a/cp3


(70.5),
11 'P21 4031
(P

-1 = P22 032
(9)
'P13
('P 12
'23 033

where (P''j = (I)ij/(p3 is the reduced cofactor of 'ij. Making use of


the well-known relations in determinant theory,
'Pi1'Pj1 + <Pi2'Pj2 + 'P1i'P1j +
<P2i(P2j
+ ati

we may verify that 4-' or 4)-1 4) give the idemfactor:


1 0 0

(10) I= 0 1 0
0 0 1

The characteristic equation of 4) is obtained by equating the


third scalar of 4' - XI to zero; hence, with 4) in nonion form, it
t The cofactor of vii in the determinant I pif I is defined as the coefficient
of Pii in the expansion of the determinant; it equals the minor obtained by
striking out the ith row and jth column with the sign (-1)'+i affixed.
§ 78 MATRIC ALGEBRA 169

becomes
(P11 - X 'P12 'P13

(11) f(X) = 'P21 'P22 - X 'P23 = 0.

(P31 <P32 IP33 - X


Example. If the dyadic 4) in (1) is symmetric, and r = xi + yj + k, then
(i) r$ r = Pllx2 + 2co12xy + P22y2 + 2w13x + 2-P23Y + V33 = 0
represents a conic section-an ellipse, parabola, or hyperbola, according as
V12 - IPiisv2z < 1, = 1, or >1. Let this conic cut the sides BC, CA, AB of
a triangle ABC in the points R1, Ri; R2, R2; R3, R3, respectively, and let
the corresponding division ratios be Plr p l; P2r P2i Par P3'- If we put r =
(b + Pc)/(1 + p) in (i) we find that P1, P1 are the roots of the quadratic
equation
b 4) b +2Pb 4) - C +P2c (D c = 0;
hence

and
c'4) 'c a'4 a
b
follow in the same way. On multiplying these equations, we have
(ii) P1P1P2P2P3P3 = 1

Now let R2'R3i R3R1, R'R2 meet BC, CA, AB in the points S1, S2, S3 which
divide the respective sides in the ratios al, 02, 03. Then by the Theorem of
Menelaus (§ 7, ex. 2)
P2P3°1 = - 1r P3P1Q2 = - 1r P1P20'3 = - 1

On multiplying these equations together and making use of (ii), we have


0.10.20.3 = -1; hence S1, S2,83 are collinear. We thus have proved Pascal's
Theorem: $ The opposite sides of a hexagon (R1R'1R2RZR3R3) inscribed in a conic
meet in three collinear points (Si, S2, S3)-
78. Matric Algebra. The sum of two dyadics A + B in nonion
form obviously is obtained by adding corresponding elements of
their matrices. As to the product C = A B, let us consider the
formation of a certain dyad of C, say c12ij. This evidently results
from the product of terms in A with antecedent i and terms in B
with consequent j: thus
C12 = a11b12 + a12b22 + a13b32 = 2;alb,2-
r

The general result is therefore


Cij = 2;airbrj ;
r
$ This proof is due to Wedderburn, Am. Math. Monthly, vol. 52, 1945, p.
383.
170 LINEAR VECTOR FUNCTIONS § 78

the element in the ith row and jth column of A B is the sum of the
products of the elements in the ith row of A by the corresponding
elements in the jth column of B-the "row-column rule."
The foregoing rules for the sum and product of dyadics in non-
ion form are precisely the classic definitions for the sum and prod-
uct of square matrices. These definitions may be extended to rec-
tangular matrices with m rows and n columns (m X n matrices).
1. The sum A + B of two m X n matrices is them X n matrix
obtained by adding their corresponding elements.
2. The product of an m X p matrix A and a p X n matrix B is
the m X n matrix C = AB, whose element in the ith row and jth
column is
P
(1) cjj = E airbrj.
r=1

Note that only similar matrices can be added; whereas in a


product the second matrix must have the same number of rows
as the first has columns.
This extension enables us to interpret scalar products of dyadics
and vectors in terms of matric algebra. We regard a vector u
with rectangular components ui either as 1 X 3 matrix (row vector)
or a 3 X 1 matrix (column vector). Since a dyadic A in nonion
form is a 3 X 3 matrix, u must be a row vector in u A, a column
vector in A u. With this proviso, the rules of matric algebra
give values of vector components in full agreement with vector
algebra:
(2) v = u A, vi = u1a1i + u2a2i + u3a3i;
(3) w = A u, wi = ai1u1 + ai2u2 + ai3u3
The matric product of a row vector u into a column vector v is
a 1 X 1 matrix consisting of a single element, the scalar product:
(4) u v = u1v1 + u2V2 + u3v3
But the matric product vu of a column vector into a row vector
is a 3 X 3 matrix, namely the dyad,
v1u1 v1u2 v1u3
(5) Vu = v2u1 v2u2 v2u3
v3u1 v3u2 V3u3
§ 79 DIFFERENTIATION OF DYADICS 171

in nonion form. Matric multiplication, although associative and dis-


tributive with respect to addition, is in general not commutative. The
proofs follow readily from (1).
79. Differentiation of Dyadics. If a dyadic (F is a function of a
scalar variable t, we define
d(= (F(t
+ At) - (F(t)
lim
dt it -o At

For a single dyad ab, let a and b become a + Aa, b + Ab when t


becomes t + At; then, since
(a + Aa) (b + Ab) - ab Ab Aa Aa
At
= a-+-b+-Ab,
At At At

we have, on passing to the limit At - 0,


d db da
(1)
- (ab) =a dt + dt
b.

This formula suggests the usual product rule with the order of the
factors preserved.
The derivative of the dyadic (D = 2;aibi is evidently the sum of
the derivatives of its dyads. If 4, is given in the nonion form
(77.1))
'P12
d(
(2) 'P22
dt
'P32

where the primes denote derivatives with respect to t. Note that


the derivative of <03, the determinant of the matrix ('i;), is not
the determinant of the matrix
The derivatives of products such as - r, fi x r, s (- r, which
conform to the distributive law, are computed just as in the cal-
culus when the order of the factors is preserved. For example,
d d(F dr
r+(D

(3) dt dt;
d ds d(F dr
(4)
dt dt dt dt
172 LINEAR VECTOR FUNCTIONS § 81

80. Triadics. A triadic is defined as a sum of triads, Eaib,ci.


A triad abc consists of three vectors written in a definite order.
We may regard a triadic as an operator which converts vectors r
into dyadics; thus
4) r = Eaibici r, r 4) = 2;r aibici.
If 4) and' are two triadics, we write - = T, when
(1) 4) r = 4, r for every vector r.
Then, for any vector s, s (D r=s 4, r, and, from the defini-
tion of equality for dyadics (61.3),
(2) s 4) = s - ' for every vector s.

Conversely, from (2) we may deduce (1). Thus from either (1) or
(2) we may conclude that 4) = T.
Using a given basis ei and its reciprocal ei to form basic dyads,
we have seen in § 76 that the 32 = 9 components of a dyadic are
of 22 = 4 types. For a triadic 4), the 33 = 27 components are of
23 = 8 types; for, for each vector in the basic, triads may be
chosen from the set ei or e' giving 23 types; and, for a given type,
each index on the base vectors may be chosen in three ways, giving
33 components.
Similarly, we define a tetradic as the sum of tetrads Maibicidi.
Two tetradics 4), I are equal when either (1) or (2) holds good.
A tetradic has 34 = 81 components of 24 = 16 types.
We shall speak of scalars, vectors, dyadics, triadics, collec-
tively as tensors of valence 0, 1, 2, 3, The equality of two
tensors of valence n, say 4) = 'Y, depends upon the equality of
two tensors of valence n - 1, as required by equations (1) or (2).
81. Summary: Dyadic Algebra. A linear vector function f(r) is
characterized by the properties,
f(a + b) = f(a) + f(b), f(Xa) = Xf(a).

A dyadic 4) = X.aibi is the sum of dyads aibi; the vectors ai are


antecedents, bi are consequents. The conjugate of 4) is 4 = 2;biai.
Any linear vector function may be expressed as 4 r (or r
and, if [ala2a3] 0 and f (ai) = bi (i = 1, 2, 3),
f(r) _ 4 r, where 4) = blal + b2a2 + b3a3.
81 SUMMARY: DYADIC ALGEBRA 173

Basic definitions (r an arbitrary vector) :


NP : 4)

(D r0;
(+'1')r=
r = (I' r);
I (idemfactor) : I r = r;
-i (reciprocal) : c (I)-1
= 4'-1 4 = I;
4) xv: Taibixv if 4) = Eaibi;
V x (D: My x aibi.
Every dyadic may be reduced to the sum of three dyads,
4) =al+bm+cn,
in which either antecedents or consequents may be an arbitrary
non-coplanar set. In this form, the principal invariants of 4) are:
Dyadic: 4, = bxcmxn+cxanxl+axblxm = (D2;

Vector: = axl+bxm+cxn,
*= (b x c) x (m x n) + cycl
Scalar: (P1 =
'P2 = `I'1 = (b x c) (m x n) + cycl,
(P3 = [abc][lmn],
4PO

4) is complete if 'P3 0 0; only complete dyadics have reciprocals.


4) is planar if <P3 = 0, 4)2 0 0; a planar dyadic may be reduced to
two dyads. (D is linear if '1)2 = 0, 4) 0 0; a linear dyadic may be
reduced to one dyad. Planar and linear dyadics are called singular.
Fundamental identities :
p), = *' 4)"
(4, ((D ,F)-1 = j-1 4)-1;

Ixv = vxl, (Ixv)c = -Ixv, I x(uxv) = vu -uv;


(Da = (Da `I' = (P3I (the adjoint 4)a = (h,);
1431 - 'P24) + tP1 (D2 - 4)3 = 0 (Hamilton-Cayley Equation).
174 LINEAR VECTOR FUNCTIONS
4) is symmetric if 4), = 4), antisymmetric if 4), = - 4). Every

and antisymmetric dyadic; the latter is - 2I x .


dyadic 4) can be expressed uniquely as the sum of a symmetric

The vector r1 is an invariant direction of - with multiplier Al if


4) r1 = A1r1. The multipliers of 4) satisfy the cubic,

503 - <02X +
P1A2
- A3 = 0 (characteristic equation).
The invariant direction rl makes (4) - X11) r1 = 0.
If 4) is symmetric its multipliers Ai are all real; and, if Al A2i
rl 1 r2. A symmetric 4) always may be reduced to the form,
`1) = A1u + A2JJ + X3kk;
the multipliers Ai need not be distinct.
Any dyadic 4) can be reduced to the normal form,
(P = ai'i + $j'j + yk'k,
in which a, 0, y all have the same sign; i, j, k and i', j', k' are two
dextral sets of orthogonal unit vectors.
When - is complete, the point transformation, r' = 4) r,
changes lines into lines and planes into planes, preserves paral-
lelism, and alters all volumes in the ratio 'P3/1. It preserves
lengths when and only when 4-1 = 4,; it is then a rotation if
03 = 1, a rotation followed by a reflection if p3 = -1.

PROBLEMS
1. Prove that
abxc +bcxa +caxb = [abc)I.
2. For any dyadic 4> show that
U.4> v - v 4> u =4. uxv.
3. If 4> has the characteristic numbers X1, X2, X3 and the corresponding in-
variant directions r1i r2, r3, prove that 4)" (n an integer) has the characteristic
numbers X', 02, 03 corresponding to the salve invariant directions.
4. If 4> has the characteristic numbers X1, X2, 1\3 for the directions r1r r2, r3,
prove that 4>2 has the characteristic numbers X2X3, X3X1, X1X2 corresponding to
the directions r2 x r3i r3 x r1, r1 x r2. [Cf. (70.2).]
5. If 4, = 4,, prove that *2 = 4>2,,, + = -4 and h = cir >k2 = v2, 4,3 = v3
6. Prove that the scalar invariants of 4' = 4>2 are
01 = 'c1 - 2'P2, 2 = 'P2 - 2'vl'P3, '3 = 'P3
[Use Prob. 3 and (71.4).]
PROBLEMS 17 5

7. Given the invariants c2, +, , , v3 of 4), find the corresponding invar-


iants for
(a) k4), (b) P x u, (c) 4-1.
8. Compute (P2, +, and the six scalar invariants of (69.13), namely,
'Pl, V2, 4'3; + ' +, + ' 4) ' +, + ' p2 - +,
for the dyadic
1 3 -2
= 2 0 4
-1 2 3

9. If D = I x e and e is a unit vector, prove that


4)2
= - (I - ee), 4'3 = -4), D4 = I - ee, 4)5 = (P.
10. Show that the symmetric part of (P, namely, 4' = 1(4) +4,), has the
scalar invariants
h = c1, 42 = V2 - 4+' 11
413 = V3 4 1
11. Prove that ('t - ')2 = `1'2 *2-
12. Prove that the first three scalar invariants of 4 and 4 4) are the
same.
m n.

13. Ifs a;b,, 4' c;d1,


1=1 J=1
we define the double-dot product 'F:4' as the scalar
m n

E(a`.c1)(b`-d,).
1-I j_1
Prove that
(a)

(b) (uv) :4' = u 4)


++4':
v, 4 ) :1

(c) If 4) is given in the nonion form (77.1),


'P11 + V12 +'P13 +'P2I + P22 + V23 +
'P21
+'P32
+P2
Hence 4, = 0 when and only when 4):4' = 0.
14. Any central quadric surface with its center at the origin has an equa-
tion of the form,
(1) r- 'F r = 1,
where 4, is a symmetric dyadic; for, if 4) is reduced to the standard form
(72.4), we have
axe + $y2 + yz2 = 1.
This represents an ellipsoid, an hyperboloid of one sheet, or an hyperboloid
of two sheets, according as a, j3. y include no, one, or two negative constants.
176 LINEAR VECTOR FUNCTIONS
If p 5*1 0, the equation (1) associates with every point (1, p) the plane
(4' p, 1), its polar plane. Prove that:
(a) If the point (1, p) lies on the quadric surface, its polar plane (4' p, 1)
is tangent to the quadric at (1, p). [Find the points where the line r = p + Xe
cuts the quadric when e 4' p = 0.]
(b) If the polar plane of (1, p) passes through (1, q), the polar plane of
(1, q) passes through (1, p).
15. The diametral plane of any point (1, p) on the quadric surface (1) is
the locus of the mid-points of all chords parallel to p. Show that the diam-
etral plane of (1, p) is (4). p, 0).
16. If three points (1, u), (1, v), (1, w) on the ellipsoid r 4) r = 1, satisfy
the equations,
U.
the position vectors u, v, w are said to form a conjugate set. Show that:
(a) The vectors u, v, w and 4' - u, 4) v, 4) w form reciprocal sets.
(b) 4'-1 = uu + vv + ww.
(c) For any conjugate set u, v, w, of the ellipsoid r (D r = 1, the sum
u u + v- v + w- w and the product u x v- w are constant.
17. Verify by direct computation that the dyadic 4' in Problem 8 satisfies
its Hamilton-Cayley Equation.
18. A rigid body with one point 0 fixed has the inertia dyadic K relative
to 0 (§ 72, ex. 2). If the angular velocity of the body at any instant is w,
show that its moment of momentum H (defined as f r x v dm) and kinetic
energy T (defined as v dm) are given by
if v

19. If the forces acting on the body of Problem 18 have the moment sum
M about 0, the equation of motion is dH/dt = M. Show, from (56.3), that
dH
dt
=K dw
dt
+wxK w.

Let K = Aii + Bjj + Ckk (72.10) when referred to the principal axes of
inertia (fixed in the body). Then if w = [WI, w2i w3], M = [Ml, M2, M3] re-
ferred to these axes, deduce Euler's Equations of Motion:
Awl - (B - C)w2w3 = M1,
Bm2 - (C - A )w3w1 = M2,
(A - B)w1w2 = M3-
20. In Problem 19 show that d7'/dt = M w (the energy equation).
21. In Problem 18, suppose that the only forces acting on the body are its
weight W and the reaction R at the support 0; then if the center of mass is at 0,
W passes through 0 and M = 0. Prove in turn that
(a) H is a constant vector.
(b) T is a constant scalar.
PROBLEMS 17 7

(C) If w = OP, the locus of P in space is the invariable plane w H = 2T;


:uul the locus of P in the body is the energy ellipsoid w K w = 27'.
(d) The energy ellipsoid is always tangent to the invariable plane at P.
(c) The body moves so that its energy ellipsoid rolls without slipping on
the invariable plane (Poinsot's Theorem). [See Brand's Vectorial Mechanics,
§ 219.]
22. The vectors of a dyadic 'F are fixed in a rigid body having the instan-
taneous angular velocity w, relative to a "fixed" frame. Show that, relative
to this frame,
d'F/dt=ca x4) -4) xw.
23. A rigid body revolving about its fixed point 0 has the angular velocity
w = [wl, W2, w3] referred to fixed rectangular axes through O. If I1, 12, I3 are
the moments of inertia about the fixed axes x, y, z and 123, I31, 112 are the
products of inertia for yz, zx, xy, show that
dIl/dt = 2(I12w3 - 113-2), .. ,

dI23/dt = (13 - 12)wl + 113W3 - 112-2, ... ,


[If K is the inertia dyadic of the body relative to 0, Il = i K i, 123 =
-j K K. k, etc. See § 72, ex. 2.]
24. The axis of a homogeneous solid of revolution has the direction of the
unit vector e. If its principal moments of inertia at the mass center G are
A, A, C, show that the inertia dyadic at G is
KG =AI+(C-A)ee.
Hence find the moments and products of inertia with respect to fixed rectan-
gular axes x, y, z through G if e = [1, in, n].
25. If G is the mass center of a rigid body of mass in and r* = OG, show
that the inertia dyadic at 0 is
Ko = m(r* r* I - r*r*) + KG.
Hence compare moments and products of inertia for parallel axes at 0 and G.
26. If X is arbitrary parameter < a2 but b2 or c2, the dyadic,
c2kk
x211 + b2 J1 { (a > b > c),

defines a one-parameter family of confocal quadric surfaces r % r = 1.


These are ellipsoids if X < c2, hyperboloids of one sheet if c2 < \ < b2, hyper-
bolas of two sheets if b2 < X < a2.
Prove that
(a) The central quadrics r T r = 1 and r O r = 1 are confocal when
and only when I-' - 0-1 = kI.
(b) Two confocal central quadrics of different species intersect at right
angles. [O - 4, = ke *.]
CHAPTER V

DIFFERENTIAL INVARIANTS

82. Gradient of a Scalar. The points P of a certain region may


be specified by giving their position vectors r = OP; and we shall
on occasion refer to P as the "point r." A scalar, vector, or dyadic
which is uniquely defined at every point P of a certain region is
called a point function in this region and will be denoted by f (r).
For example, the temperature and velocity of a fluid at the points
of a three-dimensional region are scalar and vector point functions
respectively.
A scalar point function f (r) is said to be continuous at a point
P1 if to each positive number e, arbitrarily small, there corresponds
a positive number S such that
If(r) - f(r1) <e when jr - r1 l <6;
then, as r approaches r1 in any manner, limf(r) = f(r1).
From a point P1 draw a ray in the direction of the unit vector,
e =icosa+jcos/3+kcosy.
Along this ray r = r1 + se, where s denotes the distance P1P, and
f (r) is a function of s. We now define
f (r, + se) - f(r1)
(1) of = lim
ds s -.o s

as the directional derivative of f (r) at P, in the direction e. If this


limit exists on all rays issuing from P1i f (r) is said to be diferen-
tiable at P1.
The rectangular coordinates of any point P on the ray r =
r1 -}- se (s > 0) are
x = x1 + s cos a, y = y1 -}- s cos z = z1 -{- s cos y.
178
82 GRADIENT OF A SCALAR 179

If f(r) is given as a function f(x, y, z),


df of dx of dy of dz
(2) -- _ - - --F - - + -- --
ds ax ds ay ds az ds

af of of
= cos a + cos i3 + cos y;
ax ay az
or, since
e i = cos a, e j = cos a, e - k = cosy,
=
(3) e
A \1 ax+iafy +kaz/
The vector in parenthesis is called the gradient of f (r) and is writ-
ten grad f or
49 19
(4) Vf
ax+jay+kc3z
Vf is a vector point function; thus, when (3) is written as
df
(5)
ds
the direction enters only through the factor e. The directional
derivative of a scalar function at a point P is the component of its
gradient at P in the given direction. The gradient Vf at P effects a
synthesis of all the directional derivatives of f at P. In effect, the
vector of replaces the infinity of scalars df/ds.
When P varies along a curve tangent to e at P1, f(r) is a func-
tion of the are s = P1P along the curve, and df/ds is still given
by (2) ; for, at P1, dr/ds = e (44.1), and hence
dx/ds = cos a, dy/ds = cos a, dz/ds = cos y.

Since df/ds, as defined by (1), does not depend upon any specific
choice of coordinates, (5) shows that the gradient Vf has the same
property. In fact, we determine Vf by giving the directional de-
rivatives df /dsi in three non-coplanar directions e1. For, since
ei Of = df/dsi are the covariant components of Vf, we have,
from (24.2),
df
(6) of = el + e2 df + e3 df .

ds1 ds:2 ds3


180 DIFFERENTIAL INVARIANTS § 82

We proceed to specify the length and direction of Vf independ-


ently of the coordinate system. The points for which f has a con-
stant value lie on a level surface of f. In any direction e tangent
to the level surface at P, df/ds = e - Vf = 0; hence Vf is normal
to the level surface at P. If n is a unit vector normal to the level
surface and directed towards increasing values of f, n Vf > 0.
Hence Vf has the direction of n, and its magnitude is the value of
df,/ds in this direction. Writing this normal derivative df/dn, we
have

(7)

For example, if f = r, the distance OP from the origin, the level


surfaces are spheres about 0 as center, and n = R, the unit radial
vector; the normal derivative dr/dn = 1, and
(8) Vr=R.
When f is a function f (x, y, z) of rectangular coordinates, Vf is
given by (4). In particular,

(9) Vx=i, Vy = j, Vz = k.
If f is a function f(u, v, w) of variables which themselves are func-
tions of x, y, z, we have
df of du of dv of dw
ds au ds av ds aw ds
or, in view of (5),

Vu Of
+Vvaf +Vwa-f1
\ au av aw

As this equation holds for every e,


of af Of .
(10) Vf = Vu + Vv + Vw
Ou av aw

When f = f (u, v) or f = f (u),


Of Of
Vf = Vu of + Vv , Vf = Du ,
au av au
§ 83 C1t.ADIENT OF A VECTOR 181

respectively; for example,


1
V(uv) = v Vu + it Vv, Vu' = nu"-1 Vu, V log u = - Vu.
It

When f is constant, Vf = 0; conversely, Vf = 0 implies of/ax =


of/ay = of/az = 0, and f is constant.
If f (r) is differentiable in a region R, Vf is defined at all points
of R. If moreover Vf is continuous in R, we say that f (r) is con-
tinuously differentiable in R.
Example. Gradients in a Plane. The gradient of a point function f(x, y)
in the xy-plane is

vf=afi+Of
ax ay
j.

At any point P, Vf is normal to the level curve f(x, y) = c through P.


For a function f(r, 0) of plane polar coordinates,
of of-
= or
Cf R +00r'
in the notation of § 44; for Vr = R and V0 = P/r (from (44.5) dB/ds = 1/r in
the direction perpendicular to R).
For a function f(rl, r2) of bipolar coordinates,

Vf =OrlR1+_R2,

where R1, R2 are unit radial vectors from 01, 02 directed to the point in ques-
tion.
The families of curves u = c, v = c cut at right angles when Cu Cv = 0.
For example, the ellipses rl + r2 = c and hyperbolas r, - r2 = c cut orthogo-
nally since (R1 + R2) (R1 - R2) = 0-

83. Gradient of a Vector. Let f(r) denote a vector point func-


tion in a certain region. It is said to be continuous at a point P1
if to each e > 0 there corresponds a S > 0 such that
I f (r) - f(r1) I < e when Ir - r1 I< S.
If f(r) is given by its rectangular components,
f(r) = f1i + f2j + f3k,
f(r) is continuous at P1 when the three scalar point functions
fi, f2, f3 are continuous there.
182 DIFFERENTIAL INVARIANTS § 83

Let P1P be a ray drawn from P1 in the direction of the unit


vector,
e = icosa+jcos$+kcosy.
Along the ray r = r1 + so (s > 0) and f(r) is a function of s. We
now define
df f(rl + se) - f(r1)
-- = lim
ds s-.o s
as the directional derivative of f (r) at P1 in the direction e. If this
limit exists on all rays issuing from P1, f(r) is said to be differen-
tiable at Pl. Evidently f (r) is differentiable if its rectangular com-
ponents f, are differentiable.
If f(r) is given as a function f(x, y, z) of rectangular coordinates,
we have. just as in § 82,
df of dx of dy of dz
(2)
ds axds+ayds+azds
of of of
=- cosa +
ay
cosa+
- cosy.

On replacing the cosines by e i, e j, e k,


df of of of
(3) ds = e1-+ja +kaz
y ,

a formula entirely analogous to (82.3). The dyadic in parenthesis


is called the gradient of f(r) and is written grad f or
of
(4) Vf = ofi-+ j - +k --
of
ax ay az
Vf is a dyadic point function; thus, when (3) is written
df
(5)

the direction enters only through the prefactor e. The gradient


Vf at P effects a synthesis of all the directional derivatives of f
at P. In effect, the dyadic Vf replaces the infinity of vectors df/ds.
Since df/ds, as defined by (1), does not depend upon any specific
choice of coordinates, (5) shows that Vf has the same property.
§ 84 DIVERGENCE AND ROTATION
In fact Vf is determined by giving the directional derivatives
df/dsi in three non-coplanar directions ei:
df + e2 df + e3 df
(6) of _ e'
ds1 ds2 ds3

For, if we put df/dsi = ei Vf, the right member becomes


(e'ei + e2e2 + e3e3) Vf = I I. Vf = Vf.
When f is function of rectangular coordinates f (x, y, z), Vf is
given by (4). More generally, for f(u, v, w) we have
df of du of dv of dw
ds au ds av ds aw d s
C of of of
+ VV
au av aw

for any e; and, from the definition of dyadic equality,


of of of
(7) Vf = Vu-+Vv--+Vw-
au av aw

84. Divergence and Rotation. The first scalar invariant of the


gradient Vf of a vector f is called the divergence of f and is written

The vector invariant of Vf is called the rotation or curl of f and


is written V x f, rot f, or curl f.
In terms of rectangular coordinates, we therefore have the de-
fining equations:
(1) Vf=gradf=iaxof - +j ayof- +k -azof ,

(2) V f = divf -of + j


= i ax of
ay
of
az
,

of of of
3) xf = rotf = ix-+
ax
jx-+kx__.
ay az

If f is resolved into rectangular components,


f=ifl+jf2+kf3,
184 DIFFERENTIAL INVARIANTS § 84

we obtain from (1) its gradient of in nonion form with the matrix:

afl/ax af2/ax af3/ax


(4) of = afl/ay af2/ay af31ay
8fl/az af2/az af3/az
The first scalar and vector invariants of of now become
axl+ a19h
v + ,
(5) f= az

(6) vxf=i --- -f-j(az---ax -}-k(ax-- ay)/


Gay
f3 af2

az
afl af3 af2 afl

The last expression is easily remembered when written in determi-


nant form:
j k
a a a
(7) Vxf =
ax ay az

fl f2 f3

These expressions for divergence and rotation may be written down


at once, if we regard V - f and V x f as products of the vector
operator del (or nabla),

(8) V =axai-+j-+k--,
a
ay
a
az

with the vector f = if, + jf2 + kf3.


For the position vector r = xi + yj + zk, we have
Or
yr = i-+ j -Or+k --Or= ii+ jj +kk,
ax ay az

the idemfactor; hence


(9) Vr = I, div r = 3, rot r = 0.
When f = the gradient of a scalar,
(9 (P a4'
fl=ax, f2=ay , f3=az-
§ 84 DIVERGENCE AND ROTATION 185

Using these components, (5) and (6) gives


a2
(P a2V a2,p
v v'P
( 10 ) . ax2 + ay 2 + az2

(11) V X VV = 0.

The differential operator V V or div grad is called the Laplacian


and often is written
__ a2 a2 a2
(12) v2
ax2 + ay2 +az2
When f = V x g, the rotation of a vector g,
C193 092 _a91 a93 a92 C191
A
ay az f2 az ax ' f3 ax ay

In this case (5) and (6) give


(13) 0,

(14) Vx(Vxg) _ V(V.g) - O29.


Proof of (14) : From (6),
r a292 - x291 - a291 a293
+ .. .
ay ax aye az2 + az ax Jt

= i j a tag, + a92 + a931


- v291 I + .. .
lax ax ay az J
a
=i ax
(O. g) - V291 +...

= V(V . g) - V2g.
Note that, if we regard V as an actual vector and expand
V x (V x g), but keep the V's to the left of g, we obtain V(V g) -
(V V)g, the correct result.
The proofs of (11), (13), and (14) depend upon changing the
order of differentiations in mixed second derivatives; this is always
valid when the derivatives in question are continuous.
186 DIFFERENTIAL INVARIANTS § 85

The foregoing differential relations of the second order also may


be written
(10) div grad p = V2<p,
(11) rot grad cp = 0,
(13) div rot g = 0,
(14) rot rot g = grad div g - V2g.
Example. The velocity distribution in a rigid body is given by (54.4):

vP = vo -}- w x OP = vo + w x r.
Hence, from (3),
rotvp = rot(wxr) = ix(wxi) + =3w- w =2w;
the rotation of the velocity of the particles of a rigid body at any instant is equal
to twice its instantaneous angular velocity.

85. Differentiation Formulas. Starting from the defining equa-


tions (84.1), (84.2), (84.3), we now can deduce some useful iden-
tities.
If X is a scalar point function:
(1) V(Xf) = (OX)f + XOf,
(2) div (hf) = (VX) f + X div f,
(3) rot (Xf) = (VX) x f + X rot f.
For the vector point function f x g:
(4) V (f x g) = (of) x g - (Vg) x f,
(5) div (f x g) = g rot f - f rot g,
(6) rot (f x g) = g Of - f Vg + f div g - g div f.
We give the proof of (6) :
Caf ag/ +...
rot(fxg) =1x
axxg+ fxax
=g iax-- (i
\ -g+(i
ax!
axg
\ f -f.i g
ax
+. ..

= g - Vf - (div f)g + (div g)f - f . Vg.


§ 86 GRADIENT OF A TENSOR 187

For the scalar point function f g:


(7) 0(f - g) _ (of) g+ (7g) . f
From (68.6), D, = +Ix hence
(8) (of), = Vf + I x rot f,
(9) (Vf) - a = a - Vf + a x rot f.
Making use of (9), we also can write (7) as
(10)
If '1 is a constant dyadic,
(11) v(r -1)) =I.4) =4);
(12) div (r D) = cpl,
(13) rot (r 4) _+.
Finally, if 4) is a constant symmetric dyadic,
(14) r.
From this result we can prove the
THEOREM. A symmetric dyadic I transforms any vector r = OP
into a vector r' = 4) r normal to the real central quadric surface
r 4) r = f1 * at the point P; and r' = 1/p where p is the distance
from 0 to the tangent plane to the quadric at P.
Proof. From (72.4), we see that
ax2+(3y2+yz2 = f1
represents a central quadric (an ellipsoid if a, y have the same
sign). From (14), r' has the direction of n, a unit normal to the
quadric at P; hence
r'n=
r' = 1/I 1/p.
86. Gradient of a Tensor. If f(r) is a tensor point function of
valence v (cf. § 80), the derivative of f(r) at Pl in the direction
of the unit vector e is defined by
df f(rl + se) - f(rl)
(1) -- = lim
ds s -o s
* The sign is chosen so that the quadric is real.
188 DIFFERENTIAL INVARIANTS § 86

If this limit exists for all rays drawn from P1, f(r) is said to be
differentiable at P1. When referred to a constant basis, f(r) is dif-
ferentiable when all its scalar components are differentiable.
If f(r) is given as a function f(x, y, z) of rectangular coordinates,
df of dx of dy of dz
ds Ox ds+ay ds+azds
Since dx/ds = e - i, etc., we have, just as in § 83,
df
(2) = e vf,
ds
of of of of
(3) Vf=i-+j-+k-=i,.-
ax ay az axr
t
Here Vf, the gradient of f (grad f) is tensor point function of va-
lence v + 1; of effects a synthesis of all the directional derivatives
(valence v) of the tensor f. Thus, if f is dyadic, the triadic of
replaces the infinity of dyadics df/ds.
of is independent of the choice of coordinates. At any point, of
is completely determined when df/dsi is given for three non-
coplanar directions ei:
df
of = e1df- + e2 - + e3 -
df
(4)
ds1 ds2 dsi '
for, if we put df/dsi = ej Vf, the right member becomes I of
= Vf.
When f(r) is a vector, we obtain from the dyadic of the invari-
ants div f and rot f by putting dots and crosses, respectively, be-
tween the vectors of each term. Similarly, if f(r) is any tensor of
valence v > 0, we obtain from of the further invariants,

(5) =i -of +j -of +k of of


ay
ax az axr

of + j x of + k x of _ it % Of ,
(6 ) Vx f = ix
ax ay Oz axr

of valence v - 1 and v, respectively. In forming these invariants


from Vf, the dots and crosses are inserted between the first and
second vectors to the left-we dot and cross in the first position.
t Here i, j, k; x, y, z are written il, i2, i3; xl, x2, x3, and we employ the sum-
nmation convention: a repeated index (as r) denotes summation over the index
range 1, 2, 3. See § 145.
86 GRADIENT OF A TENSOR 1S9

For example consider the dyadic


f(r) = fiiii + fi2ij + ... = fstisit
where s and t are summation indices. Then if D, denotes a/8x
Vf = i,D,f = D,fst i,isit, a triadic;
V f = i, D,f = Dfst i, is it
= (Difit + D2f2t + D3f3t)it, a vector;
here i, is = 6,s, the Kronecker delta (23.2) ;
Vxf = i,>D,f = Drfsti,xisit
= (D2f3t - D3f2t)iiit + (D3fit - Dif3t)i2it
+ (Dlf2t - D2f3t)i3it, a dyadic.
Again, if X(r) is a variable scalar and 4) a constant dyadic,
(7) V(X4)) = (VX)4), a triadic;
(8) V V. (a4)) = (VX) 4), a vector;

(9) Ox (A4)) = (VX) x 4), a dyadic.


For any tensor f(r) with continuous second derivatives we have
identities which are generalizations of (10), (11), (13), (14) of § 84:
(10) V Vf = V2f,
(11) V X Vf = 0;
and, if f(r) is not a scalar,
(12)
Vx x f) _ V(V f) - V2f.
The proofs are straightforward applications of (3), (5) and (6):
V Vf = i, D,(isDsf) = i, is DrDsf = DrDrf = V2f;
V x Vf = it x D,(isDsf) = i, x is DrDsf = 0;
V (V x f) = it Dr(ls x Dsf) = i, x is DrDsf = 0;
V x (V x f) = i, x Dr(is x Dsf) = i, x (is x DrDsf)
= is i, DrDsf - i, is DrDsf
= i8D5(i, Df) - D,D,f
= V(V f) - V2f.
Since i, x is = -is x i,, DrDsf = DsD,f, all non-zero terms in
V x of and V (V x f) cancel in pairs.
190 DIFFERENTIAL INVARIANTS § 87

87. Functional Dependence.


THEOREM 1. A necessary and sufficient condition that two con-
tinuously differentiable functions u(x, y), v(x, y) satisfy identically a
functional relation f(u, v) = 0 is that their Jacobian vanish:
a(u, v) au/ax au/ay
(1) = I = 0 or Vu x Vv = 0.
a(x, y) av/a. av/ay I

Proof. The condition is necessary ; for, from f(u, v) = 0, we


have
of of of
-Vu+--Vv = 0, Vu. Vv = 0.
au av av

If it is constant, Vu = 0, and Vu x Vv = 0. If u is not constant,


f (u, v) must contain v, of/ft is not identically zero, and again
VuxVV=0.
Conversely, suppose that Vu x Vv = 0. This relation is satisfied
if either u or v is constant; thus, if u = c, we may take f (u, v) =
it - c. If u and v are not constant, Vv is parallel to Vu; hence Vv
is normal to the curves u = c. Along these curves, dv,/ds = 0, and
v is constant. In other words, a level curve it = a is also a level
curve v = b; when u is given, v is determined, and v is a function
of it.
THEOREM 2. A necessary and sufficient condition that two con-
tinuously differentiable functions u(x, y, z), v(x, y, z) satisfy identi-
cally a functional relation f (u, v) = 0 is that
i j k
(2) Vu x Vv = au/ax au/ay au./az = a.
av/ax av/ay av/az
The proof is essentially the same as in theorem 1. Instead of
level curves of u and v, we now have level surfaces.
THEOREM 3. A necessary and sufficient condition that three con-
tinuously differentiable functions u(x, y, z), v(x, y, z), w(x, y, z) sat-
isfy identically a functional relation f (u, v, w) = 0, is that their
Jacobian vanish:
au/ax au/ay au/az
a( u, v, w)
(3) = av/ax av/ay av/az = Vu x Vv Vw = 0.
a(x, y, z)
Ow/ax aw/ay aw/az
88 CURVILINEAR COORDINATES 191

Proof. The condition is necessary for, from f (u, v, w) = 0, we


have
of -of of
-Vu+
au av
Vv+ -of Vw = O,
aw aw
0.

If u and v alone satisfy a relation f (u, v) = 0, Vu x Vv = 0, and


also Vu x Vv Vw = 0. If this is not the case, f (u, v, w) must con-
tain w, of/aw is not identically zero, and again Vu x Vv Vw = 0.
Conversely, suppose that Vu x Vv Vw = 0. If Vu x Vv = 0,
f(u, v) = 0 from theorem 2. If Vu x Vv F- 0, consider the curve of
intersection of the level surfaces u = a, v = b. Vu and Ov are
normal to this curve; and, since Vu, Vv, Vw are coplanar, Vw is
also normal to the curve. Therefore dw/ds = 0 and w = c along
the curve u = a, v = b; in other words, when u and v are given,
w is determined: w is a function of u and v.
88. Curvilinear Coordinates. In a given region let
(1) u = u(x, y, z), v = v(x, y, z), w = w(x, y, z)
be three continuously differentiable functions whose Jacobian
Vu x Vv Vw F4- 0 at all points. The functions, therefore, are not
connected by a relation f (u, v, w) = 0. Since the Jacobian is con-
tinuous, its sign cannot change in the region; and, to be explicit,
we shall suppose that
(2)

This involves no loss in generality; for, if the Jacobian were nega-


tive, an interchange of v and w (for example) would make it
positive.
Under the foregoing hypotheses a well-known theorem $ states
that in the neighborhood of any point (x0, yo, zo) the equations
(1) have a unique inverse
(3) x = x(u, v, w), y = y(u, v, w), z = z(u, v, w);
and that these functions are also continuously differentiable. At
least in a suitably restricted region of uvw space, each set of values
it, v, w yields, through equations (3), a unique set x, y, z. In this
region the correspondence (x, y, z) - (u, v, w) effected by (1) and
$ Cf. Sokolnikoff, Advanced Calculus, New York, 1939, p. 434.
192 DIFFERENTIAL INVARIANTS 488

(3) is one-to-one. Instead of specifying a point Po by the coordi-


nates (xo, yo, zo), we may use instead the three numbers,
uo = u(xo, yo, zo), v0 = v(xo, yo, zo), wo = w(xo, yo, zo)
When (uo, vo, wo) are given, Po is located at the point of inter-
section of the three coordinate surfaces,
u(x, y, z) = uo, v(x, y, z) = vo, w(x, y, z) = wo.
These, in general, will intersect in three curves, the coordinate
curves, along which only one of the quantities u, v, w can vary.
For this reason u, v, w are called curvilinear coordinates, in dis-
tinction to the rectangular coordinates x, y, z, for which the co-
ordinate curves are straight lines.
If the position vector r is regarded as a function of x, y, z,
Vr = I (84.9). But, if r is regarded as a function of u, v, w,
ar ar car
Vr = Vu-+ Vv-+ Vw-,
au av aw,
from (83.7). We therefore have the fundamental relation,
(4) Vur,,
on writing ru = ar/au, etc. From the theorem of § 66, we now
conclude that:
The vector triples Vu, Vv, Vw and ru, r, rw form reciprocal sets.
The vectors ru are tangent to the u curves, the coordinate curves
along which v and w are constant. Thus at any point P(u, v, w),
r, r, rw are tangent to the three coordinate curves meeting there;
and, from (23.6),
(5) Vv Vw] = 7
Here [rurrw] is the Jacobian a(x, y, z)/a(u, v, w) of the inverse
transformation (3); and, from (2),
(6) J= 0.

Moreover, from the properties of reciprocal sets,


Jrw,
x x ru, ru x
(7) Vu = rv Vv = rJ Vw =

J r = J Vw J x Vv.
§ 88 CURVILINEAR COORDINATES 193

The volume of the parallelepiped whose edges are ru du, r dv,


rw dw is called the element of volume dV; thus
(9) dV = [rurvrw] du dv dw = J du dv dw.
To find the gradient of a tensor f (u, v, w), we first compute
df of du of dv of dw
(10)
ds au ds + av ds + aw ds
= e (Vufu + Cvfv+ Vwfw).
But, from (86.2), df/ds = e Vf for every e; hence
(11) of = Vu fu + Vv fv + Vw fw.
Thus, in curvilinear coordinates the operator del becomes
a a a
(12) V = Vu-+ Vv -- + Vw-
au ov aw
From (11), we obtain V f and V x f by dotting and crossing:
(13)

(14) Vxf = Vuxfu+ Cvxfv+ VwXfw.


For purposes of computation, it is usually more convenient to
eliminate Vu, Vv, Vw from these formulas by use of equations (7).
We thus obtain
1
{rvxrwfu + rwxrufv + ruxrvfw} ,
J
and corresponding equations for V f and V x f. In view of the
identity,
(rv x rw)u + (rw x ru) v + (ru x rv) w = 0,
these also may be written
1
(16) Of = { (rv x r f) u + (rw x ru f) v + (ru x rv f)w },
J
1
(17) =J

(18) Vxf= 1 [(rv x rw) X flu + [(rw x ru) x f]v + [(ru x rv) x f]w }.
J
194 DIFFERENTIAL INVARIANTS § 89

When the triple products in (18) are expanded, the brace becomes
(rw r71 f).u - (r rw f)u. + (r.u rw f) - (rw ru f)v
+ (rv ru f) w - (ru r f) w
= rw(rv f)u - rv(rw f), + ru(rw f) - rw(ru f) 1,

+ f)w - ru(rv f)w.


With this value for the brace, (18) may be written compactly in
determinant form:
ru r rw
1 a a a
(19) Vxf
J au av aw

This equation reduces to (84.7) when ru, r,,, rw are replaced by


rx = i, rv = j, rz = k.
89. Orthogonal Coordinates. The curvilinear coordinates
u, v, w are said to be orthogonal if the coordinate curves (along
which one coordinate only can vary) cut at right angles. Since
ru, r, r.v, are tangent to the coordinate curves, the coordinates are
orthogonal when and only when
(1)

We choose the notation so that


(2) ru = Ua, r = Vb, rw = Wc,
where a, b, c are a dextral set of orthogonal unit vectors and
U, V, W are all positive; then [abc] = 1, and
(3) J = [rurvrv,] = UVW.
The set of vectors reciprocal to ru, rv, rw are the gradients of the
coordinates:
a b c
(4) Vu = , Vv = , Vw = .
U V W
From the general formulas (11), (17), (19) of the preceding
article we obtain Vf, V f and V x f in orthogonal curvilinear co-
ordinates:
§ 89 ORTHOGONAL COORDINATES 195

1 1 1
(5) Vf = afu + bf + cf
U V W
1 i18 a
(6) V f =UVW (VW a f) -F a (WU b f)
Lou

+_
Ua Vb We
1 a a 8
(7) Vxf =
UVW au av aw

If we put f = Vg in (6), we obtain the Laplacian V2g = V. Vg


in orthogonal coordinates,
la a WU a UV \I
(8) o2g
UVTV
1
I/V\) 1/
au \ U gu + av \ V gv/ + aw \ W gw/ 1
\1
1/

When the curvilinear coordinates are given by the equations,


x = x(u, v, w), y = y(zt, v, w), z = z(u, v, w),
we may compute the (positive) functions U, V, TV from equations
of the type,
u + yu + u
U = I ru _ xui + yuJ + zuk = 1/x2z
(9) z2

The element of volume (88.9) is now


(10) dV = UVW du dv dw.
Example 1. Cylindrical Coordinates. The point P(x, y, z) projects into the
point Q(x, y, 0) in the xy-plane. If p,p are polar coordinates of Q in the xy-
plane, u = p, v = gyp, uw = z are called the cylindrical coordinates of P (Fig.
89a). They are related to rectangular coordinates by the equations:
(10) x = p cos (p, y = p sin gyp, z = Z.
From r = ix + j y + kz, we have
rp = [cos v, sin gyp, 0],
r, = [ -p sin rp, p cos gyp, 0],
r. = [0, 0, 11.
Since these vectors are mutually perpendicular and [rrr=] > 0, cylindrical
196 DIFFERENTIAL INVARIANTS § 89

coordinates form an orthogonal system which is dextral in the order P, c, Z.


Moreover,
(11) U= rn = 1, V = rw I = p, W
(12) J = UVW = P.
The level surfaces p = a, v = b, z = c are cylinders about the z-axis, planes
through the z-axis, and planes perpendicular to the z-axis. The coordinate

FIG. 89a

curves for p are rays perpendicular to the z-axis; for p, horizontal circles cen-
tered on the z-axis; for z, lines parallel to the z-axis. The element of volume
dV =Pdpdpdz.
From (8), the Laplacian is

- -8 (Pgp) +
1 1
(13) V2g =
P aP Pa
g = log P, Pgv = 1; hence log p satisfies Laplace's Equation V2g = 0. Such
a function is called harmonic.
Example 2. Spherical Coordinates. The spherical coordinates of a point
P(x, y, z) are its distance r = OP from the origin, the angle 0 between OP
and the z-axis, and the dihedral angle ' between the xz-plane and the plane
z OP (Fig. 89b). They are related to rectangular coordinates by the equations:
(14) x = r sin 0 cos p, y = r sin 0 sin gyp, z = r cos 0.
From r = ix + jy + its, we have
rr = [sin 8 cos p, sin B sin gyp, cos 0],
rs = [r cosBcosgyp,rcos©singyp, -rsin0],
r, = rsin0cos p, 0].
90 TOTAL DIFFERENTIAL 197

since these vectors are mutually perpendicular and [rrror,] > 0, spherical co-
ordinates form an orthogonal system which is dextral in the order r, 0,
Moreover,
(15) U=jrr1, V=Irol=r, Iiir,rsin o;
(16) J = UVIV = r2 sin 0.
The level surfaces r = a, 0 = b, = c are spheres about 0, cones about the
z-axis with vertex at 0, and planes through the z-axis. The coordinate curves

Fia. 89b
for r are rays from the origin; for B, vertical circles centered at the origin;
for gyp, horizontal circles centered on the z-axis. The element of volume dV
= 7-2sin0drdodrp.
From (8), we now have
1
(17) V2g isin 0 (r2gr) + (sin 0 go) + 9
r2 sin 0 or aB sin 0 0`°

If g = 1 /r, r2gr = -1; hence 1 /r satisfies Laplace's Equation V2g = 0.


O. Total Differential. In passing from the point P to P, the
---4 --->
position vector r = OP changes by the increment Ar = PP.
Then, if f(r) is any differentiable tensor point function, the total
differential of f (r) is defined by the equation :
(1)
In particular, when f = r, Vf = I, and
(2) dr = Or.
The differential of the position vector is the same as the increment
The defining equation (1) therefore may be written
(3) df=drVf.
198 DIFFERENTIAL INVARIANTS §91

If f is a function f (u, v, w) of curvilinear coordinates,


Vf = Vufu+VvfV+Vwfw,
and, from (3), since dr Vu = du, etc.,
(4) df = fu du + f dv + fw dw.
91. Irrotational Vectors. A vector function f (r) is said to be
irrotational in a region R when rot f = 0 in R.
If V(r) is a scalar function with continuous second derivatives,
its gradient VV is irrotational; for, from (84.10),
(1) rot grad ,p = 0.
Conversely, if rot f = 0, and f is continuously differentiable in R,
we shall show that f may be expressed as the gradient of a scalar
V(r). Using rectangular coordinates, let f = f1i + f2j + f3k; then,
if rot f = 0, the three determinants of the matrix,
(2) aax a/ay a/az\

(fl, f2 f3

all vanish (84.7). Under this condition we shall determine a


scalar function ,p(r), so that Vv = f, that is,

(3) = f1(x, y, z), = f2(x, y, z), = f3 (x, y, z)


ax ay az

Let (xo, yo, zo) be an arbitrary point of R. On integrating


app/ax = f1 with respect to x and regarding y and z as constant
parameters, we have
x
(4) P=
J o
f1(x, y, z) dx + «(y, z),

_ -dx+-=
where «(x, y) is a function of x and y as yet undetermined. Hence
x afl a« /' af2 a«
app

ay ,, ay ay
-dx+-,
Ixo ax ay
or
a

-_J
49Z 1o
xafl
- dx + - _
az az

f
f2(x, y, z) = f2(x, y, z) - f2(xo, y, z) +
xaf3
- dx +
ax az
- , or
ay

f3 (x, y, z) = f3 (x, y, z) - f3 (xo, y, z) +


-

-
§ 91 IRROTATIONAL VECTORS 199

Instead of three equations (3) for gp(x, y, z), we now have two
equations,
a« a«
(5) = f2(xo, y, z), = f3(xo, y, z)
ay az

for «(x, y), with the condition aft/az = af3/ay. The problem has
been reduced from three to two dimensions.

f
On integrating a«/ay = f2 with respect to y, we have
v
(6) a= f2(xo, y, z) dy + #(z),
bo

where $(z) is a function of z to be determined. Hence


-
a«fY af2 dy
az vo az +
d(3
dz -f Y a af3

y
dy +
dz
,

f3(xo, y, z) = f3(xo, y, z) - f3(xo, yo, z) +


or

d
da
(7)
dz = f3 (xo, y0, z)
We now have one equation (7) to determine $(z); and

(8) J f3 (xo, yo, z) dz.


o

(9) '= (x, y, z) dx + f


On collecting the results (4), (6), and (8), we have finally

ffi
D irect substitution shows that Vip = f; and, if
Yo
Y z

f2 (xo, y, z) dy + f f3 (xo, yo, z) dz.

is a second func-
tion for which V4, = f, V(¢ - cp) = 0, and 4 - (p is a constant.
Thus (9) gives the solution of equations (3), determined uniquely
except for an additive constant.
There are evidently five other forms for ,p which may be obtained
from (9) by permuting 1, 2, 3 and making the corresponding per-
mutation on x, y, z; for example,
z

(10) c = ff2(x y , z) dy + ffs(x, yo, z) dz + ffi(x, yo, zo) dx.


Y

Moreover, xo, yo, zo may be given any values that do not make the
integrands infinite.
200 DIFFERENTIAL INVARIANTS §91

In mathematical physics, it is customary to express an irrota-


tional vector f as the negative of a gradient. Thus, if 4,
we have
(11) f = -V ';
¢ is then called the scalar potential of f.
Example 1. When
f = 2xzi + 2yz2j + (x2 + 2y2z - 1)k,
we find that rot f = 0; hence f = V. With xo = yo = zo = 0, we have,
from (9),
=f X2xz dx +f 2yz2 dy
0 0
- f Zdz = x2z + y2z2 - z.
0

Example 2. Exact Equation. The differential equation,


(i)
is said to be exact when f dr = dp, the differential of a scalar. If (i) is exact,
we have, from (90.3), dr f = dr vv; and, since dr is arbitrary, f =
When f is continuously differentiable, f = V implies rot f = 0, and con-
versely. Therefore we may state the
THEOREM. If f is a continuously differentiable vector, in order that f dr = 0
be exact it is necessary and sufficient that rot f - 0.
Thus, in view of ex. 1, the equation,
2xz dx + 2yz2 dy + (x2 + 2y2z - 1) dz = 0,
is exact and may be put in the form dp = 0. Its general solution is = c,
that is,
x2z + y2z2 - z = C.
If (i) is not exact, a scalar X which makes Xf dr = 0 exact is called an inte-
grating factor of (i). The preceding theorem shows that X must satisfy the
equation,
(ii) rotaf = 0 or VX x f + X rot f = 0.
On multiplying (ii) by f , we have
(iii) rot f = 0.
f
Hence, when f dr = 0 admits an integrating factor, f rot f = 0. Con-
versely, when f is continuously differentiable, the condition (iii) ensures the
existence of an integrating factor. We shall prove this in § 105. For this
reason f rot f = 0 is called the integrability condition for f dr = 0.
When f dr = 0 is integrable, ?f = V (p. Then the vector field f(r), being
parallel to Dip, is everywhere normal to the level surfaces rp = c. The con-
dition (iii) therefore implies the existence of a one-parameter family of sur-
faces everywhere normal to f. Thus the geometrical content of the condition
f rot f = 0 is that the vector field f (r) is surface-normal.
§ 92 SOLENOIDAL VECTORS 201

92. Solenoidal Vectors. A vector function f(r) is said to be


solenoidal in a region R when div f = 0 in R.
If g(r) is a vector function whose components have continuous
second derivatives, rot g is solenoidal, for, from (84.13),
(1) div rot g = 0.

Conversely, if div f = 0, we shall show that f may be expressed


as the rotation of a vector g.
Using rectangular coordinates, let f = f1i + f2j + f3k, and (84.5)
aft
(2) div f aftax= -+-+-
ay
af3
= 0.
az
We now shall determine a vector g = g1i + g2j + g3k, so that
i j k
(3) f = rot g = a/ax c1/ay a/az

91 92 93

We first find a particular solution of (3) for which g3 = 0; then


(3) is equivalent to the scalar equations:
4992 a91 4992 _ 1991
(4) fl = - az
, f2 = az
, f3 = ax ay

The first two equations of (4) are satisfied when


z

(5) 92 = - f f i (x, y, z) dz, 91 = f f2 (x; y, z) dz + 49(x, y) ;


w z Zo

in these integrations x and y are regarded as constant parameters,


and a(x, y) is a function as yet undetermined. In order that these
functions satisfy the third equation of (4),

- Zaf + J dz - as = f3,
(ax ay/Zp ay
.

aJ2

or, in view of (2),


Zafs as
dz - - = f3 (x, y, z).
I. az ay
When we perform the integration, this reduces to
as
-f3 (x, y, z0) - = 0,
ay
202 DIFFERENTIAL IN VARIANTS § 92

an equation which is satisfied by taking


v
a (X, y) = - f3 (x, y, zo) dy.
Yo

Hence the vector g, whose components are


z v
(6) gi = ff2(x, y, z) dz - ff3(x, y, zo) dy,
a
z

92 = - f fi (x, y, z) dz, 93 = 0,
zo

is a particular solution of our problem.


If G is any other solution, rot G = rot g = f, and hence
(7) rot (G - g) = 0.
But any irrotational vector may be expressed as the gradient of a
scalar p (§ 91); hence the general solution of (7) is G - g = o,p,
where ,p(r) is an arbitrary twice-differentiable scalar. When
div f = 0, the general solution of rot g = f is therefore g + Vip;
its rectangular components are obtained by adding c /cx, 8,p/cy,
8,p/8z to the components of g given in (6).
In mathematical physics, the solenoidal vector f = rot g is said
to'be derived from the vector potential g.
Example 1. When
f = x(z - y)i + y(x - z)j + z(y - x)k,
we find that div f = 0. Therefore f = rot g; if we take zo - yo = 0, the par-
ticular solution (6) is

91 = f Zy(x - z) dz =xyz - Zyz2,

z
g2 f x(z - y) dz =
0
- Zxz2 + xyz, ga = 0.

Example 2. If u and v are continuously differentiable scalars, the vector


Vu x Vv is solenoidal, for, from (85.3),
Vu x Vv = rot (uVv).
Conversely, we shall show in § 104 that a continuously differentiable sole-
noidal vector always can be expressed in the form Vu x Vv.
§ 93 SURFACES 203

93. Surfaces. A surface is represented in parametric form by


the equations:
(1) x = x(u, v), y = y(u, v), z = z(u, v).
We assume that the three functions of the surface coordinates u, v
are continuous and have continuous first partial derivatives, a
requirement briefly expressed by saying that the functions are con-
tinuously differentiable. In order that equations (1) represent a
proper surface, we must exclude the two cases:
(i) the functions x.(u, v), y(u, v), z(u, v) are constants: equations
(1) then represent a point;
(ii) these functions are expressible as functions of a single vari-
able t = t(u, v); equations (1) then represent a curve.
In case (i) all the elements of the Jacobian matrix,

(2)

vanish. In case (ii), the rows of the matrix are dx/dt, dy/dt, dz/dt
multiplied by at/au and at/av, respectively, and all of its two-
rowed determinants vanish identically. We exclude these cases
by requiring that the matrix (2) be, in general, of rank two; then,
at least one of its two-rowed determinants,
yu zu zu xu _ I xu
(3) A = B = C Yu ,
yv zv Zv xv xv yv
is not identically zero.
Even when the matrix in general is of rank two, the three de-
terminants A, B, C, all may vanish for certain points u, v. Such
points are called singular, in contrast to the regular points, where
at least one determinant is not zero.
If we introduce the position vector to the surface,
(4) r = xi + yj + zk = r(u, v),
(5)
and the condition for a regular point may be written
(6) ru x r, /- O.
204 DIFFERENTIAL INVARIANTS §94

If we introduce new parameters u, v by means of the equations,


(7) u = u(u, v), v = v(u, v),
we shall require that the Jacobian of this transformation,
a (u, v)
(8) J = FK 0.
a(u, i)

Then equations (7) may be solved for u and v, yielding


(9) u = u(u, v), v = 17(u, v),
and the correspondence between u, v and u, v will be one to one.
Since
X _ au av) x (ru au av\ X
(lp) rU ru - - rU - -- r -- J = J ru ry,
au au av aU

the requirement J 0 0 makes ru x rv 0 a consequence of (6).


Thus a point which is regular with respect to the parameters u, v
is also regular with respect to u, D.
94. First Fundamental Form. A curve on the surface r(u, v)
may be obtained by setting u and v equal to functions of a single
variable t:
(1) u = u(t), v = v(t).
A tangent vector along the curve (1) is given by
dr
dt

and
(2)
The are s along the curve is defined as in (43.4) :
fro' ds
t t dt; and
dt
=

Since du = ft dt, dv = v dt by definition, on multiplying (2) by dt2,


we have
(3) ds2 = ru ru du2 + 2ru r du dv + r r, dv2.
This first fundamental quadratic form usually is written
(4) ds2 = E du2 + 2F du dv + Gdv2,
§ 94 FIRST FUNDAMENTAL FORM
where
(5) E = ru ru, F = ru r,,, G = r r,,.
Moreover, from (20.1),
ru - ru ru rv
(ru x rv) (ru x rv) _
hence

ru x rv 12 = EG - F2

is positive at every regular point.


The curves v = const (u-curves) and u = const (v-curves) are
called the parametric curves on the surface. For these curves
dv = 0 and du = 0, respectively, and the corresponding elements
of arc are
(7) dsl = 1/E du, ds2 = N/G dv.
Since the vectors ru, rv are tangent to the u-curves and v-curves,
respectively, the parametric curves will cut at right angles when
and only when
.(8)
The vector ru x rv is normal to the surface. The parallelogram
formed by the vectors ru du, rv dv, tangent to the parametric curves
and of length dsl, ds2i has the vector area (§ 16),
(9) dS = ru x rv du dv.
We shall call this the vector element of area; the scalar element of
area is
(10) dS = I ru x rv I du dv = E du dv.

The unit normal n to the surface will be chosen as

n= ruxrt =
ruxrv
E H
then dS = n dS. At every regular point the vectors ru, r, n form
a dextral set; for
(12) [rurvn] = H = VEG > 0.
206 DIFFERENTIAL INVAItIANTS § 95

95. Surface Gradients. Let f(u, v) be a differentiable function,


scalar, vector, or dyadic, which is defined at the points of the
surface r = r(u, v). We shall compute the derivative of f(u, v)
with respect to the are s along a surface curve u = u(t), v = v(t).
Along this curve,
df du dv
(1)
ds fu ds + fv ds '
where du/ds = is/s, dv/ds = v/s, and, from (93.4),
ds
ds = V L ic2 + 2Ficv + Gv2.
-
(2)
dt
If we apply (1) to the position vector r(u, v), we obtain the unit
tangent vector e to the curve:
du dv
(3) e=r,,,+rv
ds ds

Let a, b, c denote the set reciprocal to ru, r,,, n; then, since


[rur ,n] = H,
_ rvxn _ nxru ruxrv
b II , c = = n.
(4) a H , H
Now from (3) a e = du/ds, b e = dv/ds; hence (1) may be
written
df
ds
We now define afu + bf as the surface gradient of f and denote it
by V8f t or Grad f:
(5) V8f = Grad f = afu + b f,,.
Grad f has the characteristic properties:
df
(6) e- Grad f = , n Grad f = 0;
ds

(7) ru Grad f =
of
-au , r Grad f = -
of
av
t In surface geometry we shall write of for the surface gradient.
96 SURFACE DIVERGENCE AND ROTATION 207

If f (r) is a tensor of valence v, Grad f is of valence v + 1. At


any point (u, v) of the surface, Grad f is in effect a synthesis of
all the values of df/ds for surface curves through this point. Since
Grad f depends only on the point.(u, v), df/ds is the same for all
surface curves having the unit tangent vector e at this point.
At any point (u, v) of the surface Grad f is completely deter-
mined when df/dsi is given for two directions ei in the tangent
plane at (u, v). If the set reciprocal to el, e2, n is e', e2, n,*
df + e2 df
(8) Grad f = el
dsi ds2

for, by virtue of equations (6), the right member of (8) may be


written
(elel + e2e2 + nn) Grad f = I Grad f
where I is the idemfactor (§ 66). In particular if el and e2 = n x el
are perpendicular unit vectors, el = el, e2 = e2. In view of (8),
Grad f is independent of the coordinates x, y, z and of the surface
parameters u, v.
96. Surface Divergence and Rotation. If f(r) is a tensor point
function defined over the surface r = r(u, v), its surface gradient,
(1) V8f=afu+bf,,,
has the invariants,
(2)
(3) V,-f = axfu +bxf,,.
We recall that the set a, b, n is reciprocal to ru, rv, n.
When f (r) is a vector, V8f is a planar dyadic; then the scalar
and vector invariants of oaf are called the surface divergence and
surface rotation and are written
(4) V8 f = Div f, V8 x f = Rot f.
The second of V8f is the linear dyadic,
1
(V3f)2 = axbfuxfv = Hnfuxfv;

the second scalar of V8f is therefore n f,, x f v/H.


* Since el x e2 = An, e3 = n, we have e3 = Xn/X = n.
208 DIFFERENTIAL INVAItIANTS 196

For the position vector r to the surface, we have


(6) V8r = ar.u + br v = I - nn,
(7) Div r = 2, Rot r = 0.
For the unit normal n,
(8) Rot n = 0.
To prove this, put f = n, a = r x n/H, b = n x ru/H in (3); then
H Rot n = (rz, x n) x nu + (n x ru) x n,,.
=
Now, from n n = 1, we have, on differentiation with respect to
u and v,
(9) n nu = n nv = 0;
and, from ru n = r n = 0,
(10) ru nv = rv nu = -n ruv;
hence H Rot n = 0.
From the defining equations (1), (2), (3), we may derive various
expansion formulas. Thus, if X is a scalar, f a tensor,
(11) V3(af) = (V3X)f + XV3f,

(12) V. (Af) _ (V8A) f + XV8 f,


(13) V8 x (Xf) _ (V8X) x f + XV3 x f.

If g and f are vectors,


(14) Div (g x f) _ (Rot g) f - g Rot f;
and, in particular,
(15) Div (n x f) = -n Rot f.
Proof of (14) :
Div(gxf)

(Rot g) f - g Rot f.
§ 97 SPATIAL AND SURFACE INVARIANTS 209

97. Spatial and Surface Invariants. If f(r) is a tensor function


of valence v defined over a 3-dimensional region including the
surface r = r(u, v), its spatial gradient at a point (u, v) of the
surface may be computed from (86.4). If el, e2 are vectors in the
tangent plane at (u, v) and e3 = n, then, at all points of the sur-
face,
df df df df
(1) Vf = e'-+e2-+n-
dsl ds2
=
do
V8f +n -
do ;

here the set e', e2, n is reciprocal to el, e2, n and df/dn denotes
the derivative of f in the direction of n. Moreover, if v > 0,
df df df df
(2) V
ds.l ds2 do do
df df df df
(3) Vxf = elx-+e2x-+nx-=
dsl ds2 do
V8xf +nx-.
do
From (1), we see that the tensors of valence v + 1,
(4) nx Vf = nx V j,
are the same over the surface; therefore both may be written
n x Vf. Thus n x Vf may be computed solely from the values of f
on the surface.
The same is true of the invariants obtained from n x Vf by dot-
ting and crossing in the first position. Since
(5) nxVf = nxafu+nxbft,,
from (96.1), this process yields
(n x a) fu + (nxb) f = n (a x fu + b x f,) = n V. x f,

n208f -nV8 f.
Here n V8f means that n is dotted into the second vector from
the left in each term of V8f. These invariants remain the same
when computed from the corresponding spatial quantities. Thus we
verify at once, from (3),
(6)
and, from (1) and (2),
(7) nVf-nV.f=n2V8f-nV8f.
210 DIFFERENTIAL INVARIANTS § 97

When f is a vector, these invariants become


(6)' n rot f = n Rot f,
(7)' (grad f) n - n div f = (Grad f) n - n Div f.
We now express n x Vf and its invariants (6) and (7) in terms
of the surface coordinates u, v. Since a, b, n and ru, r,,, n are re-
ciprocal sets, we have, from (5),
1 1
(8) nx Vf
- (r,fu - ruf
II
{(rvf)u - (ruf),, },

since ruv = r,,,, if these derivatives are continuous. Dotting and


crossing now yields
1
(9) n Vxf II

(10) II{(rvxf)u-(ruxf)v}.

Since ru, rv are tangent to the surface, (9) shows that: If a


vector f is everywhere normal to the surface, n rot f 0.
With f = V3g in (9) we obtain the important identity,
(11) =0;
for then
f = gv, rv ru f = gu (95.7).
When f = pq, a dyad, we have, from (9),
1
n V x (pq) = (r,, Pq) u - (ru pq)
II {

1 1
= { (r,, P)u - (ru P) v } q + p (rvqu - rugv),
H H
that is,
(12) n V x (pq) = (n rot p)q + p n x Vq.
For future use we compute the invariant (10) when f = ng and
g is an arbitrary tensor:

n2 Vf -nVf= 1(rvxng)u - (ruxng)v}


H
1
= agu + bg,, + {(rvxn)u - (ruxn)v} g.
H
§ 98 SUMMARY: DIFFERENTIAL INVARIANTS 211

Now
Grad g = agu + bg (95.5),
2
n Grad n = (an. + n=0 (96.9) ;
hence, on putting f = n in (10), we have
1
-n Div n = { (r x n)u - (ru x n) v } .
H
Thus, with f = ng,
(13) n? Vf - n V f = Grad g - (Div n) ng.
98. Summary : Differential Invariants. Let f (r) be a tensor
point function of valence v; its derivative at the point P and in
the direction of the unit vector e is defined as
df f (r + se) - f (r)
-- = lim (s > 0).
ds s -- o s
If df/dsi denote the directional derivatives corresponding to three
non-coplanar unit vectors ei, the gradient of f, namely,
df
grad f = Vf = e' + e2 + e3 ,
ds 1 2 63
has the property,
df
-- =e V.
ds
Vf, a tensor of valence v + 1, thus gives a synthesis of all the
directional derivatives of f at point r. If the vectors ei are i, j, k,
Vf = if., + jfv + kfz (fz = of/ax, etc.).
From Vf (valence v + 1) we derive the invariants V f (valence
v - 1) and V x f (valence v) by dotting and crossing in the first
position. When f is a vector (v = 1),
V f = div f, the divergence of f,
V x f = rot f, the rotation of f.
When r is a function r(u, v, w) of curvilinear coordinates
Vf = Vu fu + Vv fv + Vw f.W,
Vr = Vu ru + Vv r + Vw rw = I.
212 DIFFERENTIAL INVARIANTS § 98

The sets Vu, Vv, Vw and ru, r',,, ru, are reciprocal. With J =
we have
J Vf = (r x ru, f)u + cycl,
J V f = (r, x ru, f)u + cycl,
ru r ru,
J V x f = ((r x ru,) x f)u + cycl = a/au a/av a/aw

ru, ru, are mutually orthogonal, the coordinates u, v, w


are called orthogonal. Then if a, b, c denote a dextral set of or-
thogonal unit vectors, ru = Ua, r = Vb, ru, = Wc, and J =
UVW in the preceding formulas.
For any tensor f (r),
V Vf = V2f, V. Vf = 0,
V (V f) = 0, Vx (Vxf) = V(V _ f) - V2f.

The operator V V. V is called the Laplacian.


A vector f is called irrotational if rot f = 0, solenoidal if div f
= 0. An irrotational vector f can be expressed as the gradient of
a scalar (f = Dip) ; a solenoidal vector f can be expressed as the
rotation of a vector (f = rot g).
For the surface r = r(u, v), the fundamental quadratic form is
ds2 = E due + 2F du dv + G dv2,
where

At a regular point,
E>0,
and the unit surface normal is defined as n = ru x
The derivative of a tensor f (r) along any surface curve tangent
to the unit vector e at the point (u, v) is
df
ds
the set a, b, n is reciprocal to ru, r, n.
The surface gradient,
V8f = Grad f = a fu + b f,,,
PROBLEMS 213

has the properties,


df
e. Grad f = , n Grad f = 0;
ds
df df
r.u Grad f = - , r, Grad f =
au av
From V3f we derive the invariants,
O3 f = Oaxf = axfu+bxfv,
by dotting and crossing in the first position. When f is a vector,
V8 f = Div f, the surface divergence of f,
Og x f = Rot f, the surface rotation of f.
The tensor n x of and its dot and cross invariants,
n?of
are not altered when V is replaced by V8. They may be computed
solely from the values f(r) assumes on the surface. Thus, if
H=
1
n x of = { (ru.f) v } ;
H
and the invariants follow by dotting and crossing between r,, r
and f.
PROBLEMS
1. If R = r/r is the unit radial vector, prove that
div R = 2/r, rot R = 0.
2. For any scalar function f (r) of r alone prove that
V2f(r) = frr + 2fr/r.
If v2f (r) = 0, show that f = A/r + B.
S. If a is a constant vector, prove that
grad (a. r) = a, div (a x r) = 0, rot (a x r) = 2a.
4. If a is a constant vector, prove that
grad (a - f)
div (a x f) = -a rot f,
rot(axf)
Specialize these results when f = r.
214 DIFFERENTIAL INVARIANTS
5. For any vector point function f prove that

6. If Sp and ¢ are scalar point functions, prove that


div (Vp x v) = 0
7. If e is a unit vector, prove that
div (e r)e = 1, rot (e r)e = 0;
div (e x r) x e = 2, rot (exr) xe = 0.
8. If a is a constant vector, prove that V(r x a) = I x a.
9. Show that Laplace's Equation V2f = 0 in cylindrical and spherical co-
ordinates is
a2f 1 of 1 a2f a2f
= 0,
apt + p (9p + p2 a22
and
r{ -- sin0-J+
a2(rf)
are
a/ af\
1

sin 0 a0 a0
1 a2f
sine 0 5;i
=0.
10. Prove that
1 1
vC T =
- (3RR - I)
where R is a unit radial vector.
11. If f is a vector point function, prove that
V (Vf), = grad div f; V x (Of), _ (V rot f),.
12. If Vf is antisymmetric, prove that rot f is constant and that the dyadic
Vf itself is constant.
13. If X is a scalar point function, prove that
V(,I) = VXI, V V. (XI) = VX, V x (XI) = VX x I.
14. If f is a vector point function, prove that
V (I x f) = rot f, v x (I x f) = (Vf), - I div f.
15. For the dyad fg prove that
V V. (fg) = (div f)g + f Vg, V x (fg) _ (rot f)g - f x Vg.
In particular, if r is the position vector,
V V. (rr) = 4r, V x (rr) _ -I x r.
16. If is a constant dyadic, prove that
div 4 . r = 91; rot e r = -+ (§ 69).
17. If the scalar function f(p, gyp) in plane polar coordinates is harmonic,
prove that f(p-', tip) is also harmonic.
18. If the scalar function f(r, 0, Sp) in spherical coordinates is harmonic,
prove that r 'f (r ', 0, gyp) is harmonic.
PROBLEMS 215

19. If f, g, h are scalar point functions, prove that


V2(fg) = g02f + 2Vf Vg + fV2g;
V2(fgh) = gh V2f + hf V2g + fg v2h + 2fvg vh + 2gvh of + 2hvf vg.
20. If u, v, w are orthogona' coordinates and f(u), g(v), h(w) are scalar func-
tions of a single variable, show that

V2(fgh) = fgh { f + egg + hh

21. If a is a constant vector and f = ar", prove that


of = nrs-2ra, e'-f = n(n + 1)ri-2a.
22. If f = r"r, prove that
of = r"I + nrn-err, V2f = n(n + 3)rn-2r.
23. If f = r"r, find a scalar function p such that f = vcp.
24. Prove that rot f = 0, and find V so that f = vp:
(a) f = yzi + zxj + xyk; (b) f = 2xyi + (x2 + log z)j + y/z k;
(c) f = c r, cD a constant symmetric dyadic.
25. If a is a constant vector and f = r" a x r, prove that
div f = 0, rot f = (n + 2)rna - nrn-2(a r)r.
26. Find a vector g such that rot g = a x r (cf. § 92).
27. Prove that
ar
= - rot --
r'
.

28. The position vectors from 0 to the fixed points P1, P2 are r1, r2.
(a) If f = (r - r1) x (r - r2), show that
divf =0, rotf =2(r2 -r1).
(b) Prove that
V(r-rl) (r-r2) =2r-r1-r2.
29. If f(u, v) is a vector function defined over the surface r = r(u, 1'), prove
that
H Div f =n (ruxf -r, xfu).
30. If the vector f(u, v) is always normal to the surface r = r(u, v), prove
that Rot f is tangent to the surface or zero.
31. If u, v, w are curvilinear coordinates, prove the operational identities:
a a a
'le = ruxrw-+ru,xru-+ruxrv-;
[rurvru7

au av aw
a
(ruxr,,)xV =rv -- ruav-;
a
ru.V =_.
a
au au
CHAPTER VI

INTEGRAL TRANSFORMATIONS

99. Green's Theorem in the Plane. Let R be a region of the


xj-plane bounded by a simple closed curve C which consists of a
finite number of smooth arcs. Then, if P(x, y), Q(x, y) are con-
tinuous functions with continuous first partial derivatives,

(1) J I1
CIQ
x - y) dx dy = f (P dx F Q dy),

where the circuit integral on the right is taken in the positive


sense; then a person making the circuit C will always have the
region R to his left.

Fra. 99a

Consider first a region R in which boundary C is cut by every


line parallel to the x- or y-axis in two points at most (Fig. 99a).

f
Then, if a horizontal line cuts C in the points (x1, y), (x2, y),
a
f f aQ dx dy = {Q(x2, Y) - Q(xi, y)) dy
fQ(X2,
a
= y) dy + fQ(xiy) dy

= fQ(x, y) dy.
216
§ 99 GREEN'S THEOREM IN THE PLANE 217

And, if a vertical line cuts C in the points (x, yl), (x, y2),
b

If ay
dy dx = f I P(x, Y2) - P(x, Y1)) dx
fb( yl) dx - P(x, y2) dx
J6

= - fP(x, y) dx.
On subtracting this equation from the preceding, we get (1).
We now can extend this formula to more general regions that
may be divided into a finite number of subregions which have the
property that a line parallel to the x- or y-axis cuts their boundary

Frc. 99b Fia. 99c

in at most two points. For in each subregion formula (1) is valid,


and, when these equations are added for all the subregions, the
surface integrals add up to the integral over the entire region; but
the line integrals over the internal boundaries cancel, since each
is traversed twice, but in opposite directions, .leaving only the line
integral over the external bounding traversed in a positive direc-
tion (Fig. 99b).
The boundary of R may even consist of two or more closed
curves: thus in Fig. 99c the region R is interior to Cl but exterior to
C2; we may now make a cut between Cl and C2 and traverse the
entire boundary of R in the positive sense as shown by the arrows.
The line integrals over the cut cancel, for it is traversed twice
and in opposite directions; and the resultant line integral over C
consists of a counterclockwise circuit of Cl and a clockwise circuit
of Cam.
218 INTEGRAL TRANSFORMATIONS § 100

Although Green's Theorem is commonly stated for scalar func-


tions P, Q, it is evidently true when P and Q are tensors with
continuous first partial derivatives; for, when P and Q are referred
to a constant basis, the theorem holds for each scalar component.
100. Reduction of Surface to Line Integrals. A surface is said
to be bilateral if it is possible to distinguish one of its sides from
the other. Not all surfaces are bilateral. The simplest unilateral
b
surface is the Mobius strip; this
a,

may be materialized by taking a


a b'
rectangular strip of paper, giving
FIG. 100a
it one twist, and pasting the ends
ab and a'b' together (Fig. 100a). This surface has but one side;
if we move a point P along its median line and make a complete
circuit of the strip it will arrive at the point P directly under-
neath. Since we can travel on a continuous path from one side
of the Mobius Strip to the opposite, these sides cannot be distin-
guished. This is not the case with a spherical surface, which has
an inside and an outside, or any portion S of the surface botinded
by a simple closed curve C; for we cannot pass from one side of S
to the other without crossing C.
Let S be a portion of a bilateral surface r = r(u, v) bounded by'
a simple closed curve C which consists of a finite number of smooth
arcs. The surface itself is assumed to consist of a finite number
of parts over which the unit normal n is continuous. The positive
sense on C is such that a person, erect in the direction of n, will
have S to the left on making the circuit C. If such an oriented
surface is continuously deformed into a portion of the xy-plane
bounded by a curve C', n becomes k (the unit vector in the direc-
tion +z) and the positive circuit on C' forms with n = k a right-
handed screw.
Let f (r) be a tensor point function (scalar, vector, dyadic, etc.)
whose scalar components have continuous derivatives over S.
Such a tensor is said to be continuously differentiable. When the
foregoing conditions on S and C are fulfilled, we then have the
BASic THEOREM I. If f(r) is a continuously differentiable tensor
point function over S, the surface integral of n x Of over S is equal
to the integral of Tf taken about its boundary C in the positive sense:

(1) fn Vf dS = JTf ds.


§ 100 REDUCTION OF SURFACE TO LINE INTEGRALS 219

Proof. Since
1
n x Vf = { (ref),}, dS = H du dv,
II
fn x Vf dS = { (r ,f).u - du dv,
s'
IS,

where S' is the region of the uv-plane which contains all parameter
values u, v corresponding to points of S. We now may apply

FIG. 100b

Green's Theorem in the plane to the last integral, letting u and v


play the roles of x and y; thus

fs, { (rtf)u - du dv = f(rtf du + dv),

where the circuit integral is taken about the curve C', which forms
the boundary of S', in the positive sense, that is, in the direction
of a rotation of the positive u-axis into the positive v-axis. If we
regard u and v as the functions of the are s on the curve C, it
becomes
/
I \
du
ds
dv
ruf-+rf-)ds =ic Tfds, ds

du dv dr
ru ry = T,
ds + ds A
the positive unit tangent along C. Formula (1) thus is established.
220 INTEGRAL TRANSFORMATIONS §100

If we introduce the vector elements of surface and of arc,


dS = n dS, dr = T ds,
the basic theorem for transforming surface to line integrals be-
comes

(2) fdS X Vf =fdr f,


in which dS and dr must be written as prefactors.
If f is a vector, dyadic, etc., we may obtain the other integral
transformations from (1) by placing a dot or a cross between the
first two vectors in each term (dyad, triad, etc.) of the integrands;
for both of these operations are distributive with respect to addi-
tion and therefore may be carried out under the integral sign.
This process applied to the tensor n x Vf gives n V X f and n z Vf
- nV f; we thus obtain the important formulas:

(3) ffl.VxfdsfT.fds,
(4)
f (n z Vf - n V f) dS = if T X f ds.
When f is a vector function, (3) becomes Stokes' Theorem:

(3)' fn.rotfds=fT.fds ,

a result of first importance in differential geometry, hydrodynamics


and electrodynamics.
If S is a closed surface which is divided into two parts S1, S2
by the curve C, we may choose n as the unit external normal;
then, from (1),

fnxVfds = fTlf ds, fn,< Vf dS = fT2f ds,


c
where T2 = - T1, since the positive sense on C regarded as
bounding S1 is reversed when regarded as bounding S2. On adding
these integrals, we find that the integral of n X Vf over the entire
surface is zero. We indicate this by the notation,

(5) fnx Of dS = 0,
§ 101 ALTERNATIVE FORM OF TRANSFORMATION 321

which denotes an integral over a closed surface. On taking dot


and cross invariants in (5), we have also

(6) fn.VxfdS=0,
(7) f(n'4Vf_nV.f)dS=0.
Example 1. Put f = r in (4) and (7); then, since grad r = I, div r = 3,

J'ndS = 2irxdr, fnds = 0.


The last result, which states that the vector area of a closed surface is zero,
generalizes the polyhedron theorem of § 17.
Example 2. When div f = 0, we have seen that f = rot g (§ 92); hence

f
Thus, if f = Vu x Vv, div f = 0, and we may take g as uVv or -vVu; hence

Jn. Vu x Vv dS = it dv = - Jv du = 2 f(u dv - v du).


In particular, when u = x, v = y, n = k, this formula expresses an area in the
xy-plane as a circuit integral 2 J(x dy - y dx) over its boundary.
101. Alternative Form of Transformation. If we replace f by
ng in (100.4), where g is any continuously differentiable tensor,
we have, from (97.13),

(1) f { Grad g - (Div n) ng } dS = fT x n g is.


This integral transformation is quite as general as that given by
basic theorem I. For, if we replace g by of and then cross in the
first position, the integrand on the left becomes
a x (nuf + n f.u) + b X n fv) _ -n x Grad f,
since Rot n = 0; whereas, on the right,
(Txn) Xnf = -Tf.
We thus retrieve the basic transformation (100.1).
222 INTEGRAL TRANSFORMATIONS § 102

We now shall express (1) in slightly different notation. On the


right the vector T x n = m is the unit normal to C tangent to S
and pointing outward. If we write f instead of g and J = - Div n
(J is not the Jacobian of § 88), (1) now becomes

(2) f (vsf + J nf) dS = fm f ds.


On dotting and crossing in the first position of the tensor inte-
grands, we obtain the additional formulas:

(3) f(vs.f+Jn.f)ds =
f
f f f=
Div n is called the mean curvature of the surface (§ 131).
Surfaces for which J = 0 are called minimal surfaces; for such
surfaces the integrands on the left reduce to the first term.
On the plane, n = k, a constant vector, and Div n = 0; then
J = 0 in (2), (3), and (4). When f is a vector, these equations
become (if we write dA for dS) :

(5) fGrad f dA = i(m f ds,

(6) fDivfdA
(7) JRotfdA = fm f ds,
-
where m is the external unit normal to the closed plane curve
forming the boundary. We may deduce (6) from (7) by replacing

rf
fbykxf;for
Rot (k x f) = k Div f,m x (k x f) = km f.
102. Line Integrals. When f(r) is a vector, consider the line
integral,
(1) J'f.dr=
o
to
-
dr
dt
dt,

ta ken over a curve C: r = r(t) from the point P0 to P. The value


of this integral depends on the curve C and the end points Po and
§ 102 LINE INTEGRALS 223

P, but not on the parameter t. For, if we make the change of


parameter t = t(T),

Jcu
t

f -dr
dt
dt= pT

TO
f---drf
dr dt

dt dr
- dr.
ar
,,
r
dr

Let us consider under what conditions the line integral (1) is in-
dependent of the path C-that is, when its value is the same for
all paths from Po to P.
We shall call a closed curve reducible in a region R if it can be
shrunk continuously to a point without passing outside of R.
Thus in the region between two concentric spheres all closed
curves are reducible. However in the region composed of the
points within a torus, all curves that encircle the axis of the torus
are irreducible. A region in which all closed curves are reducible
is called simply connected. Thus the region between two concen-
tric spheres is simply connected, whereas the interior of a torus
is not.
If f is continuously differentiable, and rot f = 0 in a region R,
the line integral of f around any reducible closed curve in R is zero;
for we can span a surface S over R that lies entirely within R
(shrinking the curve to a point generates such a surface), and
then, by Stokes' Theorem,

ff . dr = fn. rotfdS=0.
We may now prove the
THEOREM. If f is a continuously differentiable vector, and rot f
= 0 in a simply connected region, the line integral J'f dr between
any two points of the region is the same for all paths in the region
joining these points.
Proof. If C1 and C2 are any two curves POAP, POBP joining
PO to P, the line integral of f around the circuit POAPBPO is zero;
for all closed curves in a simply connected region are reducible.
Hence

J oA P
fPBP,
P Po
f dr=0.
224 INTEGRAL TRANSFORMATIONS § 103

Under the conditions of this theorem, the line integral (1) from
a fixed point Po to a variable point P of the region depends only
upon P; in other words the integral defines a scalar point function,

(2) p(r) = Jf dr.


o

Let us compute d<p/ds = e Vg for an arbitrary direction e. We


have
,p(r + se) - p(r) =
f +8e
f 0
8

since dr = e ds along the ray from r to r + se. By the law of the


mean for integrals,

ff.eds = sef(r+e), (0 < 9 < s),

and hence
d(P + se) - p(r)
= lim p(r = e f(r).
ds s -o s
Since e - Vp = e f for any e,
(3)

Thus an irrotational vector f may be expressed as the gradient


of scalar gyp, given by (2). The determination of p given in § 91 is
the line integral of f taken over a step path from (xo, yo, zo) to
(x, y, z). Thus in (91.9) the integrals, in reverse order, are taken
over the straight segments from
P0(xo, Yo, zo) - (x0, Yo, z) - (x0, y, z) - P(x, y, z).
There are six such step paths; for the first segment can be chosen
in three ways and the second in two.
In mathematical physics it is customary to express an irrota-
tional vector as the negative of a gradient:
jTO
(4) f = - V4,; then ¢(r) = dr
is called the scalar potential of f.
103. Line Integrals on a Surface. We next consider line and
circuit integrals over curves restricted to lie on a given surface,
or within a certain region S of the surface. A region S of a surface
§ 103 LINE INTEGRALS ON A SURFACE 225

is called simply connected if all closed surface curves in S are


reducible-that is, can be shrunk continuously to a point without
passing outside of S. In the plane, for example, the region within
a circle is simply connected, but the region between two concentric
circles is not.
If the vector f is continuously differentiable, the circuit integral
if dr about any reducible curve of S is zero when rot f is every-
where tangential to the surface; for

f
When S is a portion of the level surface u(x, y, z) = c of a scalar
function, Vu on the surface is parallel to the surface normal and
the condition n rot f = 0 may be written
(1) 0.
Just as in § 101, we may now prove the
THEOREM 1. If f is a continuously differentiable vector, and
n rot f = 0 in a simply connected portion S of a surface, the line
integral ff dr between any two points of S is the same for all
surface curves joining these points.
Under the conditions of this theorem the line integral,
r
(2) p(r) = f dr,
ra

is the same over all surface curves from Po to P and therefore


defines a scalar point function V(r) in S. Along any definite curve
r = r(s) issuing from P0 (s = 0) c is a function of the are s,
s dr 8

p(s) =J'ff Tds,


where T is the unit tangent vector; hence at any point P d(p/ds
= f T. Since the curve may be varied so that T assumes any
direction e at P, we have the relation,
d
(3)
s
226 INTEGRAL TRANSFORMATIONS § 104

for all unit vectors e in the tangent plane at P. Now if e1, e2


are two such vectors, and el, e2, n the set reciprocal to e1, e2, n,
we have, from (95.8),

Grad cp = el
d + e2 d = ele1 f + e2e2 f.
ds1 ds2
Since f = (ele1 + e2e2 + nn) f,
(4) Grad 'P = ft,
the projection of f on the tangent plane at P. We have thus
proved the

THEOREM 2. If f is a continuously differentiable vector, and


n rot f = 0 in a simply connected region of a surface, the tangential
projection of f on the surface is
ft = Grad 9 where 9 = f dr.
ro

104. Field Lines of a Vector. When f(r) is a continuously dif-


ferentiable vector function, the curves tangent to f(r) at all their
points are called the field lines of f. If r = r(t) is a field line,
dr/dt is tangent to the curve (§ 41) and consequently a multiple
of f. The field lines have therefore the differential equation:
(1) fxdr = 0.
If f = f1i + f2j + f3k, (1) is equivalent to the system,
dxdydz
(2)
fl f2 f3

Any integral u(x, y, z) = a of this system represents a surface locus


of field lines. For, from du = dr Vu = 0 and (1), we have
(3)
and, since Vu is normal to the surface u = a, the vector f at any
point of the surface is in the tangent plane.
If v(x, y, z) = b is a second integral of (2), we have also
(4)
If v is independent of u, Vu x Vv 0 0 (§ 87, theorem 1). From (3)
and (4) we conclude that f is parallel to Vu x Vv, and hence
(5) Vu x Vv = Xf.
104 FIELD LINES OF A VECTOR, 227

But the curve in which the surfaces u = a, v = b intersect is every-


where tangent to Vu X Vv. Hence we have the
THEOREM 1. If u = a and v = b are independent integrals of the
system f x dr = 0, the surfaces they represent intersect in the field lines
of f.
From (3) and (4), we see that u and v are independent solutions
of the partial-differential equation:
aw aw caw
(6) f- Vw =flax + f2 a + f3 az = 0-
ax
In view of (5), this equation is equivalent to
(7) Vu X Vv Vw = 0.
From § 87, theorem 3, (7) is satisfied when and only when u, v and
w are connected by a functional relation. Hence we have
THEOREM 2. The general solution of the partial differential equa-
tion f Vw = 0 is w = cp(u, v), where cp is an arbitrary function and
u = a, v = b are independent integrals of f X dr = 0.
When the vector field f(r) is solenoidal, div f = 0. In § 92 we
found that we could express f as rot g. We now deduce another
form for f which gives at once its field lines.
THEOREM 3. A solenoidal vector f which is continuously differ-
entiable can be expressed in the form:
(8) f = Vu X Vv.

Proof. If we assume the relation (8), we have f Vu = 0,


f Vv = 0; thus both u and v are solutions of f Vw = 0. If we
choose for u some integral of the system (2), f Vu = 0. Now,
if dr is a differential on the level surface u = a., we have, from (8),
(9) f X dr = Vv du - Vu dv = - Vu dv.
In the function u(x, y, z), at least one variable is actually present.

(10) v= -
If z is present, au/az = uz is not identically zero; hence, on multi-
plying (9) by k , we have k f X dr = -uz dv and
kxf dr
uz
,
228 INTEGRAL TRANSFORMATIONS § 104

the integral being taken over a curve on the surface u = a. This


integral is independent of the path; for, from (85.3) and (85.6),
kxf 1\ 1
rot = O - J X (kXf) + -rot (kXf)
uz uz U,

Of
2(Vuz)X(kxf) -k
uZ uZ

-(f V i )k + (k Vu,)f f2
2
ui uz

kXf fZ - cu -1 a
Vu -rot = _-- (f - Vu) = 0.
uZ uZ uZ az
If u(x, y, z) contains x or y, we may replace k and uz in (10) by
i and ux or j and uv.
With the values of it and v thus obtained, we now have, from
theorem 2 of § 103,
(k x f)c -1
Vu X Vv = Vu x Grad v = - V it X = - Vu x (k X f) = f,
uZ uZ

where Grad v refers to the surface u = a, and (k X f)t is a tangen-


tial projection upon it. Moreover f - Vv = 0.
The field lines of the vector Vu X Vv are the curves in which the
surfaces u = a, v = b intersect.
Example 1. The field lines of the vector,
f = xzi + yzj + xyk,
have the differential equations:
dx dy dz
(i)

From these, we obtain the equations,


ydx-xdy=0, y dx + x dy - 2z dz = 0,
which have the integrals:
y/x = a, xy - z2 = b.
These one-parameter families of surfaces intersect in the two-parameter family
of field lines.
§ 104 FIELD LINES OF A VECTOR 229

When one integral, as y/x = a, is known, we may find a second by obtain-


ing the field lines on the surface y/x = a. These must satisfy the equation
obtained from (i) by putting y = ax, namely, ax dx = z dz. Its integral
ax2 - z2 = b gives the field lines on the surface y/x = a. On replacing a by
y/x, we obtain a second family of surfaces xy - z2 = b which intersects the
first family y/x = a in the field lines.
Example 2. The field lines of the vector,
f = xi + 2x 2j + (y + z)k,
have the differential equations:
dx dy dz
(ii)
x 2x2 y+z
From dy = 2x dx, we obtain the family of surfaces,
y-x2=a.
As no other integrable combination is evident, we put y = a + x2 in (ii); then
dx _ dz dz z a
x a+x2+z' or dx-x=x-I-x.
This linear equation, with the integrating factor 11x, has the solution,
z = -a+x2+bx.
A second family of integral surfaces is therefore
z +Y
- 2x = b.
x
Example 3. The vector,
f = x(y - z)i + y(z - x)j + z(x - y)k,
is solenoidal; for div f = 0. In order to express f as Vu x Vv we choose for u
an integral of the system,
dx _ dy _ dz
x(y - z) y(z - x) z(x - y)
Since the sum of the denominators is zero,
dx+dy+dz=0, x+y+z=a;
hence we may take
u = x + y + Z.

We now use (10) to compute v. On the surface u = a, z = a - x - y, and


-kxf dr = -x(y -z)dy+y(z - x)dx
= y(a-2x-y)dx+x(a-x-2y)dy.
230 INTEGRAL TRANSFORMATIONS § 105

Since this is an exact differential, the method of § 91 gives

v= f(aY_2xY_Ydx=axY_?J_x?J
= xy(a - x - y) = xyz.
With u = x + y + z, v = xyz, we may readily verify that f = Vu x vv.
105. Pfaff 's Problem. Since rot f is solenoidal, we have from
theorem 3 of § 104,
(1) rot f = Vu x Vv = rot (u Vv),
rot (f - uVv) = 0;
hence (§ 91) f - uVv is the gradient Vw of a scalar, and
(2) f = Vw + uvv.
When f is a continuously differentiable vector, we may find three
scalars u, v, w so that (2) holds good. The determination of these
scalar functions is known as Pfaff's Problem. We proceed to give
a simple and direct solution.
Assuming the truth of (2), we at once deduce (1); hence
(3) Vu rot f = 0, Vv rot f = 0,
so that both It and v satisfy the same partial differential equation :
V. rot f = 0. Let v = a be some integral of the system dr x rot f
= 0; then Vv rot f = 0. Now on the surface v = a we have,
from (2),

this integral, taken over a curve of the surface v = a, is inde-


pendent of the path since Vv rot f = 0. On substituting the func-
tions v and w thus obtained in (2), this equation uniquely deter-
mines u.
We now must show that, with It, v, w thus determined, Vw +
uVv = f. Let Grad w and n denote the gradient and unit normal
on a surface of the family v = const; then (103.4)
dw dw
Vw=Grad w+ndn
ft+ndn
§ 105 PFAFF'S PROBLEM 231

and the projection of Vw + uVv tangential to the surface is ft.


Moreover u was chosen to make the normal projections of Vw +
uVv and f the same, that is,
dw
n-+uVv=f-ft.
do
From (1) and (2),
a (u, v, w)
f- rot f =
C )(X' y, z)

hence f rot f = 0 implies that u., v, w are functionally dependent


(§ 87, theorem 3). In this case,
(4) f dr = dw -i-- u(v, w) dv = 0
is an ordinary differential equation which admits solutions under
general conditions-as when u(v, w) and au/caw are continuous in
a region R. * Hence the condition
(5) f-rot f=0
shown in § 91, ex. 2, to be necessary for the integrability of f dr
= 0, is also sufficient.
When (5) is fulfilled, let (p(w, v) = C be the general solution of
(4).Then
amp
d(p = dw + - dv = 0
aw av

must yield (4) upon division by app/8w. In other words, X =


a,p/8w is an integrating factor of (4) :

Then Xf = VV, and f is everywhere normal to the surfaces p =


const. The field lines of f are then the orthogonal trajectories of
these surfaces.
A family of curves is said to form a field in a region R if just
one curve of the family passes through every point of R. The unit
vectors T, N) B along the curves are then vector point functions
in R; and the first Frenet Formula dT/ds = KN can be written
(6) T- VT = KN.
* See Agnew, Differential Equations, New York, 1942, p. 310, et seq.
232 INTEGRAL TRANSFORMATIONS § 105

Now, from (85.7) and (85.9),

TT=
T = -KN.
For a surface-normal field
T rot T = 0, (T x rot T) X T = rot T,
and, from (7), we have
(8) rot T = KB, rot T I = K.
The Darboux vector of a surface-normal field of curves is there-
fore 8 = rT + rot T.
Example. Find u, v, w so that
f = [2yz, zx, 3xy] = Vw + u Vv.
Solution. rot f = [2x, -y, -z]; since the system,
dx _ dy _ dz
2x -y -z
has the solution z/y = a, we take v = z/y. On the surface v = a, z = ay;
hence
w =ff dr = f (2yz dx + zx dy + 3xy dz)

= f (2ay2 dx + 4axy dy)

2axy2 = 2xyz.
Now u is given by
[2yz, zx, 3xy] = [2yz, 2zx, 2xy] + u[0, -z/y2, 1/y],
whence u = xy2. Thus our solution is
u = xy2, v = z/y, w = 2xyz.
Since f U. rot f = 0, u, v, and w must be functionally related; in fact w = 2uv.
The total differential equation,
(i) 2yzdx+zxdy+3xydz =0,
is thus equivalent to
dw + udv = 3udv + 2vdu = 0,
and has the integral u9 = C, or x2yz3 = C.
106 REDUCTION OF VOLUME TO SURFACE INTEGRALS 233

106. Reduction of Volume to Surface Integrals. Let S be a


closed surface and V the volume it encloses. Then S has two
sides, an inside and an outside, and at all regular points of S we
have a definite unit normal n directed towards the outside. We
shall suppose that S consists of a finite number of parts over which
n is continuous.
Consider, first, a surface S which is cut in at most two points
by a line parallel to the z-axis; denote them by (x, y, z1) and
(x, y, z2), where zl < z2. Then S has a lower portion Sl con-
z

*_y

FIG. 106

silting of all the points (x, y, z1), and an upper portion S2 con-
sisting of the points (x, y, z2). We suppose also that the points
for which zl = z2 form a closed curve separating Sl from S2 (Fig.
106). The equations of Sl and S2 may be taken as z = zl(x, y),
z = z2(x, y).
Now let f (x, y, z) be a tensor function of valence v whose scalar
components have continuous partial derivatives throughout V.
Then, if the volume integral of of/8z over V is written as a triple
integral with the element of volume dV = dx dy dz, we may effect
a first integration with respect to z,

(i) ff az dx dy dz

= fff ( x, y, z2) dx dy -fff (x, y, z1) dx dy,

whe re the double integrals are taken over the common projection
A of Sl and S2 on the xy-plane.
234 INTEGRAL TRANSFORMATIONS §106

If we regard x, y as the parameters u, v on the surfaces S1 and


S2, the vector element of surface is ry x ry dx dy (94.9). Now from
r = ix + jy + kz(x, y), rz = i + kzz, ry = j -}- kzy,

Over S2 (z = z2) this vector has the direction of the external nor-
mal n; but over Sl (z = zl) it has the direction of the internal
normal -n. Hence, if we denote the vector element of area in
the direction of the external normal by dS = n dS,
n dS = ± (k - izz - jzy) dx dy, k . n dS = fdx dy,
where the plus sign applies to S2 and the minus to S1. The two
integrals over A now may be combined into a single integral over
8, so that we may write

(1)
j/,f dx dy dz =f I

This formula is also valid when S is bounded laterally by a part


of a cylinder parallel to the z-axis and separating Sl from S2. For
(i) holds as before; and, in (1), k n = 0 over the cylinder, so that
it contributes nothing to the integral over S.
We now may remove the condition that S is cut in only two
points by a line parallel to the z-axis. For, if we divide V into
parts bounded by surfaces which do satisfy this condition and
apply formula (1) to each point and add the results, the volume
integrals will combine to the left member of (1); the surface inte-
grals over the boundaries between the parts cancel (for each ap-
pears twice but with opposed values of n), whereas the remaining
surface integrals combine to the right member df (1).
Finally we may extend (1) to regions bounded by two or more
closed surfaces, that is, regions with cavities in them, by this same
process of subdivision. Additional surfaces must be introduced so
that the parts of V are all bounded by a single closed surface, and
the surface integrals over these will cancel in pairs as before.
When x, y, z form a dextral system of axes, the same is true of
y, z, x and z, x, y. Hence, if in (1) we make cyclic interchanges in
x, y, z, we obtain the corresponding formulas:

(2) fff a dx dy dz = y, z) i n dS,


§ 106 REDUCTION OF VOLUME TO SURFACE INTEGRALS 235

(3) fff ay dx dy dz = ff f(x, y, z) j n dS.

If we insert the prefactors k, i, j in the integrands of (1), (2), and


(3), respectively, add the resulting equations, and note that

iof +jof +kof = Vf,


ax ay az
we obtain finally
(4) f of d V =in f dS,
using J( to denote integration over a closed surface. The inte-
grands are tensors of valence v + 1. We have thus proved
BASic THEOREM II. If f(r) is a continuously differentiable tensor
point function over the region V bounded by a closed surface S, whose
unit external normal n is sectionally continuous, then the integral of
of over the volume V is equal to the integral of of over the surface S.
If f(r) is a vector or tensor of higher valence, we may obtain
from (5) other integral transformations by placing a dot or a cross
between the first two vectors in each term of the integrands; for
both of these operations are distributive with respect to addition
and therefore may be carried out under the integral sign. We thus
obtain the important formulas:

(5) fv.fdv=Jn.fds,
(6) fv<fdv = JnxfdS.

integral
surface.
f
When f is a vector, (5) is known as the divergence theorem; the
n f dS then is called the normal flux of f through the

Example. If rot f = 0 in a simply connected origin, f = 102), and

ffdV =f = fncodS.

Thus the volume integral of an irrotational vector can be expressed a surface


integral over the boundary.
236 INTEGRAL TRANSFORMATIONS § 107

For example, the center of mass of a homogeneous body of mass m is fixed by


the position vector,
r*=mfrdm=yfrdV;
for m = pV, where p denotes the constant density. Since rot r = 0, we have,
from (102.2),
r
r=V, where P= e
U

Hence
Vr* = i 0 nr2 dS.

107. Solid Angle. The rays from a point 0 through the points
of a closed curve generate a cone; and the surface of a unit sphere
about 0 intercepted by this cone is called the solid angle & of the
cone.
The reciprocal of the distance r = OP is a harmonic function
(§ 89, ex. 2) ; hence
1 1 R
div grad
r
= 0, and grad
r
= - r2
is a solenoidal vector.
Let us now apply the divergence theorem to the vector f = R/r2
in the region interior to a cone of solid angle 0 and limited exter-
nally by a surface S, internally by a small sphere a about 0 of
radius a. Within this region f is continuously differentiable,
div f = 0 and fn f dS = 0. The external normal n = - R over
0', and over the conical surface n R = 0; hence
a R 1 Sa
(1) J r2 dS =,a2 dS = a2 = S2,
where Sa is the area cut from or by the cone. Note that the ratio
Sa/a2 is independent of the radius and may be computed with
a=1.
If S is a closed surface, we have
nR 14r, 0 inside of S
(2)
2
dS =
r l0, 0 outside of S.
When 0 is outside of S, f is continuously differentiable throughout
its interior, and the foregoing result is immediate. In this case
§ 108 GREEN'S IDENTITIES 237

the elements of solid angle,


(3) d1t = 2
dS,
r

corresponding to the same ray cancel in pairs.


108. Green's Identities. We now apply the divergence theorem
(106.5) to the vector f = p0j,, where <p and ' are scalar functions
having continuous derivatives of the first and second order, re-
spectively, in a region R bounded by a closed surface S. Now
div (,p0') _ V p V4, + p div V ' (85.2),
n d¢/dn,
where the normal derivative d¢/dn is in the direction of the external
normal to the bounding surface. Hence, on writing the operator
div grad as V2, we have

(1) V p V. V4, dV + f g' V24,dV do dS.

In case ' is not defined outside of S, we replace d ,/dn by the


negative of the derivative along the internal normal -n. Formula
(1) is known as Green's first identity.
If both cp and 4, have continuous derivatives of the first and
second orders, we may interchange (p and 4, in (1). On subtract-
ing this result from (1), we obtain Green's second identity:

(2) f(v2G - 4'V2


v) dV =
dd4,
\ do - 41 dnl dS.
We now take ,' = 1/r, where r is the distance OP. If 0 is inte-
rior to S, we cannot apply (2) to the entire region enclosed since
4, becomes infinite at 0. We therefore exclude 0 by surrounding
it by a small sphere v of radius e and apply (2) to the region R'
between S and v; then
1
V24pdV =
dl 1dipdS.
dl ldc dS- cP-----
co-----
JS \ dn r r dn J, dn r r dn
On the sphere r = e,
d d 1 1 dp
(1l (1l dS = eQ,
2d
do \ r/ dr \ r) r2 e2 ' do ar '
238 INTEGRAL TRANSFORMATIONS § 108

dil denoting the solid angle subtended by dS. The integral over v
is therefore

J(o E2 + 1
E ate)
ar dS
dSt + E f a
ar
M= Ef a

or
dg,

where p is a value of <p at some point of o. Now, as e -* 0,


d1 1
dS -
do r r do

-v2(p dV-- -f-v2(P dV


R1r

Rr

j
for the integrand remains finite if we use the spherical element of
volume dV = r2 sin 0 dr dB d4p. We thus obtain Grreen's third iden-
tity:
(3) Jr v2r dV + i dip
\r do
d it
do r/
dS.

When 0 is exterior to S we may put ¢ = 1/r directly in (2); in


this case the right member of (3) equals zero.
We may deduce three analogous identities from (101.6),

fDivfdA
the divergence theorem in the plane. With f = (p Grad ¢, we find

(4) fv.vdA+fv2dA = f (P do ds,


where we now interpret V and V2 as Grad and Div Grad and
di/dn as a derivative in the direction of the external normal m.
By an interchange of cp and , we find as before
d dip
f(cv24, - ,'tv2 p) dA
(5) = /C do - # d.n) ds.

To obtain the third identity from (5), we take 4, = log p, where


p is the distance OP in the plane. Since log p is a solution of
Laplace's Equation in the plane,
Div Grad # = 0,
§ 109 HARMONIC FUNCTIONS 239

v24, = 0 in (5). As before, we must exclude the origin from the


region by surrounding it by a small circle y of radius E. Applying
(5) to the region remaining, we have

- flog p
\
v2cp dA =
R
d d J
log p - log p ds + ,. (P do log p - log p ds.
c do do do
On the circle p = E,
d d 1 1 dip aP
do
log p
dp
log p
p E ' do
=
- ap ' ds = E d6;

the integral over y is therefore

where
f\ log E - - -
ap
E dB = E log c-

is a value of p at some point of y. When e -> 0, E log t


acv
- d8 - 27r gyp,
ap

--* 0, and we obtain


d duo
(0) 27r p(0) = log p V2,p dA + (`p log p - log p -) ds.
R C do /
109. Harmonic Functions. A solution of Laplace's Equation,
(1) div grad cp = 0,
is called a harmonic function. A function <p is said to be harmonic
in a region R if it has continuous derivatives of the first and
second orders and V2p = 0 in R.
If a vector f is both irrotational and solenoidal, its scalar poten-
tial 4, is harmonic; for, from (102.4),
f = -Vi, divf = -V2' = 0.
If in Green's second identity (108.2) cp and 4, are harmonic
throughout R, we have

(2)
f\odn - dn/dS=0.

Thus, if '= 1, we have, for any harmonic function cp,

(3) do dS = 0.
240 INTEGRAL TRANSFORMATIONS § 109

If cp is harmonic in the region bounded by a closed surface S,


Green's third identity (108.3) gives the value v(P) at any interior
point P,
(4)
v l dS,
is \r do - do r/
where r = PQ, the distance from P to points Q on the surface.
Thus a function gyp, harmonic in the region enclosed by S, is deter-
mined completely at any interior point P by the values of <p and
dcp/dn on the boundary.
When P is exterior to the surface S, we have

(5) 0 - .Js (r do do r) dS'


on putting i = 11r in (2).
If S is a sphere of radius r about P as center and lying entirely
within the region R, we have, from (4),

4irv(P) =
or, in view of (3),
r f do dS
+ r2 JCp dS;
--
(6) f p dS.
r
c'(P) = 4r2
Since the surface of the sphere is 47rr2, we have the
MEAN VALUE THEOREM. The value of a function, harmonic in a
region R, at any point P is equal to the mean of its values on any
sphere about P as center and lying entirely within R.
This theorem shows that a function which is harmonic in a
closed bounded region R, but not constant, attains its extreme
values only on the boundary. For let P be a frontier point of the
set for which cp attains its minimum value m. If P were an interior
point of R, there would be a sphere about P, lying within R, on
which V > m at some points; hence the mean value of c over the
sphere would be greater than (p(P) = m.
If in Green's first identity (108.1) we take cp = 4' and assume
that (p is harmonic in R, we have

(7) fi o1 2 dV
=
- is do
dS.
§ 110 ELECTRIC POINT CHARGES 241

Hence, if either cp = 0 or dcp/dn = 0 on S, the volume integral in


(7) vanishes; and, since I nip 12 is continuous and never negative,
we must conclude that V p = 0, and cp has a constant value through-
out R. If 'p = 0 on S, 'p = 0 in R; and, if d'p/dn = 0 on S, p = C
in R.
Now let 'i and 'P2 be two harmonic functions in R; then their
difference <p = V, - 'P2 is also harmonic. If 01 = 'P2 on S, cp = 0
on S and cpl = 'p2 throughout R. If d'pl/dn = dSP2/dn on S, dcp/dn
= 0 on S and <pl = IP2 + C throughout R.
If a harmonic function 'p has a constant value C on S, 'p = C
throughout R; for the harmonic function cp - C is 0 on S.
110. Electric Point Charges. By Coulomb's Law, an electric
charge el at 0 exerts a force of

(1) F = el2 R f
r2

upon a charge e2 at P; r = OP, and R is a unit vector in the direc-


tion OP. Thus the charges repel when elel > 0 (charges of same
sign) and attract when ele2 < 0. The force exerted by a charge e
at 0 upon a unit charge at P, namely,
e
(2) E = a R,
r

is called the electric intensity at P due to the charge e. Since


R = Vr,
e
(3) E _ -V-,
r

so that E has the scalar potential (102.4),

(4)

The vector E is both irrotational and solenoidal; for


1
(5) rot E = 0, div E _ -eV2 - = 0.
r
t In a vacuum, if the charges are measured in statcoulombs and the distance
in centimeters, the force is given in dynes.
242 INTEGRAL TRANSFORMATIONS § III

If S is a closed surface, we have, from (107.2),


nR
(6) n E dS = e
J r2
dS =
147re
0
0 inside of S,
0 outside of S.
We assume that the intensity due to a system of point charges
e1, e2, , e,, is the sum of their separate intensities: E = EEi.
Since Ei = - Vei/ri, where ri is the distance from the charge ei
to P,
E _ - -+-+ e1

r1
e2

r2 r,
en/

and the potential of the system is


el e2 e
(7)
r1 r2 rn

If this system of charges is within the closed surface S, we have,


from (6),
(8) Jn.EdS = 47r1ei.
The normal flux of the electric intensity through a closed surface is
equal to 4ir times the sum of the enclosed charges.
111. Surface Charges. If a surface S carries a distributed
charge a per unit of area, the potential at a point P due to the
charged element dS at Q (regarded as a point charge a dS) is
a dS/r, where r is the distance QP. If we assume that the surface
density a is continuous or piecewise continuous over S, the total
potential at P is
dS

The electric intensity at a point P outside of S, due to the


charge element a dS at Q, is -a dS Vp 1/r; the total electric in-
tensity at P is therefore

(2) E fov-r dS = -Vpp.


The notation Op means that P must be varied in computing the
gradient. Since P is outside of S, 1/r and its first partial deriva-
3 112 DOUBLETS AND DOUBLE LAYERS 243

tives are continuous for all positions of Q on S; hence in computing


E = -Vpcp, we may differentiate (1) under the integral sign.
In differentiating functions of r = PQ, we may vary either P
or Q, holding the other point fixed. Thus, if R is a unit vector in
-4
the direction PQ,
VQr = R, Vpr = - R,
when Q and P, respectively, are varied; and, in general,
(3) Vpf(r) = f'(r)Vpr = -f'(r) VQr = -VQf(r)
Consequently, we also may write (2) as

(2)' E_ 1 dS.
r
In integrals, such as this, the subscript on V may be omitted on
the understanding that the variation is at dS.
From (1) and (2), we see that cp and E are continuous at all
points P not on the surface. At such points cp is harmonic; for
/'
vp"p = I crop dS = 0.
s r

At a point P on the surface, the integrals for <p and E are im-
proper since r passes through zero. However it can be shown t
that if v is piecewise continuous on S, p is defined on S and is
everywhere continuous. Moreover, as P approaches a surface
point Q from the positive side (toward which n points) or the nega-
tive side, under general conditions E approaches limiting values
E+ and E_, such that
(4) E+ - E_ = 47rv n,
where o- and n are the surface density and unit normal at Q. Thus
the normal component of E experiences a jump of 4irv as P passes
through the surface.
112. Doublets and Double Layers. The potential at P due to
a charge -e at Q and a charge +e at Q' is (110.7)
-e e //1 1

QQ'PG r
t See O. D Kellogg, Foundations of Potential Theory, Berlin, 1929, Chapter
3, §5.
244 INTEGRAL TRANSFORMATIONS § 112

If Q' approaches Q and at the same time the charges increase, so


that the product,
e QQ' = m,
remains constant, the limiting result is called a doublet of moment m.
The potential of this doublet is
1 1

1
(1) P=QimQje(QQ')PQQQ,PQ} =m - OQ - ;
frallction r
for the limit of the in the second member is the directional
derivative of 1/r in the direction of m.
A continuous distribution of doublets over a surface with mo-
ments everywhere in the direction of the normal n is called a
double layer. If /in dS is the moment of the doublet at the surface
element dS at Q, the potential of the double layer at an outside
point P is
1
(2) _
fs r
where r = PQ. Since
1
dS=dSl
r r2

is the solid angle subtended by dS (107.3), we also may write

(3) 1P _ -f s
Ada

When µ is constant, this reduces to -µf2, where Sl is the total solid


angle subtended by S at P.
When µ, the moment density, t is piecewise continuous, p is con-
tinuous at all points P not on the surface. At such points cp is
harmonic; for

rs
At a point Q where µ is continuous and the surface has contin-
uous curvature, it can be shown * that cp has definite limits cp+, (p_
t In the case of a magnetic shell, u is called the density of magnetization.
* See 0. D. Kellogg, op. cit., Chapter 6, § 6.
113 SPACE CHARGES 245

according as P approaches Q from the positive or negative side of


the surface, and that
(4) p+ - Cpl = 4rµ.
At a point P outside of S, the electric intensity E = -Vpca is
continuously differentiable. Since pn is a function of Q (not P),
we have, from (2),
1
(5) E cep = - µn CQ VP dS.
fs r
But, from (85.6),
GP X n x G'Q (-1
\ =
-n VP6Q - ,
1

r
so that we also may write
(6) E = Vp x 1 µn x VQ -1 dS.
r s
Consequently the intensity due to a double layer has, besides the
scalar potential gyp, also a vector potential A (§ 92) :
(7) E = rot pA,

(8) A = fLn VQ dS.


r
When ja is constant, A may be transformed into a circuit integral
about the boundary of S; for, from the basic theorem (100.1),

(9) A=A fnxV


s r cr
Example. Let (P be a function harmonic in a region bounded by a closed
surface S; then, at any interior point P (109.4),

w(P) dS d (1
47rr do r/ dti.

Comparison with (111.1) and (112.2) shows that p may be regarded as the
potential due to a surface change of density a = (d'/dn)/4a and a double
layer of moment density IA = -c/4,r.

113. Space Charges. If a region V carries a distributed charge


p per unit of volume, the potential at a point P due to the charged
element dV at Q (regarded as a point charge p dV) is p dV/r, where
246 INTEGRAL TRANSFORMATIONS § 113

r is the distance QP. If we assume that volume density p is piece-


wise continuous over V, the total potential at a point P outside
of V is

(1) (P
= I aT

The electric intensity at P due to the charge element p dV at


Q is -p dV Vp 1/r; the total electric intensity at P is therefore

(2) E fp
X ©p dV = - Vpp.
Since P lies outside of V, 1/r and its first partial derivatives are
continuous for all positions of Q within V; hence, in computing
Vpcp, we may differentiate (1) under the integral sign.
When P lies within the charged region V, the integrals for
and E are improper since r passes through zero. But it can be
shown ¶ that, when p is piecewise continuous, (p and E exist at the
points of V and are continuous throughout space. Moreover the
potential p is everywhere differentiable, and E = - Vptp.
The equation (110.8) also may be proved for space charges;
namely,
(3) fn.EdS = 47r f p dV,

the integral on the right covering all space charges within S. The
closed surface S may either completely enclose the charged region
V or cut through it. If S encloses no charges, the integral (3) is
zero.
When p is continuously differentiable, it can be shown that E
has the same property. Now if Sl is any closed surface enclosing
a subregion Vl of V, the divergence theorem shows that

fn.EdS = fdiv E dV.


Yi
But, from (3),
fdiv Ed V = 47r J pdV;
Yi

¶ See O. D. Kellogg, op. cit., Chapter 6, § 3.


§ 114 HEAT CONDUCTION 247

and, since this holds for any subregion V1 of V,


(4) div E = 47rp,
or, if we put E
(5) -4Trp.
This partial-differential equation is called Poisson's Equation. At
points outside of the charged region V, p = 0, and the potential P
satisfies Laplace's Equation:
(6) v2 P = 0.
114. Heat Conduction. In mathematical physics the integral
theorems often are used in setting up differential equations. As
an illustration, consider the flow of heat at a point P of a body.
According to Fourier's Law, the direction of flow is normal to the
isothermal surface through P; and the flow F (calories per second)
per unit of surface is proportional to the temperature gradient at P.
Thus F may be represented by the vector,
(1) F = -k VT cal./sec./cm.2,
where T is the temperature and k the thermal conductivity of the
body at P. Since k is positive, the minus sign in (1) expresses the
fact that heat flows in the direction of decreasing temperature.
Let p and c denote the density and specific heat of the body at P.
Then the rate at which heat is being absorbed in a region R bounded
by a closed surface S is
a

at
cpT dV = fcp aT
at
dV.

If n is the outward unit normal to the surface S, the rate at which


heat flows into R through S is

- in FdS = f div (kVT) dV.

Hence, if heat is being generated in R at the rate of h calories per


unit of volume,

fcp!dva=f {div (kVT) + h} dV.


248 INTEGRAL TRANSFORMATIONS § 115

Since this holds for an arbitrary region R within the body,


OT
(2) cp - = div (kVT) + h.*
at
This is the differential equation of heat conduction.
If the body is homogeneous, k is constant, and (1) becomes
aT
(3) cp = k V2T + h;
at
and, if there is no internal generation of heat, we have
aT
(4) = K V2 T,
at
where K = k/cp is called by Kelvin the "diffusivity."
When the heat flow becomes steady the temperature distribu-
tion is constant in time. Hence in the steady state (4) reduces to
Laplace's Equation V2T = 0.
Example. Find the temperature distribution in a homogeneous hollow
sphere whose inner and outer surfaces are held at constant temperatures.
When the flow is steady, T is a function of r alone; hence, from (89.17),

V 2T r dr (r2T r), dr
(r2)__O, T= r + B.
A and B may be determined from the conditions T = Tl when r = rl, T = T2
when r = r2.

115. Summary : Integral Transformations. If f (r) is a continu-


ously differentiable tensor point function, the basic integral trans-
formations are:

(A) fvfdv=fnfds ( f dV Vf =ids f);


(B) fri x Vf dS = fTf ds ( f dS x Vf = fdr f).
* If f(r) is a continuous scalar function in a region V and ff(r) dV = 0
for an arbitrary subregion of V, then f(r) = 0 throughout V. For, if f(r) 96 0
at a point P, we can surround P with a sphere a so small that f does not change
sign in a, and hence f f (r) dV 0, contrary to hypothesis.
0
§ 115 SUMMARY: INTEGRAL TRANSFORMATIONS 249

In addition,
J"T2ofds=
(C)
l
Jrl

may be regarded as a third basic type. From these, other integral


transformations may be deduced by dotting and crossing within
tensor functions. If we dot and cross in the first position, we get,
from (A) :
fv.fdv=fn.fds,
(Ax) fvxfdv =inxfdS.
Similarly, from (B),
fn.vxfds=fT.f,
(Bx) f(n? Vf - nV -f) dS = fT f ds,

in which V may be replaced by V8. In the B transformations the


surface integrals vanish when taken over a closed surface.
When f is a vector function, (A - ) is known as the divergence
theorem and (B - ) as Stokes' Theorem.
If we replace f by of in (B x) we obtain

(B') f (osf + J nf) dS = fm f ds,


a transformation of the same scope as (B); J = - Div n is the
mean curvature of the surface, and m = T x n is the unit external
normal to the bounding circuit. From (B') we derive

(B'.) f(vs.f-i-Jn.f)ds=fm.fds,
(BI X) f(Vsxf+Jllxf)dS = f mxfds.
On a plane, n is constant, and J = 0; if we write dA for dS, the
B' transformations become

(P) fV8f dA = fm f ds,


250 INTEGRAL TRANSFORMATIONS

fv3.fdA
(Px) fvsxfdA=fmxf(1s.
When f is a vector, (P - ) is the divergence theorem in the plane.
If f (r) is a continuously differentiable vector and rot f = 0 in a
simply connected region of 3-space, the line integral of f - dr is in-
dependent of the path, and
r
f=V , p(r) = f dr.
ro

If f (r) is a continuously differentiable vector, and a rot f = 0


in a simply connected portion of a surface, the line integral of
f - dr over surface curves is independent of the path, and
rf . =frf, .
ft = Grad (p, (r) = dr dr,
ra

where ft = f - (n f)n is the tangential projection of f on the


surface.
PROBLEMS
1. Show that a closed curve lying in a plane with unit normal a encloses an
area A given by
nA = frxdr
2. Compute the integral / n rot f dS over that portion of the sphere
r - a lying above the xy-plane when f = p(r)c.
3. Prove that J'r x n dS = 0 over any closed surface S. If a body bounded
by S is subjected to a uniform pressure -pn per unit of area, show that these
forces have zero moment about any point in space.
4. If the closed curve C encloses a portion of a surface S, show that

I n x r dS = I JTr2ds.

5. If f = ia(x, y) + jv(x, y), prove that

ff x dr = kf f div f dx dy

where C is closed curve in the xy-plane enclosing the region A.


PROBLEMS 251

6. If c is a constant vector, prove that

nx(cxr)dS=2Vc
is
where V is the volume enclosed by the surface S.
7. The closed surface S encloses a volume V. If the vector f is everywhere
normal to S, prove that rot f dV = 0.
V

8. Show that the centroid of a volume V bounded by a closed surface S


is given by
Vr* = z _fr2n dS.

Apply this formula to find the centroid of a hemisphere of radius a, center


0, lying above the xy-plane.
9. If p denotes the distance from the center of the ellipsoid r 4) r = 1
to the tangent plane at the point r, show that the integrals of p and 1/p over
its surface have the values

fPds = 3V, IdS/p = cv1V,


where V is the enclosed volume [p = r n, 1/p = r 4) n]. Express these
results in terms of the semiaxes a, b, c of the ellipsoid.
10. Find a vector v = f(r)r such that div v = rm (in 96 -3). Prove that

frmdv rmr n dS.


11. Show that

(a) rrmR dV = - 1 rm+in dS (m76 -1);


M

R
(b)
r
dV = flog r n dS.
12. If p and q are vector functions, prove that

(a)

13. If r = PQ, the solid angle subtended at P by the surface S (over which
Q ranges) is
i
Slp = - n VQ - dS (107.1),
is r
252 INTEGRAL TRANSFORMATIONS
where V0l /r is computed at Q. Prove that

Vp12p
1TXVQdS
r ra

where r = PQ and the curve C is the boundary of S.


14. If C and C' are two unlinked closed curves, show that if r = PP' the
double-circuit integral,
r TxT'
ds (is' = 0 .
a
r

If C and C' are two simple loops, linked as in a chain, show that the preceding
double integral is ±4a, the sign depending on the sense of C'.
15. Show that tangential forces of constant magnitude o- acting along a closed
plane curve C are equivalent to a couple of moment 2QA, where A is the vector
area enclosed by C. What is the moment of the couple when C is a twisted
curve? [Put f = r in (100.4).]
16. Let f(r) be a solenoidal vector function. Prove that f = rot g (§ 92)
where
t
"OD

g = -r X fxi(,r) dX or r x j Xf(ar) dX
o t
provided the integrals exist. (The integrand of the former may become infi-
nite at the lower limit.)
In particular if f(r) is homogeneous of degree n 0 -2, f(,\r) = X"f(r) and
g is given by g = (f x r) /(n + 2). Prove this formula directly by making use
of r Vf = of and (85.6).
CHAPTER VII

HYDRODYNAMICS

116. Stress Dyadic. Let S be any closed surface inside a de-


formable body. It divides the body into two portions, A within S,
and B without S. The forces acting on A are of two kinds: (i)
body or mass forces which act on the interior particles, and (ii)
surface forces, which act on the bounding surface S.
Let k denote the average body force on the element of mass
Am; then, if Am shrinks to zero while always enclosing a point P,
the limit of P/Om is defined as the body force R, per unit of mass,
at P.
Let F,, denote the average surface force acting on A over the
vector element of surface nOS of the boundary S; then, if AS
shrinks to zero while always enclosing a point P, the limit of
Fn/AS is defined as the surface force F, per unit of area, acting
on A at P. Here n denotes a directed line normal to S at P and
directed from A to B. The notation Fn thus associates a surface
force with a surface element of unit normal n. The force F dS is
exerted on the element n dS by the matter toward which n points.
The surface force acting on B on the same element n dS is, from
the law of action and reaction,
(1) F_n = -Fn.
Consider now a small tetrahedron (Fig. 116) with three faces
parallel to the coordinate planes; and let outwardly directed lines
normal to the faces be denoted by -x, -y, -z, n. If the area
of the inclined face is A, the faces normal to x, y, z have areas
A cos (n, x), A cos (n, y), A cos (n, z), and the volume of the
tetrahedron is 3Ah where h is the altitude of P above the base A.
If body and surface forces are replaced by averages and p denotes
the average density, the equilibrium of the tetrahedron requires
that
F,,A + F_IA cos (n, x) + F_NA cos (n, y)
+ F_ZA cos (n, z) + 3Ahpr2 = 0.
253
254 HYDRODYNAMICS §I II

If we divide out A and let h ---. 0 so that the tetrahedron shrink,


to the vertex P, we have in the limit
F,, + F_x cos (n, x) + F_ cos (n, y) + F_Z cos (n, z) = 0.
If n is a unit vector along the line n, cos (n, x) = n i, etc.; and,
since F_x = -Fr, we have
(2) Fn = n (i Fx + i Fy + k Fe).
All surface forces in (2) now refer to the point P. Thus a plan(
through P normal to n divides the body into two parts; and thE

-x
Fla. 116

part toward which n points acts upon the other part with the fort(
F,, per unit of surface at P. The dyadic.,
(3) c= i Fx + j Fy + k F=,
is called the stress dyadic at P; it effects a synthesis of all surfac(
forces at P by means of the relation,
(4)

If the deformable body is a fluid, we assume as an experimenta


fact that the stress across any surface element of a fluid in equilibrium
is a pressure normal to the element; hence
Fn = -pn; F. = - 711, Fv = - p2J, Fz = - p3k.
With these values, (2) becomes
pa=nipli-f-nj P2j+nkp3k;
117 EQUILIBRIUM OF A DEFORMABLE BODY 255

and, on multiplying by i , j , k , in turn, we have


(5) p=P1=P2=P3
At any point within a fluid the pressure is the same in all directions.
At any point where the pressure is p, the stress dyadic is
(6) `, = -p(ii + ii + kk) = -pI.
For fluids in motion it is no longer true that only normal stresses
exist. Viscous fluids in motion do exert tangential stresses; but in
many problems these tangential stresses are small and may be
neglected. We shall develop the mechanics of fluids on the hypoth-
esis of purely normal stress. We imply this assumption by speak-
ing of perfect or non-viscous fluids; for a perfect fluid the stress
dyadic is -pI.
117. Equilibrium of a Deformable Body. If a body in a strained
state is in equilibrium, any portion of it bounded by a closed sur-
face S is in equilibrium under its body forces and the surface forces
acting on S. Let the body forces be R per unit of mass, and let the
surface forces Fn per unit of area be given by the stress dyadic 4.
Then, from (116.4), F,, = n 4, and the equations of equilibrium are

(1) fRdv+fn.ds=o ,

(2) J'r x Rp dV + jr x (n (b) dS = 0,

where (2) is the moment equation about the origin.


If we transform the surface integrals by means of (106.5),

in - -1) dS = f v 4dV,

frx(n.4)ds= xrdS= -
(1) and (2) become
f(R + V )dV=0,

f[PRxr+ V.(xr)]dV = 0.
256 HYDRODYNAMICS § 117

Since these integrals vanish for an arbitrary choice of S, their inte-


grands must be identically zero. The equations of equilibrium are,
therefore
(3) =0,
(4) pRxr+V.(cxr) =0.
On eliminating pR from these equations, we obtain
(5) V.(4)xr) - (V -4)) -r = 0.
From (5) we conclude that 4), the vector invariant of c, is zero;
for, if we write 4) = i F,z + j Fy + k F2,
a
(Fxxr) +a (F,-r) +az
a
(FZxr)
-
(aFx a yy aFZ\
ax y ax + a + az xr-0,
Fx'< +Fyx-+F2xar=
Fxxi+Fyxj+FLxk=0,
ax ay az
and hence 4) = 0. In view of the theorem of § 68, we see that the
moment equation (2) requires that the stress dyadic (D be symmetric.
If we write
Fx = i Xx -f - j Yx + k Zx, . . .
the stress dyadic may be written in the nonion form,
Xx Yx Zx
(6) c = Xy Y, Zy ,

XZ YZ ZZ
where, by virtue of its symmetry,
(7) Xy = Yx, YZ = Zy, Zx = XZ.
The normal components of stress Xx, Yy, ZZ occur in the principal
diagonal, whereas the tangential components, or components of
shear, are equal in pairs. More specifically, in perpendicular
planes, the components of shear perpendicular to their line of inter-
section are equal.
Since 4) is symmetric, we always can find three mutually or-
thogonal vectors i, j, k, so that
(8) 4) = a ii + j3 jj + y kk
Then i, j, k give the principal directions of stress, and a, are
the principal stresses at the point in question.
§ 119 FLOATING BODY 257

118. Equilibrium of a Fluid. For a fluid the stress dyadic is


4_ -pI and -Vp,
from (86.8); the equation of equilibrium (117.3) is therefore
(1) Vp = pR.
Consider a liquid of constant density p in equilibrium under
gravity. The body force per unit mass is then g, the acceleration
of gravity, and
(2) Op = pg = pgk = pg Oz = V (pgz)
when the z-axis is directed downward along a plumb line Hence
(3) p = p.9z + po,
where p = po when z = 0. If the origin is at a free surface, po is
the atmospheric pressure, and pgz is called the hydrostatic pressure.
119. Floating Body. Let the sur-
face S of a floating body V be di-
vided by its plane section A at the
water line into two parts: S1 sub-
merged, S2 in air (Fig. 119). If po is
the atmospheric pressure, pi = pgz
the hydrostatic pressure at depth
z, po + pi acts on S, po on S2; or
we may say that po acts over S,
while p1 acts on the closed surface Fia. 119
S1 + A (pl = 0 on A) enclosing the
volume V1 below the water line. Hence the equations of equi.
librium of the body are:
(1) fgdm -ispondS -i+ p1ndS = 0,
is
(2) frxgdrn - f
S
A

rxnpodS - f rnpl dS = 0.
S,+A
We consider in turn the integrals in (1) :
fgdm = W, the weight of the body;

fpondS = po in dS = 0;
258 HYDRODYNAMICS § 120

n pl dS = Vpi dV = fpg dV = W1,

the weight of the displaced liquid. Therefore (1) states that


W = W1i this is the
PRINCIPLE OF ARCHIMEDES: A floating body in equilibrium dis-
places its own weight of liquid.
In order to interpret (2), we note that the center of mass of a
discrete set of particles Pi of mass mis defined as a centroid (9.3);
its position vector r* is given by
(3) mr* =
where m is the total mass. Similarly, the center of mass of a
continuous body is defined by
-(4) mr* = f r dm.

If we change signs throughout in (2), we have the following


integrals to consider:

f v
gxrdm = gx f v
rdm = gxmr* = Wxr*,

ojnxrdS = po Ifrot r dV = 0,
s s

n x pir dS = Trot (p1r) dV = j(Opl) x r dV


s+A vt v

=gx pr dV = g x m1r= W1 x r.
vt
Thus (2) reduces to W x (r* - r) = 0; this states that the center
of mass of the body and of the displaced liquid are on the same vertical.
120. Equation of Continuity. Let p and v denote the density
and velocity of a fluid at the point P and at the instant t. Con-
sider the mass of fluid dV within a fixed but arbitrary closed
surface S. This mass is increasing at the rate,
a ap
dV =
at J at dV'
§ 120 EQUATION OF CONTINUITY 259

the time differentiation being local. This rate must equal the rate
at which fluid is entering S, namely, - in pv dS, where n de-
notes the outward unit normal. Hence

Iapat d V = - fn pv dS = - fdiv (pv) dV,

when the divergence theorem (106.5) is applied. Since the inte-


gral of ap/at + div (pv) over any closed region within the fluid is
zero, we conclude that
ap
(1)
at + div (pv) = 0.

This equation of continuity may be put in another form by intro,


ducing the substantial rate of change dp/dt instead of the local time
rate Op/at. Along the actual path, or line of motion, of a fluid
particle,
x = x(t), y = y(t), z = z(t),

the density p(t, x, y, z) becomes a function of t alone, and


dp OP OP dx ap dy ap dz
dt at + Ox dt + ay dt + az dt

OP OP ap
+
at +° \l OP
+ k OZ
gives the time rate at which the density of a moving fluid particle
is changing. We thus obtain the important relation,
dp OP
(2) + v - op'
dt at

connecting the substantial and local changes of p.


For any tensor function f(t, r) of time and position, associated
with a fluid particle moving with the velocity v, we have, in the
same way,
df of
(3) +v Of.
dt at
260 HYDRODYNAMICS § 121

From (85.2),
div (pv) = v - Vp + p div v;

hence the equation of continuity (1) also may be written


dp
(4) dt+pdivv=0.

Since div v = - (dp/dt)/p we can interpret div v as the relative


time rate of decrease of density of a fluid particle having the
velocity v. Thus a positive value of div v implies a negative
dp/dt and, consequently, an attenuation of the fluid at the point
considered; hence the term divergence.
For an incompressible fluid, dp/dt = 0, and
(5) div v = 0.

This equation has the integral equivalent,

(6) in - v dS = fdiV v dV = 0;

the "flux" of an incompressible fluid across the boundary of a


fixed closed surface within the fluid is zero. If an incompressible
fluid is also homogeneous, p is constant.
121. Eulerian Equation for a Fluid in Motion. In the Eulerian
or statistical method of treatment we aim at finding the velocity,
density, and pressure (v, p, p) of the fluid as functions of the time
at all points of space (r) occupied by the fluid. Consider the fluid
within a fixed closed surface S at any instant t. By D'Alembert's
Principle, the body and surface forces, together with the reversed
inertia forces (-ma), may be treated as a system in statical equi-
librium. In a perfect fluid the surface force is a normal pressure:
Fn = -pn (§ 116). Hence, if the body force per unit mass is de-
noted by R, D'Alembert's Principle applied to the fluid enclosed
by S gives the equation,

(1) f(R_a)Pdv_fnpds=o,
(2) frx(R - a)pdV - f rXn pdS = 0,
5 121 EULERIAN EQUATION FOR A FLUID IN MOTION 261

where a denotes the acceleration of the fluid particles. On trans-


forming the surface integrals in (1) and (2), we have

f {(R-a)p-Vp}dV=0,
f {r x (R - a)p + rot (pr) } dV = 0.

Since these integrals vanish for any choice of S, the integrands are
identically zero. From (85.3), rot (pr) = (Vp) x r; the equations
of motion are therefore
(3) (R - a)p - Vp = 0,
(4) rx(R-a)p-rxVp=0.
When (3) holds good, (4) is identically satisfied and may be
omitted. The Eulerian Equation of Motion is therefore
1
(5) a=R--Vp.
P

Here a = dv/dt is the acceleration of a moving fluid particle and


must be distinguished from av/at, the rate of change of fluid ve-
locity at a fixed point. These substantial and local rates of change
are connected by the relation (120.3) :
dv av
+v - Vv.
(6) a dt at
When the density p is a function of p only, we introduce the
function:
P dp dP 1
(7) P = fe - ; then VP = - Vp = - Vp.
P dp P

Moreover, if the body force R has a single-valued scalar potential


Q (§ 102),
(8) R = - VQ;
such forces are said to be conservative. Under these conditions
(5) becomes
dv
(9) = - v (Q + P).
dt
262 HYDRODYNAMICS § 122

In the Eulerian method r and t are independent variables, so


that ar/at = 0. Equation (120.3) applied to r,
dr or

dt
I=v,
is simply an identity.
The.lines of the fluid which at any instant are everywhere tan-
gent to v are called its stream-lines. They are not the actual paths
of the fluid particles except when the flow is steady, that is, when
v is constant in time (av/at = 0). The stream-lines have the dif-
ferential equation v x dr = 0.
Example. Revolving Fluid. If the fluid is revolving with constant angular
velocity w = wk about a vertical axis, its velocity distribution is that of a
rigid body; hence, with the origin on the axis of revolution, we have, from
-4
(54.3), vp = w x OP, or simply v = w x r. Then a = w x v, and (9) becomes
(i) wx(wxr) = -V(Q+P).
Now
co x (w x r) = r - (ww - w2I)
- 2v{r (w2I - ww) r} (85.14)

- 1 w2 v(r2 - z2)
- 2w2 0(x2 + y2);
hence, from (i),
V[Q + P - 2W2(x2 + y2)1= 0,

Q + P - 21w2(x2 + y2) = cont.


For gravitational body forces, R = g, and Q = gz, if the z-axis is directed
upward; and, if the density is constant, P = p/p. In this case, (ii) becomes
gz + p/p - iw2(x2 + y2) = cont.
At a free surface p is constant; a free surface is therefore a paraboloid of
revolution.

122. Vorticity. Starting with the Eulerian Equation in the form.


av
vv = - V (Q + P),
at + v
(1)

we transform v Vv by means of (85.9) and (85.7) :


v Vv = (Vv) v - v x rot v = 2 V(v v) - v x rot v;
§ 123 LAGRANGIAN EQUATION OF MOTION 263

hence (1) may be written


av
(2) - - vxroty = -o(Q + P + 2v2).
at
Now, from (84.11),

rot av
-- vxroty = 0,
at
arot V
= rot (v x rot v)
at =
(rot v) Vv - v - V rot v - (rot v) (div v),
when we make use of (85.6) and (84.13) ; and, since
arot y drot y
at dt '
drot y
(3) dt + (rot v) (div v) = (rot v) Vv.

When v gives the velocities of a rigid body having the instan-


taneous angular velocity w, rot v = 2w (§ 84, ex.). For a liquid
we may regard
(4) w=yrot V
as the molecular rotation or vorticity of the fluid particles.
In (3) we now replace rot v by 2w; and, from the equation of
continuity (120.4),
dp d /1\
(5) + p diV v =0, diV v = p
dt dt p
Then (3), after division by p, becomes
ldw d(1\ w
p dt
+wdt\p/ = P
vv,

d (w _ co

This is known as Helmholtz's Equation.


123. Lagrangian Equation of Motion. In the Lagrangian or his-
torical method of dealing with a moving fluid, the motion of the
individual fluid particles is followed from their initial positions ro
264 HYDRODYNAMICS § 123

to their position r after a time t. Thus the history of each fluid


particle is traced. If the function f(r) is associated with a fluid
particle, r is a function of the independent variables ro and t.
In space differentiation we may form gradients relative to r or
ro; these are denoted by the symbols V and vo. Thus we may
compute df (§ 90) as either
df = dro vo f or df = dr Vf;
and, since
(1) dr = dro vor,

and
drovof = dro Dor of
for arbitrary displacements dro, we have
(2) vol = vor vf.
In particular, when f = ro, we have
(3) I = vor vro,
so that the dyadics vor and vro are reciprocal. Consequently, if
we multiply (2) by vro as prefactor, we have
(4) of = vro vof
An element of fluid volume dVo is altered by the transformation
(1) in the ratio J/1, where J denotes the third scalar invariant of
vro (§ 70) : thus
(5) dV = J dVo.
If we use rectangular coordinates,
axo/ax ayo/ax azo/ax

(6) vro = axo/ay ayo/ay azo/ay

axo/az ayo/az azo/az

and J is the determinant of this matrix, namely, the Jacobian


a(xo, yo, zo)/a(x, y, z). The element of mass Po dVo becomes
p dV = pJ dVo; and, since mass is conserved,
d
(7) PJ = Po or (pJ) = 0.
dt
This is the equation of continuity in the Lagrangian method. Since
§ 123 LAGRANGIAN EQUATION OF MOTION 265

dp/dt = -p div v from (120.4),


d (dJ
J div v = 0,
dt (PJ) = P dt
and (7) is equivalent to
dJ
(8) - = J div v.
dt
The Dynamical Equation of Lagrange corresponding to the
Eulerian Equation (121.9) is obtained by multiplying the latter
by Vor as prefactor; thus
dv
(9) Vor
dt
= -Vor V(Q + P) _ -oo(Q + P),
from (2). Since ro and t are the independent variables, the sym-
bols Vo and d/dt commute; hence the left member of (9) may be
written
d d
(Vor) v} - (Vov) v = { (cor) v} - vozv2,
-{ dt
and (9) becomes

(10) ( (oor) v} = - Vo(Q + P-


d
zv2).

With the aid of the dyadic Vor we may integrate Helmholtz's


Equation (122.6). If we make use of (4), this becomes
d/ w\ w w d w d
dt p p p dt p dt
) Vor,
since, from (3), the time derivative of Vro Vor is zero. Multiply-
ing this equation by Oro as postfactor now gives
d
'
Oro + p fro/ - 0,
dt \p10 / \dt orb/ dt \p
w (00
- Vro = -,
P PO

the constant wo/Po being the value of (w/p) Vro when t = 0.


Multiplication by Vor gives finally

P PO

This is Cauchy's integral of Helmholtz's Equation.


266 HYDRODYNAMICS 124

To verify this integral, we need only differentiate (11) with re-


spect to t: thus
d(w = wo
pr = 0
Ov=
coo
G'r
P P0 Po Po P

The lines of a fluid which are everywhere tangent to w are


called its vortex-lines.Their differential equation is cox dr = 0.
Consider a vortex line at the instant t = 0, and let its differential
equation be wo x dro = 0. After a time t, coo and dro become
P
w= coo Vor, dr = dro Vor.
Po

Hence, if coo = X dro, then co = (p/po) X dr; that is, wo X dro = 0


implies w X dr = 0. Thus we have proved the
THEOREM. If the body forces have a potential and p = f (p) or
constant, the vortex-lines move with the fluid. Vortex-lines always
consist of the same fluid particles.
124. Flow and Circulation. The tangential line integral of the
velocity of a fluid along any path is called the flow along that path.
If the path is closed, the flow is called the circulation.
If the path AB at time t was AoBo when t = 0, the flow over
AB is

f B
v dr f dro
and, since ro and t are independent variables,
Ao
Bo
(Vor) v;

f v6 dr = jOdro { (Vor) v
a

When the body forces have a potential Q and p = f(p), we have,


from (123.10),
J.B Bp
dr= - Ao
vo(Q-- I+ ' -v)
Bo

fo
LBdr V( Q+P-2v2);
§ 125 IRROTATIONAL MOTION 267

or, since the last integrand is a perfect differential,


d
(1) -- fv.dr= [-Q - P + v2]and,

in particular,
d
(2)
dt
fv.dr=0.
The last result is Kelvin's
CIRCULATION THEOREM. If the body forces have a potential and
p = f(p) or constant, the circulation over any closed curve moving
with the fluid does not alter with the time.
125. Irrotational Motion. When the body forces have a poten-
tial and p = f(p) or constant, Cauchy's Equation (123.11) shows
that, if the vorticity of a fluid vanishes at any instant, it will
remain zero thereafter. Then rot v = 0 in space and time, and
the motion is termed irrotational. In any simply connected por-
tion of the fluid, we may express v as the gradient of a scalar
(§ 102) ; thus
fr
(1) v = -Vs , P=- v dr,
ro

and (p is called the velocity potential. Then the velocity is every-


where normal to the equipotential surfaces (p = const and is di-
rected toward decreasing potentials. Hence, along any line of
motion, cp continually decreases; in a simply connected region the
lines of motion cannot form closed curves.
By Stokes' Theorem "the circulation iv dr = 0 over any re-
ducible curve (§ 102) ; and, as this curve moves with the fluid, the
circulation around it remains zero (circulation theorem, § 124).
From this fact we again may deduce that, if the motion is irrota-
tional at any instant, it remains irrotational thereafter.
The equation of continuity (120.4) now becomes

(2) p02(p = 0.
dt -
For an incompressible fluid, dp/dt = 0, and V2(P = 0; then <p is a
harmonic function.
268 HYDRODYNAMICS § 126

The equation of motion (122.2) reduces to

(3) v (Q +P+ 2v2


- ail = 0,
since av/at = - V (a(p/at) ; for a local time differentiation, a/at (r
constant) and a space differentiation V (t constant) commute with
each other. Hence

(4) Q+P+1 2
- at = f (t)
an arbitrary function of the time; and, if the flow is steady,
Q + P + 2 v2 is an absolute constant.
Every harmonic function cp represents some irrotational flow of
an incompressible liquid whose velocity v = -VV. The problem
consists in finding a solution of V2V = 0 which conforms to the
given conditions. For example, in a flow with central symmetry,
<p must be a function of r alone; hence, from (89.17),

v2 V 1d 2 d`pl
r2dr Cr drJl
0, r dr=b-,r
2 =a,
d`p a

Now v = aR/r2 where R is a unit radial vector; and the flux per
second through any sphere of radius r about the origin is constant:
47rr2 (a/r2) = 47ra. When this flux is given, a is determined. The
value of b is immaterial, since it disappears in v = - vcp; as it is
customary to have <p vanish at infinity, we choose b = 0.
126. Steady Motion. A flow is said to be steady when it is in-
variable in time. Then all local time derivatives are zero. Thus
ap/ at = 0, and the equation of continuity (120.1) is simply
(1) div (pv) = 0.

Also av/at = 0 and the stream-lines are also lines of motion. The
Eulerian Equation (122.2) now becomes
(2) vxroty = v(Q+P+1v2).
If T is a unit tangent along a stream-line (T x v = 0) or a vortex-
line (TX rot v = 0), T v X rot v = 0; hence T V(Q + P + 2v2) _
0, or
d
(3) (Q + P + Zv2) = 0 along a stream- or vortex-line.
ds
3126 STEADY MOTION 269

We thus have proved


BERNOULLI'S THEOREM. Let the body forces have a potential and
P = f(p) or constant; then, for a steady flow,
(3) Q + P + v2 = const
z
along any 'stream-line or vortex-line.
The constant will vary, in general, from one line to another.
However, if the motion is irrotational as well as steady,
V(Q+P+zv2) = 0,
(4) Q+P+zv2=C,
where C is an absolute constant-the same throughout the fluid.
Now suppose that the fluid has a constant density p; this is
nearly fulfilled in the case of a liquid. The equation of continuity
is div v = 0 or fn - v dS = 0. Along any tube whose surface
consists of stream-lines, let the normal cross section be denoted
by A. If the tube is sufficiently thin, and we apply -f n - v dS
= 0 to the portion between A 1 and A2, we have approximately
v1A1 - v2A2 = 0, or vA = const
as the equation of continuity along a tube of flow.
If the liquid is subject only to gravitational body forces, their
potential is gz, where z is measured upward from a horizontal
reference plane. Thus with
P=p/P, Q=gz,
we have

(5) gz + p + v2 = const
P 2
along a stream-line. If z, p, v are known at the section A0
(zo, po, vo), we have, from (5),
(6) Pg(z-Z)+(p-po)+zP(v2 -vo)=0.
For a thin tube we may take vA = v0A0, and (6) becomes
2
(7) p - po = Pg(zo - z) - Pvo C2 - 1) .

2
270 HYDRODYNAMICS § 127

Thus the pressure is least at the narrowest part of the tube.


Example. Torricelli's Law. When liquid escapes from an orifice near the
bottom of a vessel which is kept filled to a constant level, the flow may be
regarded as steady. Consider a stream tube extending from the orifice of the
area A to the upper surface where its area is Ao. At this surface PO is the
atmospheric pressure, and vo = vA/Ao; at the orifice p = po, and the outflow
speed is v. With these values we have, from (6),
v2(1 - A2/A0) = 2g(zo - z) = 2gh,
where h is the distance from the free surface to the orifice. When Ao is large
compared with A, we have approximately v2 = 2gh, a result known as Torri-
celli's Law.

127. Plane Motion. When the flow is the same in all planes
parallel to a fixed plane it and the velocity has no component
normal to Tr, the motion is said to be plane. Such motion is com-
pletely determined by the motion in 7r, which may be taken as
the xy-plane. Velocity, density, and pressure are all functions of
x and y alone; and, for any tensor function f(x, y), we have, from
(97.1),
Of
grad f = Grad f + k - = Grad f;
49Z

and, similarly,
div f = Div f, rot f = Rot f.
The flux across any curve C in the xy-plane is defined as the
volume of liquid per second crossing a right cylindrical surface of
unit height based on C. For an observer
who travels the curve in the positive sense
of s, the flux crossing C from (his) right to
left is

(1) fkxT.vds=fvxk.dr.
x here k, T, k x T are unit vectors along the
FIG. 127 z-axis, the tangent and normal to the curve
(Fig. 127).
In any simply connected portion of the fluid the flux across a
plane curve joining two points will be independent of the path,
provided
k rot (v x k) = 0 (§ 103, theorem 1).
§ 127 PLANE MOTION 271

From (85.6),
rot (v x k) = k - Vv - k div v = -k div v.
For an incompressible fluid, div v = 0 (120.5) and F will be inde-
pendent of the path. Hence
(2) ¢(r)
ro

defines a scalar point function; and, from § 103, theorem 2,


Grad 4, = v x k, since v x k lies in the xy-plane. Writing V for
Grad, we have
(3) V,k=vxk, v=kxVik.
V,' is everywhere normal and v everywhere tangent to the
curves t = const; these curves are therefore stream-lines (§ 121).
Consequently, the function is called the stream function.
If we set g = -4k, we have
(4) rot g=kxV,,=v, divg=0;
thus the velocity has the vector potential -,pk (§ 92).
From (85.6), we have
(5) rot v = rot (k x k V2VI.

Therefore the plane motion of an incompressible fluid will be irrota-


tional when and only when the stream function is harmonic:
(6) V2W = 0.
In this case v = - Vq: the velocity has a scalar potential (p; and,
since
(7) V2(p= -dive=0,
the velocity potential is also harmonic.
From
v= -V<p=kxV#,
we have the relations,
(8) Vp = (VP) x k, VL. = k x Vsp;
or, in terms of rectangular coordinates,

i --=i--j-
ax
3
app

ay
a,k
ay
491P

ax
272 HYDRODYNAMICS 5 127

Thus (8) is equivalent to the equations:


at 49 (P alp
(9)
ax ay ' ay ax
But these are precisely the Cauchy-Riemann Equations which
connect the real and imaginary parts of the analytic function
w = <p + i>y of the complex variable z = x + iy. We thus have
proved the important
THEOREM. In any plane, irrotational motion of an incompres-
sible fluid, the velocity potential p and the stream function ' are two
harmonic functions which combine into an analytic function V + i¢
of a complex variable x + iy.
Since -,y + icp = i1G), we see that, if cp + iy is analytic,
-4, + i<p is also; consequently, if cp and ' are the velocity potential
and stream function for an irrotational plane flow, -4, and V are
the corresponding functions for another flow of this type.
From (8), we have Ocp 0¢ = 0: the stream-lines (4i = const)
cut the equipotential lines (,p = const) at right angles.
Since the complex potential w = cp + i>G is an analytic function
of z,
a
w'=-=-+i-=--i-;
dw app .9,P ag'
dz ax ax ax ay
or, since the velocity has the components,
ag app
vx -1 Vc , vY -j Vc ,
ax ay
(10) w' _ -vz+ivy, - w' = vx+ivy)
where w' denotes the conjugate of w'. Thus the velocity at any
point is given by the complex vector -w'; its magnitude is I w' 1.
Example 1. Assume the complex potential w = azn (a real); then, if we
write z = rete,
w = arner'ne = arn(COS n9 + i sin no);
,p = amcosn9, 4, = amsinn9.
The stream-lines are the curves whose polar equations are rn sin n9 = const.
For the cases n = 1, 2, -1, we put z = x + iy, using rectangular coordi-
nates.
(a) n = 1: w = az = ax + iay.
§ 128 KUTTA-JOUKOWSKY FORMULAS 273

The stream-lines are the lines y = const, and the flow has the constant ve-
locity -t3' = -a in the direction of -x.
(b) n = 2: w = az2 = a(x2 - y2) + i2axy.
The equipotential and stream-lines are the two families of equilateral hyper-
bolas,
x2 - y2 = const, xy = const.

Since the stream-line xy = 0 may be taken as the positive halves of the x-axis
and y-axis, these may be considered as fixed boundaries and the motion re-
garded as a steady flow of liquid in the angle between two perpendicular
walls. The velocity at any point is -zb' _ -2a2; its magnitude varies directly
as the distance from the origin.
(c) n = -1: w = a/z = a(x - iy)/(x2 + y2).
The equipotential and stream-lines are two families of circles:
x/(x2 + y2) = const, y/(x2 + y2) = const,
tangent to y-axis and x-axis, respectively, at the origin. The velocity -w'
= a/22 becomes infinite at the origin.
Example 2. With the complex potential,
w = V(z +a2/Z) = V(reie + a2r 'e-'°), (V real),
a2
= V r + a2
r
cos B, >G= V r-a2a2r sin 0.
The stream-line 4, = 0 includes the circle r = a and the x-axis sin 0 = 0.
Since w' = V(1 - a2/z2), the complex velocity,
/ -2b'=-V(1-a22/-.-V
as z --+ oo.

Therefore we may regard the motion as a flow to the left about an infinite
cylindrical obstacle of radius a. At a great distance from the obstacle, the
flow has a sensibly uniform velocity - Vi.t
When body forces are neglected, the Bernoulli Equation (126.4) gives
p/p + 1v2 = const. From the symmetry of the flow about the cylinder, it is
clear that the total pressure exerted by the fluid on the cylinder is zero. We
shall consider this matter for cylinders of arbitrary section in the next article.
128. Kutta-Joukowsky Formulas. Consider an incompressible
fluid flowing past an infinite cylindrical obstacle of arbitrary cross
section. The flow, in a plane perpendicular to the generators of
the cylinder, is assumed to have the sensibly uniform velocity v
at a great distance from the obstacle. Then, if the motion is
t See Lamb, Hydrodynamics, Cambridge, 1916, p. 75, for the stream-lines.
274 HYDRODYNAMICS § 128

steady and irrotational, and the body forces on the fluid are neg-
lected, we have, from (126.4),
(1) p=K --v - v,
where K is an absolute constant.
If n is a unit internal normal to the boundary C of the obstacle
(Fig. 128),
(2) F =jnpds, M =frxnpds
c c
give the resultant force and moment about the origin exerted by
Ay

0 x
FIG. 128

fluid pressure on a unit length of cylinder. Substituting p from


(1) in these integrals and noting that

in ds = 0, Jr x n ds = J rot r dA = 0,
from (101.5), (101.7), we have

(3) F=
2 J
When the flow is given by the complex potential w = + i4,, we
shall compute these integrals after converting them into circuit
integrals in the complex z-plane.
The vectors r = xi + yj and r = xi - yj correspond to z and
its conjugate 2:
r z = x + iy = r(cos d + i sin 0) = re'6,
7 2 = x - iy = r(cos 0 - i sin 0) = re-t0.
§ 128 KUTTA-JOUKOWSKY FORMULAS 275

By definition,
r1 r2 = r1r2 cos (02 - 01), r1 x r2 = rlr2 sin (02 - 01)k;
hence

Thus r1
r 1 r2 ' it 1 x rk =
2 12
r2 and r1 x r2 k correspond to the real and imaginary
rre,(B2--B,)
-= 202-
parts of 21z2:
(4) r1 r2 - (R(21x2) = 2 (21z2 + x122),
(5) r1 x r2 . k -9(202) = 2 (21z2 -
Note also that k x r ti iz; for the multiplications by k x and
i (= eia/2) both revolve a vector in the xy-plane through 7r/2.
When r1 and r2 are perpendicular or parallel we have, respec-
tively,
(6) r1 r2 = 0 - 21z2 + z122 = 0,

(7) rl x r2 = 0 - 21z2 - z122 = 0.

Turning now to the integrals (3), we have the internal normal


n = k x T for a counterclockwise circuit of C; hence
nds = kxTds = kxdr-idz.
Moreover, if v'' v, then v v ,..' v v, and

c
The moment M, normal to the plane, is completely specified by
the scalar,
-2 P

c
and, since
k r x (k x T) ds = (k x r) (k x T) ds = r dr '-' 01(2 dz),

M P v v 0? (2 dz) vi 2 dz.
2 c 2
P G1 c
t The real and imaginary parts of any complex number z are (R(z) _
2(z + 2) and 4(z) = 2(z - 2). Note also that the conjugate of a product is
the product of the conjugates: z = 2122.
276 HYDRODYNAMICS § 128

But v and dz are parallel along C, which forms part of a stream,


line; hence v dz = v d2, from (7), and we have

F = -
P
i
JV2 dz, M dz
2 2 c
for the corresponding complex force and moment. Finally we form
F and replace v2z dz in M by its conjugate v2z dz; thus

(8)
P
j J02 dz= P
-z dw 2dz,
2JC 2 c dz

P (dw 2
v2
(9) M = - 2 6l z dz = - a? dz) z dz,
C 2 C \
since v = -dw/dz (127.10). These integrals exist if dw/dz remains
finite over C.
Now, outside of C, v is an analytic function which approaches
v at infinity. The Laurent series for v therefore has the form:
(10) 2' = P. + al/z + a2/z2 + a3/z3 + .. .
If C. is a large circle of radius R about the origin and enclosing C-
we have, by Cauchy's Integral Theorem,

ff(z) dz = ff(z) dz, if f (z) is analytic between C and C .$

Thus the counterclockwise circulation about C is

y= dr ti i J(P dz - v dz) dz dz.


fcv c cW

Moreover the static moment of the circulation is

µ= fry dr-fzvdz = fz3dz. C

If we replace v by the series in (10), all integrals vanish except


those involving
21r
dz Reie i dB
JC. z - Reie - i f dB = 2ri.
t See, for example, Franklin, A Treatise on Advanced Calculus, New York,
1940, §§ 267, 268.
§ 128 KUTTA-JOUKOWSKY FORMULAS 277

We thus find
(11) y = 27ri al, µ = 27ri a2.
We now can compute the integrals in (8) and (9) in terms of
v,,, y, and /2. From (10), we have
v2
= v2 + 2a1v /z + (2a25 + a12 ) /z2 + .. .
When this series is substituted in (8) and (9), all integrals vanish
except fdz/z = 2rri; hence
(12) F=2i ipyv,,, F = -ipyv,,;

(13) M = - 2 6?(2a2v + a2l)2-7ri = -pR?()u0)


since a2 = -y2/4r is real, and ff?(2iri a2) = 0. In vector notation
these give
(14) F = -pykxv,,,
(15) M= -pL xF,
y
for the resultant force and moment on a unit length of cylinder.
From (15) we see that F acts through the point,

(16) µ/y = Jr v dr/ Jv - dr,


c
which may be called the centroid of the circulation.
The equations (14) and (15) are called the Kutta-Joukowsky
Formulas. From (14) we see that, for a counterclockwise circula-
tion (y > 0), the force F is upward if the flow is horizontal and to
the left (v = - Vi); we then have a lift of pyVj per unit length
of cylinder. The counterclock circulation diminishes the flow ve-
locity below the cylinder and increases it above; by Bernoulli's
Theorem this results in an excess of pressure below with a conse-
quent upward lift.
In the absence of circulation about the cylinder (y = 0), F = 0,
and there can be no lift. This is the case in § 127, ex. 2, where
w = V(z + a2/z), v = -dw/dz = - V(1 - a2/z2),
and a1 = 0 in the Laurent series.
278 HYDRODYNAMICS § 129

129. Summary: Hydrodynamics. For a perfect (non-viscous)


fluid the stress dyadic is -pI; at any point within a fluid the
pressure is the same in all directions and exerted normal to a sur-
face element.
If f (t, r) is any tensor function associated with a fluid particle
moving with the velocity v, its substantial rate of change is
df of
Vf,
dt at + V
where of/8t is the local rate of change.
In the Eulerian method of dealing with fluid motion, the aim is
to compute v, p, and p as functions of the independent variables
r, t. The (kinematic) equation of continuity is
8p dp
+div(pv) = +pdivv=0;
at dt

and, for an incompressible fluid (dp/dt = 0), becomes div v = 0.


The Eulerian Equation of Motion is

(E) dt=-V(Q+P), P=f dp,


when the body forces have a potential Q and the density a func-
tion of p only. From this we can deduce the Differential Equa-
tion of Helmholtz for the vorticity co = 1 rot v:

Vv.
dt \p/ = P

In the Lagrangian method the aim is to follow the motion of the


fluid particles from their initial positions ro (t = 0) to their posi-
tions r after a time t. The independent variables are now ro and
t; the position r of a particle at time t is a function of ro and t.
We may take gradients relative to ro(V0) or r(V); and these con-
form to the relations,
Vof = Vor Of, Vf = Vro Vof; Vor Vro = I.
If J is the third scalar of the dyadic Vro, the equation of continuity
becomes
pJ = po, or dJ/dt = J div v.
§ 129 SUMMARY: HYDRODYNAMICS 279

The Lagrangian Equation of Motion corresponding to the Eulerian


Equation (E) is
dv
(L) Vor = - Vo(Q + P).
dt
From this we deduce Cauchy's integral of Helmholtz's Equation,
_=-.Vor;
(0 0)0

P Po

the fluid. We find, moreover, that the circulation f


and this in turn shows that vortex-lines (co x dr = 0) move with
v dr over
any closed curve moving with the fluid does not alter with the time.
When the motion is irrotational (rot v = 0), v = - VV, where
p is the velocity potential, and iv dr = 0 over any reducible
curve. When the body forces are conservative and p is a function
of p alone,
Q+P+ Zv2
- a = f (t), an arbitrary function of t.
For a steady-motion equation (E) gives
vxroty=V(Q+P+Zv2);
whence Bernoulli's Theorem: Q + P + Zv2 is constant along a
stream-line or vortex-line, and this constant is absolute if the
steady motion is also irrotational.
For an incompressible fluid in plane motion, the integral inde-
pendent of the path,

¢(r) = x k dr (k I. plane of motion),

defines the stream function. The curves 4, = const are the stream-
lines (v x dr = 0). If the motion is also irrotational, both stream
function and velocity potential are harmonic (V2# = 0, V2V = 0),
and V4, = k x VV. This is the vector equivalent of the Cauchy-
Riemann Equations which guarantee that w = p + iii is an ana-
lytic function of the complex variable z = x + iy in the plane of
motion. The complex potential w therefore has a unique derivative
dw/dz; and the negative of its conjugate gives the complex velocity
vector: v = -w'.
280 HYDRODYNAMICS

PROBLEMS
1. A mass of liquid is revolving about a vertical axis with the angular speed
f(r), where r is the perpendicular distance from the axis. With cylindrical
coordinates r, 0, z (we replace p, p in § 89, ex. 1 by r, 0 to avoid conflict with
the notation of Chapter VII), let a, b, c be unit vectors in the directions of
r,., rg, rZ (Fig. 89a). If the angular velocity w = f(r)c, prove that

v = rf(r)b, rot v = r dr [ref (r)]c. [Cf. (89.7).]

2. If the motion in Problem 1 is irrotational, show that


a a
PC, v=rb,
and that the velocity potential v = - ao is not single valued (a, $ are
constants).
For a liquid of constant density p under the action of gravity alone, show
that the pressure is given by
gz+p+-const.
3. If a fluid is bounded by a fixed surface F(r) = 0, show that the fluid must
satisfy the boundary condition v VF = 0.
More generally, if the bounding surface F(r, t) = 0 varies with the time,
show that the fluid satisfies the boundary condition

F + v 7F = 0 [Cf. (120.3).]

4. A sphere of radius a is moving in a fluid with the constant velocity u.


Show that at the surface of the sphere the velocity of the fluid satisfies the
condition
!v-u) (r - ut) = 0.
5. A fluid flows through a thin tube of variable cross section A. Show that
for a tube PoP of length s
a
J8PA ds + pAv 0;

hence deduce the equation of continuity

at (pA) + as (pAv) = 0.

6. From the equation of fluid equilibrium (118.1) show that when p is a


function of p alone, rot R = 0 and that the potential of R is -P.
If p is not a function of p alone, show that equilibrium is only possible when
R R. rot R = 0. [Cf. (121.7) for P.]
PROBLEMS 281

7. A gas flows from a reservoir in which the pressure and density are po, po
into a space where the pressure is p. If the expansion takes place adiabatically,
p/py = const. (y is the ratio of specific heats), show that
P= y p
y-1p
Neglecting body forces and the velocity of the gas in the reservoir, show that
the velocity v of efflux when the motion becomes steady is given by
y=1
v2 = 2y 7?0
1 - (p/po) 7 } .
y - 1 PO

The velocity of sound in a gas is c = 1'yp/p; hence show that


v2 _ 2
- 1 (c0 - c2).
y

8. If a body of liquid rotates as a whole from r = 0 to r = a with the con-


stant angular velocity coo, and rotates irrotationally with the angular velocity
to = woa2/r2 when r = a, we have the so-called "combined vortex" of Rankine.
If z = za at the free surface when r = a, show that the free surface is given by

Z = za.+-(r2-a2) r 5 a,
a2`
Z = Z. +w2a2
- 1! 1 - 2J , r a.
2 \ r
Prove that the bottom of the vortex (r = 0) is a distance w2a2/g below the
general level (r = oc) of the liquid.
9. If the vorticity is constant throughout an incompressible fluid, prove
that V2v = 0. [Cf. (84.14).]
10. When the motion of an incompressible fluid is steady, deduce from
Helmholtz's Equation (122.6) that w Vv = v Vw.
11. If an incompressible liquid in irrotational motion occupies a simply

f
connected region, show that

fv2dv
= dodS'
where p is the velocity potential and the normal derivative dp/dn is in the
direction of the external normal to the bounding surface. [Cf. (108.1).]
12. Prove Kelvin's theorem: The irrotational motion (v) of an incompres-

z
p

the same boundary conditions.


f
sible fluid occupying a simply connected region S with finite boundaries has
less kinetic energy (T = v v dV) than any other motion (vl) satisfying

[Put vi = v + v'; then div v' = 0 and n v' = 0, where n is the unit nor-
mal vector to the boundary S. Show that Ti = T + T'.]
282 HYDRODYNAMICS
13. Under the conditions of Problem 11, show that if
(a) v = const. over the boundary, or
(b) dip/dn = 0 over the boundary,
then V is constant throughout the region and v = 0.
Hence show that the irrotational motion of a liquid occupying a simply
connected region is uniquely determined when the value of either w or d'p/dn
is specified at each point of the boundary.
14. If the body forces have a potential Q, the integrals,

T f a pv2 dV, U= f pQ dV,

represent the kinetic and potential energy of the fluid within the region of
integration. Show that for an incompressible fluid

d(T-[-U)_-
15. Assuming that the earth is a sphere of incompressible fluid of constant
density p and without rotation, show that the pressure at a distance r from its
center is
p = 2gpa(1 - r2/a2),
where a is the radius of the earth.
[If -y is the constant of gravitation, the attraction on a unit mass at the
distance r is
47rtipr3/r2 = gr/a
where g is the attraction when r = a; hence the body force -grR/a has the
potential 'gr2/a.]
Compute the pressure at the center if p is taken equal to the mean density
of the earth, (pg = 5.525 X 62.4 lb./ft.3).
16. In example 1 of § 127, consider the motion when n = r/a(0 < a < r).
Since the lines 0 = 0, 0 = a are parts of the same stream-line = 0, we have
the steady irrotational motion of a liquid within two walls at an angle a. Find
the radial and transverse components of v at any point (r, 0).
17. Discuss the plane, irrotational motion when the complex potential
w = rp + i¢ (§ 127) is given by z = cosh w. Show that the equipotential lines
and stream-lines are the families of confocal ellipses and hyperbolas:
xz y2
1,
c2 cosh' v + c2 sinh2 ,p -
x2 2

C2 COS2 ' c2 sin2 - 1'


with foci at (fc, 0).
Show that the stream-lines 4, = nor, where n is any positive integer, corre-
spond to the part of the x-axis from x = fc to x = f co. If we regard this as
a wall; we have the case of a liquid streaming through a slit of breadth 2c in an
infinite plane.
CHAPTER VIII

GEOMETRY ON A SURFACE

130. Curvature of Surface Curves. Let C be any curve on the


surface r = r(u, v) with unit normal n (94.11). As the point P
traverses C with unit speed, the Darboux vector,
(1) 6 = TT+KB,
gives the angular velocity of the moving trihedral TNB.
Consider now the motion of the dextral trihedral Tnp (p = T x n)
associated with the curve (Fig. 130). Since the trihedrals Tnp and

T
T points upward
from paper

FIG. 130

TNB have T in common, the motion of Tnp relative to TNB is a


rotation about T. If p = angle (N, n), taken positive in the sense
of T, the angular velocity of Tnp relative to TRB is (d<p/ds)T.
Therefore the angular velocity of Tnp is

(2) w=S -{- KB.


ds \T + ds /
Now N X n = T sin p from the definition of p and n x (N x n)
- p sin gyp; hence
N=ncosV - psingyp, B =TXN=pcosp+nsincp.
Substituting this value of B in (2) gives
(3) (T -f dco/ds) T + K sin p n + K cos (p p.
283
284 GEOMETRY ON A SURFACE § 130

The three scalar coefficients in (3) are written t, y, k and are


named as follows:
(4) t = r -}- d<p/ds, the geodesic torsion,
(5) y = K sin (p, the geodesic curvature,
(6) k= K cos rp, the normal curvature.
With this notation,
(3)' w = tT+-yn+kp.
If we reverse the positive sense on the curve, we must replace
T, s, <p by - T, -S7 -P (since T determines the sense of sp), but K
and r are unaltered (§ 45). Therefore a change of positive sense
leaves t and k unaltered, but reverses the sign of the geodesic curva-
ture y.
Since P is moving along the curve with unit speed, ds/dt = 1,
and we take s = t. For any vector u fixed in Tnb we have du/ds
= w X u (56.4) ; hence
do
(7) -=wxn=
ds
-kT-tnxT.
Now n is uniquely defined by (94.11) at all regular points of the
surface and do/ds = T Grad n is the same for all surface curves
through a point having a common unit tangent T. From (7),

(8) k = --do
ds
T, t = --do
ds
nxT,
we may therefore state
THEOREM 1. At a given point of a surface, the normal curvature
and geodesic torsion are the same for all surface curves having a
common tangent there.
Since cos p = n N, we see from (6) that the curvature K =
k/cos rp is completely determined by T and N at the point con-
sidered, provided p 5-6 a/2. But, since T and N determine the
osculating plane at the point (§ 48), we have
THEOREM 2. All curves of a surface passing through a point and
having a common osculating plane, not tangent to the surface, have
the same curvature there, namely, the curvature of the plane curve cut
from the surface by the osculating plane.
§ 131 THL DYADIC vn 285

In view of this theorem, we now confine our attention to the


curvature of plane sections of the surface.
Consider now the normal section Cn of the surface cut by a plane
through n and T at P. For this curve Nn = fn, cos cp = =E-1, and
Kn = ±k, according as N,, and n have the same or opposite direc-
tions. If k 0, the center of curvature of C at P, namely,
Cn = r + N. 1K. = r + n/k,
is called the center of normal curvature for the direction T. But
any surface curve C tangent to T at P has c = r + N/K for its
center of curvature. Hence c - cn = N/K - n/k, and
1 cost
(C - Cn.) T = 0, (c - C,,) N=-- = 0,
K k

from (6), so that c - cn is perpendicular to the osculating plane


of C.
THEOREM 3 (Meusnier). The center of curvature of any surface
curve is the projection of the corresponding center of normal curvature
upon its osculating plane if the latter is not tangent to the surface.
Curves on a surface along which t, y, or k vanish are named as
follows:
t = 0: lines of curvature,
-y = 0: geodesic lines or geodesics,
k = 0: asymptotic lines.
When t = 0,. do/ds = -k T, from (7); hence
dr do
(9) - X - = 0, along a line of curvature.
ds ds
When k = 0, do/ds = -t n x T, from (7); hence
dr do
(10) -ds - ds= 0, along an asymptotic line.

The parametric line v = const is a line of curvature when ru x nu


= 0, and an asymptotic line when r,, n,, = 0.
131. The Dyadic Vn. Let e and e' = n X e be two perpendicular
unit tangent vectors to the surface r = r(u, v) at the point Pl.
286 GEOMETRY ON A SURFACE
If k, t and k', t' are the normal curvature and geodesic torsion
associated with these directions, we have, from (130.7),
do
_ -ke - tnxe = -ke - te',
ds
do
_ - k'e' - t' n X e' k'e' + t'e;
ds'

do do
(1) Grad n = e
ds
+ e'
-- = e(-ke - te') + e'(-k'e' + t'e).
ds'
This dyadic is symmetric, since its vector invariant, Rot n = 0
(96.8). We have, in fact,
(2) Rot n = - (t + t')n = 0, t' = - t;
(3) Vn = -k ee - k'e'e' - t(ee' + e'e).
For brevity, here and elsewhere in this chapter, we use V instead
of Grad since no misunderstanding is possible. We state the re-
sult (2) as
THEOREM 1 (Bonnet). The geodesic torsions associated with any
two perpendicular tangents at a point of a surface are equal in magni-
tude but opposite in sign.
From (1) we find that the first and second scalar invariants of
On are - (k + k') and kk' + tt'. We write
(4) J = k + k' = - Div n,

(5) K = kk' - t2;

J is called the mean curvature and K the total curvature of the sur-
face at P. Since On and its invariants are point functions over
the surface, the values of J and K are independent of the choice
of e; we therefore have
THEOREM 2. For any pair of perpendicular tangents at P, k + k'
and kk' + tt' have the same value.
From (3), we have
(6) k= t= -eVne';
§ 131 THE DYADIC C-n 287

hence k and t are not altered when e is replaced by -e. If we


assume that k remains finite at P and is not constant, there are
certain directions for which k attains its extreme values. To find
these we examine the variation of k as e revolves about P in the
tangent plane. If the angle 0 between e and a fixed line in the
tangent plane is taken positive in the sense determined by n, we
have (§ 44)
de de'
do
=nXe = e', -=nXe'=
do
-e;
and, from (6),
dk
(7) de

The extremes of k therefore occur when t = 0. If the direction


el gives an extreme value k1, tl = 0. Then the perpendicular
direction e2 = n x e1, for which t2 = 0 from (2), gives another ex-
treme value k2. But since k + k' is constant, if k = k1 is a maxi-
mum, k' = k2 is a minimum, and vice versa.
If we take e = e1, e' = e2 in (3), we have
(8) Vn = -k1e1e1 - k2e2e2,
and the symmetric dyadic Vn appears in the standard form (72.4).
Evidently el and e2 are the invariant directions of Vn with multi-
pliers -k1, -k2. Moreover, if k is not constant, ±e1 and fee
are the only invariant directions at P. We state these results in
THEOREM 3. If the normal curvature is finite and not constant at
a point of the surface, it attains its maximum and minimum values
for just two normal sections, at right angles to each other, and char-
acterized by the vanishing of the geodesic torsion.
The orthogonal directions e1, e2 are called the principal direc-
tions at P, and the corresponding normal curvatures k1, k2, the
principal curvatures. From (4) and (5),
(9) J = kl + k2, K = k1k2;
consequently k1 and k2 are roots of the quadratic,
(10) k2-Jk+K=0,
the characteristic equation (§ 71) for the symmetric planar dyadic
On.
288 GEOMETRY ON A SURFACE § 131

Consider finally the case when k is constant at P; from (7),


t = 0 for all directions, and (3) becomes
(11) On = -k(ee + e'e') = -k(I - nn).
Any direction in the tangent plane at P is an invariant direction
with multiplier - k, and P is called an umbilical point or simply
an umbilic.
If k is constant over the entire surface, we have
Vn = -k or, V (n + kr) = 0,
and n + kr is constant over the surface. If k = 0, n is a constant,
and the surface is a plane. If k 0,

kr=c,
n
-- Ir-c12 = 1/k2;

and the surface is a sphere (center c, radius 1/k). Therefore the


only surfaces whose points are all umbilical are the plane and sphere.
All curves on the plane or sphere are lines of curvature (t = 0).
Example 1. Formulas (6) give k and t for any direction e. If the angle
(ele) = 0 is reckoned positive in the sense of n (el x e = n sin 0),
e =elcos0+e2sin0, e' = -elsin0+e2cosO.
Taking Vn in the form (8), we have
(12) k = e (klelel + k2e2e2) e = ki eos2 0 + k2 sin2 0,
(13) t=e (k1elel + k2e2e2) e' _ (k2 - k1) sin 0 cos 0.
These equations are due, respectively, to Euler and Bonnet. Since 2t =
(k2 - ki) sin 20, it is clear that t attains its extreme values ±(k2 - kl)/2 for
the directions 0 = fir/4.
Example 2. At a surface point the directions for which k = 0 are called
asymptotic. From (12), the asymptotic directions are given by tan20 =
-ki/k2 and are real and distinct only when K = klk2 < 0. When kl = 0,
k2 0 0, both asymptotic directions coalesce with the principal direction el.
In an asymptotic direction, t2 = -K, from (5).
Along an asymptotic line, k = K cos p = 0; and, if K , 0, cos v = 0, and
,p = =1= 7r/2. Along a curved asymptotic line, the osculating plane remains
tangent to the surface; moreover -y = ±K, and t = T.
Referred to the principal direction el the asymptotic directions have the
slopes f and are perpendicular when and only when these slopes
are ±1; then -kl/k2 = 1, and J = kl + k2 = 0.
§ 132 FUNDAMENTAL FORMS 289

Example 3. If e is the unit tangent vector of a surface curve, the trihedral


ee'n has the angular velocity,
(14) w = to - ke' + yn (130.3)',

along the curve, and


de de'
= cu x e = kn + ye', = w x e' = to - ye.
ds ds
On differentiating equations (6), we now have

ds UdVn / e,
ds d-s

Since dVn/ds is the same along all surface curves having the common tangent
e at a point, the same is true of the expressions,

(15)
ds - 27t and + y(k - k'),
d-s

respectively discovered by Laguerre and Darboux.


132. Fundamental Forms. On the surface r = r(u, v), the gra-
dient of a tensor function f(u, v) is given by (95.5),
(1) Vf = afu + bf,
where a, b, n is the set reciprocal to ru, r, n:
(2) aru + br + nn = I.
From (1), we have, in particular,
(3) Vu=a, Vv=b;
(4) Vr = aru + br = I - nn.
The dyadic Vr is symmetric and may also be written rua + rub;
it transforms any vector f(u, v) into its projection ft on the tan-
gent plane at (u, v) :
(5)
Vr acts as an idemfactor on vectors tangent to the sur-
face at the point (u, v) and on dyadics whose vectors lie in the
tangent plane.
If we form the product,
(Vr) (Vr) = (aru + (rua + r,;b),
290 GEOMETRY ON A SURFACE § 132

we obtain
(6) yr = E as + Flab + ba) + G bb,
where
(7) E = ru ru, F = ru rv, G = rv r,
are the coefficients of the first fundamental form,
(8) dr dr = E due + 2F du dv + G dv2,
defined in § 94.In fact (8) follows from (6) when we form
dr yr dr and note that a . dr = du, b dr = dv, from (3).
If we compute
(vn) - (vr) = (anu + bn,) (rua +
we obtain, in similar fashion,
(9) - Vn = L as + M(ab + ba) + N bb,
where
L = - nu ru = n ruu,
(10) M = -nu r _ -nv ru = n ruv,

The relations between scalar products in (10) are obtained by


differentiating the equations,
=0,
with respect to u and v :
nu ru + n ruu = 0, nv ru + n ruv = 0,
nu rv+n rvv=0, n,,
From (9) we obtain the second fundamental form,
(11) -dr do = L du2 + 2M du dv + N dv2,
by computing the product dr Vn dr.
Remembering that Vn is symmetric, we next compute
(Vn) (Vn) = (anu + bnv) (nua + nab)
to obtain
(12) (Vn)2 = e as + flab + ba) + g bb,
§ 132 FUNDAMENTAL FORMS 291

where
(13) e = nu nu, f = nu n,,, 9=nn
From (12), we obtain the third fundamental form,
(14) do do = e due + 2f du dv + g dv2,
by computing the product dr - (Vn)2 dr.
The quantities (7), (10), and (13) are known, respectively, as
the fundamental quantities of the first, second, and third orders.
If e1, e2 are unit vectors in the principal directions at P,
Vr = e1e1 + e2e2,
Vn = -k1 e1e1 - k2 e2e2,
(on)2 = ki e1e1 -1- k2 e2e2,
and, on multiplying these equations in turn by K = k1k2, J. _
k1 + k2, 1, and adding, we get
(15) (on)2+JVn+KVr=0.
This is the Hamilton-Cayley Equation for the planar dyadic Vn;
for its first and second scalar invariants are -J, K, and Vr is the
idemfactor in the tangent plane. Substituting from (6), (9), and
(12) for the dyadics in (15), we obtain the following relations be-
tween the fundamental quantities:
(16) e-JL+KE=0, f -JM+KF=0,
g-JN+KG=0.
Example 1. From the reciprocity of a, b, n and ru, r,., n,
r,, x H
(17) , bnxru,
ru x r,, n; hence
(18)

We now may compute the invariants -J and K of Vn:


Vn = anu +bn,,, (On)2 = axbnuxn,,;

ruxrv n
K = (a x b) . (nu x ne,) = nu x nv a
ruxr,- n
292 GEOMETRY ON A SURFACE
Since the cross products in these equations are all parallel to n, these equa-
tions may be written also in the vector form:
(19) -J ru x r,. = nu x r,, + ru x n,.,
(20) K ru x rn = nu x nv.
If we multiply these equations by (ru x r,;) , apply (20.1), and introduce
the fundamental quantities from (7) and (10), we have
(21) J(EG - F2) = GL - 2F31 + EN,
(22) K(EG - F2) = LN - 312.
These values of J and K also may be found from
(9) -Vn = a(La + Mb) + b(Ma + Nb),
by making use of (18).
Example 2. The surface z = z(x, y), with the position vector,
r=xi+yl+z(x,y)k,
can be written in the parametric form
x=u, y=v, z=z(u,v).
If we denote the partial derivatives z.,, zy, zxx, zxy = zyx, zyy by p, q, r, s, t,
respectively, we have
ru=rx=i+pk, r,, =ry=l+qk;
hence, from (7),
E = 1 + p2, F = pq, G = 1 + q2;
H=EG-F2=1+ p2+g2;
Hn=ruxru= -pi - gl+k.
Furthermore,
ruu = rk, ru = sk, r,.y. = tk;
hence, from (10),
L=r/H, M=s/H, N=t/H.
We may now compute the mean curvature J from (21), the total curvature
K from (22):
(23)
T- (1+q2)r-2pgs+(1+p2)t
(1 + p2 + q2)
rt - s2
(24)
K = (1+p2+q2)2
The principal curvatures k1, k2 are the roots of the quadratic (131.10):
(25) k2 - Jk + K = 0.
§ 134 TI-HE' FIELD DYADIC 293

133. Field of Curves. A one-parameter family of curves on a


surface r = r(u, v) is said to form a field over a portion S of the
surface if one and only one curve of the family passes through
every point of S. After the positive direction on one of the curves
has been chosen, the positive direction on the others is taken so
that the unit tangent vector to the curves is a continuous vector
point function over S.
If the curves of the field have the equation p(u, v) = const,
their differential equation is
ccudu
From this we may compute Al = dv/du at all points of S. Then
the vector,
dr ar ar av
ru -I Alrv,
du a u + av au -
is tangent to the field curve through the point (u, v).
Similarly if A2 = dv/du for a second one-parameter family of
curves, the vector ru + A2r is a tangent to the curve through
(u, v). When the second family cuts the first everywhere at right
angles,
(ru + Air,.) (ru + A2r,.) = 0,
or, in terms of fundamental quantities,
(1) E + (A1 + A2)F + A1A2 G = 0.

Since Al = -,pu/p is known, (1) is the differential equation of the


orthogonal trajectories of the field curves. Standard existence theo-
rems for differential equations of the first order guarantee, under
very general conditions, a solution of (1).
Since ru.,r are tangent vectors to the curves v = const, u =
const, the parametric curves cut at right angles when
(2)
134. The Field Dyadic. Consider a field of curves C1 and their
orthogonal trajectories C2 on a portion S of the surface r = r(u, v).
At every point (u, v) their unit tangents e1, e2 = n x e1, and the
surface normal n form a dextral trihedral of unit vectors e1e2n.
As a point traverses CI with unit speed, e1e2n has the angular
velocity (130.3),
(1) wl = t1e1 - k1e2 + yin;
294 GEOMETRY ON A SURFACE § 134

since p1 = el x n = -e2. Similarly, along C2,


(2) (02 = t2e2 + k2e1 + y2n,

since p2 = e2 x n = e1. Therefore


de1
= w1 x e1 = kin + yle2,
ds1

de1
= w2 x e1 = -t2n + y2e2
ds2
We now have, from (95.8),
de1 de1
Grad e1 = e1
ds1 + e2 ds2

= e1(k1n + y1e2) + e2(t1n + 72e2)


= (k1e1 + t1e2)n + (y1e1 + y2e2)e2,
since t2 = -t1 by Bonnet's Theorem (§ 131). Now, from (130.7),
do
-k1e1 - t1 n x e1;
ds1
and, if we write
(3) R = y1e1 + 72e2,
the field dyadic Grad e1 for the curves Cl becomes
do
(4) vet = - n + R e2-
ds1

If we replace e1 and e2 in (3) by n x e1 = e2, n x e2 = -e1,


R is unchanged, since y2e2 + (-y1)(-e1) = R. Therefore the
field dyadic for the curves C2 is
do
(5) vet =- ds2
-n - Re,.
We have moreover
do do do do
(6) Vn=e1-+e2-=-e1+-e2.
ds1 ds2 ds1 ds2
With the notation,
do do
(7) P = ds2
ds1
§ 1.34 THE FIELD DYADIC 295

the preceding equations become


(4)' Vet = Re2 - Pn,
(5)' De2 = - Re 1 - Qn,
(6)' Vn = Pel + Qe2.
From (4), we have

or, in view of (3),


y1 = n rot el.
The geodesic curvature is therefore a surface invariant.
THEOREM. A field of curves having the unit tangent vector e(u, v)
has at every point the geodesic curvature,
1
(8) y = n rot e = { (r e),, - (ru
H
The last expression follows from (97.9).
Consider now a, curve C which cuts the curves Cl at an angle
6 = (eli e), reckoned positive in the sense of n. Along C the tri-
hedral ee'n has the angular velocity
(9) w = to - ke' + yn (131.14);

and, since the trihedral e1e2n has the angular velocity - (de/ds) n
relative to ee'n, the angular velocity of e1e2n along C is
dB dB\
(10) w -asn=to - ke'-F y
-ds n'
/ w -n- Xe1,
hence
de1 \
ds ds

dO \ dO

The left member of (11) shows that y - dB/ds is the same for all
surface having e as common tangent vector.
296 GEOMETRY ON A SURFACE § 134

Since Vel e2 = R, from (4), we may write (11) in the form:


de
(12) e R =
ds

If C is a member of a field of curves, the angle 0 = (el, e) is a


point function over this field, and
(13) R= =ye+y'e'-OB.
If the field curves C, and C cut everywhere at the same angle,
VO=0,and R=ye+y'e'.
Example. For a field of surface curves p(u, v) = c, the unit tangent vec-
tor is
e =dr/ds where puu.+cvv =0.
Therefore
y
-U _ - 'Pu _ ` ;
'Pv
1

X
u _ cv- , v
(Pa

and, if we substitute these values in


e e =Eci2+2Fuv+Gi,2 = 1,
we have
= E Vn - 2F vupt, + G wo
to find X. Thus
e=
where the sign is chosen to give the positive sense desired. The geodesic
curvature now is given by (8) :
(FP,, a (EPv - Fvu)1
(14)
H lau Gpu) _ av

This formula is due to Bonnet.


Let us apply it to find the geodesic curvature of the parametric curves in
the sense of increasing u and v. For the curves v = const, vu = 0, pv = 1,
a = 1/E, el = ru/N/-r,,; for the curves u = const, vu = 1, 'RU = 0, a
e3 = r/1'G; hence in the respective cases,

(15) Y1 =
H S
av E} ya = -G 1 }
l iau G av
\1

These formulas also follow at once from (8) when the foregoing values of el
and e3 are used.
The subscript 3 refers to the curves u = const when they cut the curves
v = const at an arbitrary angle; when this angle is it/2, we use the subscript 2.
§ 135 GEODESICS 297

135. Geodesics. The geodesic curvature of a curve C has been


defined (§ 130) as
(1) K sin gyp,

where p, the angle (N, n), is taken positive in the sense of T. Since
dT/ds = KN,
dT
(2) -xn = KNXn = KSin VT = yT;
ds
and, as dT/ds is unaltered by a change in positive direction, y must
change sign with T (§ 130).
A surface curve for which y = 0 is called a geodesic. Thus, if a
straight line can be drawn on a surface, it is necessarily a geodesic
since K = 0. If a geodesic is curved (K 0), y = 0 implies sin <p
= 0 and <p = 0 or 7r. We thus have
THEOREM 1. Along a curved geodesic the principal normal is
always normal to the surface.
Along a curved geodesic,
dip
(3) k = K COS p = ±K, t = T + - = T;
ds

and, if the geodesic is straight (K = r = 0) and the same equations


apply (k = t = 0).
THEOREM 2. Along a geodesic the normal curvature is numeri-
cally the same as the curvature and the geodesic torsion equals the
torsion.
On putting y = 0 in (2), we have
d2r d2r
(4) - x n = 0, or or = 0,
ds2 ds2
since the tangential projection of d2r/ds2 is zero (132.5). If we
replace Vr by aru + br,,, we obtain the scalar differential equa-
tions of a geodesic:
(5) ru
-
d2 r
d82
= 0,
d2r
r - = 0.
dS2

These differential equations of the second order show that the geo-
desics on a surface form a two-parameter family. In general, a
geodesic may be determined by two conditions, by specifying (a)
298 GEOMETRY ON A SURFACE a 1:35

that it shall pass through a given point in a given direction, or (b),


that it shall pass through two given points.
That a geodesic on a surface in general can be found to fullill
conditions of the type (a) or (b) is plausible in view of the follow.
ing theorems from mechanics.
THEOREM 3. A particle constrained to move on a surface and fief
from the action of any tangential forces will describe a geodesic writi
constant speed.
Proof. In view of (52.5), the equation of motion ma = F may
be written
(ddt
m T + Kv2 N) = pn,
where v is the speed and pn the normal force. Multiplying by T
gives dv/dt = 0, v = const. If K = 0, the particle describes a
straight line. If K F4- 0, N = ±n, and sin cp = 0. Since either
K or sin (p must vanish, y = 0 in both cases, and the particle de-
scribes a geodesic. The pressure I p I = mKV2.
THEOREM 4. If a weightless flexible cord is stretched over a smooth
surface between two of its points, its tension is constant, and the line
of contact is a geodesic.
Proof. Since the cord is in equilibrium, the vector sum of all
the forces acting upon any portion of it vanishes. Let F denote

FT

Fia. 135

the magnitude of the tension at P, distant s (arc POP) from the


fixed end of the cord, and pn the normal reaction of the surface
per unit length. Then from the equilibrium of the length s of the
cord (Fig. 135),
FT - FOTO + o s pn ds = 0.
0
§ 136 GEODESIC FIELD 299

On differentiating this equation with respect to s, we have


dF
- T+FKN+pn = 0;
ds
hence dF/ds = 0, and F is constant. If K = 0, the line of contact
is straight; if K n= ±n, and I p I = KF. In either case,
-y = 0, and the line of contact is a geodesic.
As a cord can be stretched between any two points of a convex
surface, the line of contact is a geodesic fulfilling condition (b).
On concave surfaces we must imagine the cord replaced by a thin
strip of spring steel laid flatwise.
136. Geodesic Field. Consider now a one-parameter family of
geodesics that form a field of tangent vector el over a portion S
of the surface. Then, since
(1) y1 0,

we have, from Stokes' Theorem,

fT e1 ds = fn rote,dS=0,
for any sectionally smooth closed curve lying entirely within S.
Writing 0 = angle (e1, T) gives

(2) fcos 0 ds = 0
as the integral equivalent of (1). This formula leads to
THEOREM 1. An arc of a geodesic that is one of the curves of a
geodesic field is shorter than any other surface curve joining its end
points and lying entirely within, the field.
Proof. Let AI'B be a geodesic arc of the field and AQB any
other curve of surface covered by the field (Fig. 136a). Applying
(2) to the circuit APBQB formed by these arcs, we have

f ds +
PB BQA

f PB(ls = f QBcos 0 ds < /QBI cos 0 1 ds < fQBds;

that is, are APB < are AQB.


300 GEOMETRY ON A SURFACE § 137

We next consider the orthogonal trajectories of a geodesic field.


Their basic property is given by
THEOREM 2 (Gauss). The orthogonal trajectories of a geodesic
field intercept equal arcs on the geodesics.

Q
_A'
B!

B-
-A

FIG. 136a FIG. 136b

Proof. Let AB, A'B' be geodesic arcs intercepted between the


curves AA', BY cutting them at right angles (Fig. 136b). Then
applying (2) to the circuit ABB'A'A, and, noting that
cos0 = 1,0, -1,0
over AB, BY, B'A', A'A, we have are AB = arc A'B'.
137. Equations of Codazzi and Gauss. These celebrated equa-
tions in the theory of surfaces simply state that
(1) n - Qx(Vn) = 0,
(2) n - V x (Ve1) = 0.
If we remember that V means V8, both identities follow from
(97.11).
In order to express these identities in terms of the vectors,
(3) P = do/dsl, Q = dn/ds2, R = y1e1 + '2e2,
we make use of (97.12) :
(4) n - Vx(pq) _
On substituting
(5) Vn = Pe1 + Qe2
in (1), we have, from (4),
(n - rotP)e1 + 0;
and, if we put (§ 134)
(6) Vet = Re2 - Pn, Ve2 = -Re1 - Qn,
§ 137 EQUATIONS OF CODAZZI AND GAUSS 301

we have
(n rot P - Q n x R)e, + (n rot Q + P n x R)e2 = 0.
This equivalent to the two equations:
(7) n- rot P = -n - QxR,
(8) n rot Q = -n R x P.
These are the Equations of Codazzi.
We next put Ve, = Re2 - Pn in (2) and obtain
(n - rot R)e2 - (n -
Replacing vet and Vn by the foregoing values, we have

This is also equivalent to two equations. One of these is the same


as (7) ; the other,
(9)
is the Equation of Gauss. In the right-hand member,
K,
the second scalar invariant of Vn or the total curvature (§ 131).
Gauss's Equation thus becomes
(10) n rot R = -K.
This is perhaps the most important result in the theory of surfaces.
The equations of Codazzi and Gauss can be expressed in terms
of the quantities ki, tt;_,yi along the field curves.t From (130.7),
do do
-
i
= -kie, - tin x e,, as = -k2e2 - t2n x e2,
2

and, since t.2 _ n,, -ei,


(11) -P = kie, + tie2i -Q = tie, + k2e2 (134.7).
In order to compute the left members of (7), (8), (10), we apply
the identity,
(12)
t Now k1 and k2 denote the normal curvatures of the orthogonal field curves;
only when t1 = t2 = 0 are k1 and k2 principal curvatures (§ 131).
302 GEOMETRY ON A SURFACE § 13S

which follows at once from (85.3), to both terms of -P, -Q and


R. Remembering that n rot ei = yt, we thus obtain
dk1 dt1
-n rot P = y1k1 + y2t1ds2- - +dsl-
dt1 dk2
-n rot Q = 'Y1t1 + -12k2 - +
ds2 dsl

yi+y2 - dyl dye


ds + ds
2 1
and, since
n Q x R = ylk2 - y2tl, n R x P = -yltl + y2k1,
the Equations of Codazzi and Gauss become
dt1 dk1
(13) y1(k1 - k2) + 2724 + -ds1- -ds2= 0,
dk2 dt1
(14)
dsl
- ds = 0,
72(k2 - k1) + 2y1t1 +
2

(15) yi+yi+---+h'=0.
2

ds2
dye
ds1
d-11

Equation (14) states the same property for the e2-field that (13)
states for the el-field. To show this, change the subscripts 1, 2
in (13) into 2, -1 (n x e2 = -e1) ; then, since
k-1 = k1i t2 = -t1, y-1 = -71, ds_1 = -ds1,
we obtain (14).
138. Lines of Curvature. The curves on the surface which are
everywhere tangent to the principal directions (§ 131) are called
lines of curvature. Along a line of curvature the geodesic torsion
is zero (t = 0), and the torsion z = - dcp/ds.
From (130.7), we have do/ds = -kT along lines of curvature.
Their differential equation is therefore
dr do dr do
= 0, or = 0;
(1) dsxds n'dsxds
for (dr/ds) x (dn/ds) is always parallel to n. In particular, the
parametric curves are lines of curvature, if
ruxnu = 0, 0;
§ 138 LINES OF CURVATURE 303

then, if k1, k2 are the principal curvatures,


(2) nu = -klru, n = -k2r,.
On the plane and sphere, all curves are lines of curvature (§ 131).
On other surfaces there are in general two orthogonal principal
directions at each point; and the lines of curvature form two fields
cutting each other at right angles. If el and e2 = n X el are the
corresponding unit field vectors, the Equations of Codazzi become
dk1 dk2
//
(3) = 'Y1 \ki - k2), = y2(kl - k2).
dsl

These lead to a simple proof of


THEOREM 1. If K = 0, J 0 0, the lines of curvature along which
the normal curvature vanishes are straight lines.
Proof. Since K = klk2 = 0, J = k1 + k2 0 0, we may suppose
that k1 = 0, k2 0 0. Hence yl = 0, from (2); and, since
kl = K1 COS cp = 0, yj = Kj sin c = 0,
we conclude that K1 = 0.
Along the rulings of the surface (which are asymptotic lines as
well as lines of curvature) t1 = 0, kl = 0, and do/dsl = 0 (130.7);
hence n has a fixed direction along a ruling. This also follows
from (134.1); for co = 0, and the trihedral ele2n has a fixed orien-
tation along a ruling. In general a surface has a different tangent
plane at each point and is therefore the envelope of a two-param-
eter family of planes. However the ruled surface under considera-
tion has the same tangent plane along an entire ruling and is
therefore the envelope of a one-parameter family of planes. Such a
surface is called developable.
Consider now the lines of curvature cutting the rulings orthog-
onally. Their tangent vector e2 = n x el is constant along a ruling.
From (137.14) and (137.15),
dk2
122)
dye
= -y2i2. _d
_
'Y2
= 0.
ds1 dsl dsl k2

But y2/k2 = tan P2; hence X02 = (N2, n) is constant along a ruling,
and N2 as well. This property lends plausibility to the fact that
the developable surface may be bent into a plane (after certain
cuts are made) without stretching or tearing.
304 GEOMETRY ON A SURFACE §138

We now may amplify theorem 1: If K = 0, J 54 0, the surface


is developable. We shall see in § 142 that developable surfaces are
of three types; cylinders (including planes), cones, and tangent sur-
faces (generated by the tangents to a curve which is not a straight
line).
THEOREM 2. The normals to a surface along one of its curves have
an envelope when and only when the curve is a line of curvature; in
this case the envelope is the locus of the centers of normal curvature
along the curve.
Proof. Let s denote the are measured along the curve C of the
surface S, and r(s) and n(s) the position vector and unit surface
normal at a point of C. The points on the family of normals have
position vectors r(s) + X n(s), where X is a variable scalar. If we
take X as a function of s, say X = X(s), the curve C,
f = r(s) + X(s)n(s)
will be the envelope of the normals if
df do dX do
- =T+X
+
dsn or
T+X --
is parallel to n. But since both T and do/ds are perpendicular to
n, the envelope will exist only when
do
T+X = (1 - Xk)T - Xt n x T = 0 (130.7),

that is, when t = 0 and X = 1/k. Therefore the normals have an


envelope only along lines of curvature; and then the envelope is
the curve f = r + n/k, the locus of the centers of normal curva-
ture along C.
We turn now to some theorems relative to the lines of intersec-
tion of surfaces.
THEOREM 3. If two surfaces S1, S2 cut under a constant angle 0,
the curve of intersection has the same geodesic torsion whether regarded
as a curve of Sl or of S2.
Proof. If N is the unit principal normal along the curve of
intersection,
P2 - Pl = (N, n2) - (N, nl) = (n1, N) + (N, n2) = (n1, n2) = 0,
§ 138 LINES OF CURVATURE 305

and

t2-tl =T -}---- T -(lg'ds-_-do


d.s
=O.
ds

From this result, we have at once


THEOREM 4 (Joachimsthal). If two surfaces cut under a constant
angle, their curve of intersection is a line of curvature of both or of
neither; and, conversely, if the line of intersection is a line of curva-
ture of both, the surfaces cut under a constant angle.
We are now in position to prove the celebrated
THEOREM 5 (Dupin). Two surfaces belonging to different fami-
lies of a triple orthogonal system cut one another in lines of curvature
of each.
Proof. Denote the geodesic torsion at a point P on the curve
of intersection of the surfaces Si and Sj by tij or tji, according as
the curve is regarded as belonging to Si or Sj. Now at the point
P we have
tij = tji (theorem 3), tij = -tik (§ 131, theorem 1),
where i, j, k represents any permutation of the indices 1, 2, 3.
Then
tij = - tik = - tki = tkj = tik = - tji = - tij,

so that tij = 0.
Example. A surface of revolution about the z-axis has the parametric
equations,
z = u cos v, y = u sin v, z = z(u),
where u, v are the plane polar coordinates (p, gyp). These give the vector
equation,
r = u R(v) + k z(u),
where R is a unit radial vector (R k = 0). Now
ru=R+kz', r =uP;
ruXr k - z'R
n = ru x r _ (t + z,2)

z" z'
nu = - (1 + z,2)2
ru, n + , r,,.
u(1 z'2)=
306 GEOMETRY ON A SURFACE § l39

The last equations show that the parametric lines (the parallels and meridians
of our surface) are the lines of curvature, and that the principal curvaturv
are
Z/1
z'
kl =
(1 + z) 1'
k2 =
u(1 + z,2)z
Consequently,
z' z"
(4) K = k l k2 = u(l + z,2)2

(5) J=k1+k2= uz" + z'(l + z'2)


u(l + z'2)2

Let us apply these results to find K and J for the torus generated by revolv-
ing the circle,
(u-a)2+z2=b2,
about the z-axis. On differentiation, we find
a-u 1 +z'2
b2 b2
zt =
z z2
z" _ -
2
3
hence, from (4) and (5),
_u - a a-2u
J
K b2u bu

If we introduce the latitude 0 on the generating circle, u = a + b cos 0, and

(6) K_
b(a+bcose)'
Cos0 -J ba+bcos0
+
cos0
1

139. Total Curvature. The Gauss Equation (137.15),

(1) 1Y1 f- yz
+ dyl
- + K = 0,
ds1 ds2

shows that, if orthogonal geodesics can be chosen as fielt ^urves,


71 = 72 = 0, and K = 0; hence
THEOREM 1. Orthogonal geodesic fields can exist only on
of zero total curvature.
We now compute K from the Gauss Equation (137.10); using
(97.9), we have
1
(2) - K = n rot R =
- { (r R)u - (ru
H
§ 139 TOTAL CURVATURE 307

Let the parametric curves cut at an angle 0 = (el, e3) where el =


ru/-/E, e3 = r, Then, if e2 = n x el, e4 = n x e3i we have,
from (134.13),
R = ylel + y2e2 = y3e3 + y4e4 - V0;
and

yj
/
-Vi, -/ ae
y3VG-- (95.7).
av

Substitution in (2) now gives the elegant Formula of Liouville for


thi' total curvature,

(3) K
1 ja
- H 1av (yl
1/E) -- a

au
(-I"-\/G' ) +
020

au av.
in which II = 1/EG - F2, yl and 73 are given by (134.15), and
F
(4)
-\/EG EG
When the parametric curves are orthogonal, 0 = it/2 and F = 0,
II = EG. Equations (134.15) now give
1 aE/av 1 aG/au
(5) 73
yl = - 21i NIT ' - 211 \/G
and Liouville's Formula becomes
1 I aG a / 1 aE)l
aau (il_
(6) h'
211 au + av \II a v f
From (3), we obtain an immediate proof of
THEOREM 2 (Gauss). The total curvature K of a surface depends
only upon ' e coefficients E, F, G of the first fundamental form and
their fir f .,and second partial derivatives with respect to u and v.
This iu.idamental theorem was first proven by Gauss after long
and t' iious calculations. Gauss's name of "Theorema egregium"
(La' n egregius means literally "out of the herd") for this result
shows that he was fully aware of its importance.
Example. To find a surface of revolution of constant negative total curva-
ture, we set K = -1/a2 in (138.4). The resulting differential equation,
2z'z" 2u
(1 + z'2)2
a2 '
308 GEOMETRY ON A SURFACE § 140

has the first integral,


u2
1

1 + z'2 a2
A or cost i = u2
a2 + A,

where >G = tan -l z' is the inclineftion of the tangent to the u-axis (Fig. 139).
If we impose the condition u = -a when 4, = 0, A = 0 and u = -a
Now
dz dz du
= = tank a sin 0 = a(sec ¢ - cos 0,
d,k du dik

z = a log (sec V, + tan p) - a sin ¢ + B;

FIG. 139

and B = 0 if z = 0 when ¢ = 0. Therefore the meridian of our surface of


constant negative K has the parametric equations:
u = -a cos ¢, z = a log (sec V, + tan ik) - a sin ik.
This curve, asymptotic to the z-axis as 4, --> a/2, is called a tractrix. The
equation u = -a cos yG shows that the segment of any tangent between the
curve and z-axis has the constant length a. This property is characteristic of
the tractrix (ef. § 50, ex. 5).

140. Bonnet's Integral Formula. The differential formula for


the total curvature,
(1) K = -n rot R,
has an important integral equivalent. If we integrate K over a
portion of a surface S bounded by a simple closed curve which
consists of a finite number of smooth arcs, we have, by Stokes'
§ 140 BONNET'S INTEGRAL FORMULA 309

Theorem,

fKdS= -
f(. - ds,

from (134.12); hence

(2) fKds = f fdo - y ds.

If the bounding curve has a continuously turning tangent, fdo


= 2ir and (2) becomes

(3) fKdS = 2ir fds.

eter, Bonnet, in 1848. The integral


integral curvature of S.
f
This very important formula was discovered by the French geom-
K dS over S is called the

Let us first apply (2) to the figure bounded by two geodesic arcs
APB and AQB meeting at A and B. Then denoting the interior
angles at the corners by A and B (Fig. 140a), we have

(4) fK dS = fdo = 2ir - (ir - A) - (ir - B) =A+ B.


Since A + B > 0 this equation is impossible when K = 0; hence,
on a surface of negative or zero total curvature, two geodesic arcs

Fu}. 140a FIG. 140b


310 GEOMETRY ON A SURFACE 3 14t)

cannot meet in two points so as to enclose a simply connected area.


We may state (4) as follows:
THEOREM 1. The integral curvature of a geodesic lune is equal to
the sum of its interior angles.
We next apply (2) to a geodesic triangle ABC (Fig. 140b) : that
is, the figure enclosed by three geodesic arcs. Again denoting the
interior angles at the corners by A, B, C, we have

fKdS = fdO=27r- (7r-A) - (ir-B) - (rr-C),


(5) fKds=A+B+C_.
THEOREM 2 (Gauss). The integral curvature of a geodesic triangle
is equal to its "angular excess," that is, the excess of the sum of its
angles over 7r.

When K is constant, the integral curvature f K dS is the prod-


uct of K by the area S. If K is identically zero, as on a plane, the
sum of the angles of a geodesic triangle is 7r. If K 34 0, we have
THEOREM 3 (Gauss). On a surface of constant non-zero total curva-
ture, the area of a geodesic triangle is equal to the quotient of its
angular excess by the total curvature.
Example 1. A sphere of radius r has the constant total curvature K = I /r2.
For all normal sections at a point are great circles of curvature 1/r; all points
are umbilical points and k1k2 = 1/r2. Formula (5) now states that: The area
of a spherical triangle is r2 times its angular excess in radians.
If e denotes the angular excess in right angles, we therefore have the formula,
4,rr2
Area = r2 2 e = e,
8

that is, the area of a spherical triangle equals the area of a spherical octant times
the angular excess in right angles.

Example 2. Let us decompose a closed bilateral surface S, consisting en-


tirely of regular points (§ 93), into a number f of curvilinear "polygons" Si
whose sides are analytic arcs. These sides meet in vertices at which at least
three arcs come together. Since S is regular and bilateral, the surface has a
continuous unit external normal n which defines a positive sense of circuit on
each polygon by the rule of the right-handed screw. When positive circuits
§ 140 BONNET'S INTEGRAL FORMULA 311

are made about two adjoining polygons, their common side is traversed in oppo-
site directions. Now for each Si we have, by Bonnet's Theorem (2),

J(i KdS=27r- (7r-ai;)-Js


where ail are the interior angles at the vertices of Si. If these equations for
all of the f polygons Si are added, we have
i j

fKdS 2irf - (a -

for the integrals f y ds over adjoining sides cancel in pairs since y changes
sign with change of direction. In each polygon the number of interior angles
equals the number of sides; hence 2;2:7r = 2eir, where e is the total number of
sides or edges. Moreover the sum of the interior angles about any vertex is
27r; hence TdIai; = 2v7r, where v is the total number of vertices. We thus
obtain
(6) f-e+v=1 KdS.
is,
From the right member of (6) we see that the number on the left is independent
of the manner in which S is subdivided into polygons. From the left member
we conclude that cb K dS is not altered when S is deformed into another
completely regular surface.
If S can be continuously deformed into a sphere, we have

(7) fK dS = 47rr2 = 4a

for a sphere, and


(8) f - e + v = 2.
Evidently f - e + v is not altered by any continuous deformation of S, even
though the resulting surface is not completely regular. Thus, if S is trans-
formed into a polyhedron (a closed solid with plane polygons for faces), the
relation (8) still holds good and constitutes the famous Polyhedron Formula
that Euler discovered in 1752: Faces + vertices - edges = 2.
Next suppose that S can be continuously deformed into a torus. From
(138.6), for a torus,

(9) f"KdS=ffKb(a+bcoso)dedv =
0
2
f
0
2"
cos e do dv, = 0;

hence f - e + v = 0 for any surface continuously deformable into a torus.


This relation holds, for example, for any ring surface with polyhedral faces.
312 GEOMETRY ON A SURFACE §140

Example 3. Parallel Displacement. A vector f, that always remains tangent


to a surface S, is said to undergo a parallel displacement along a surface curve
C when the component of df/ds tangential to the surface is zero; then
df df
(10) nx--=0 or ds=>`n.

Since

the length of f must remain constant during a parallel displacement.


Now the trihedral Tnp associated with the curve C has the angular velocity,
w = tT + yn + kp (130.3').

If the angle 0 = (T, f) is reckoned positive in the sense of n, the trihedral


fng (g = f x n) has the angular velocity n dB/ds relative to Tnp. Hence, as
f moves along C with unit speed, the trihedral fng revolves with the angular
velocity,
SZ = to + n - = tT + (y + dB/ds)n + kp,
hence df/ds = St x f, and

dfxn = (12 xf) xn = (n - fl)f = +do> f.


C'Y

Hence f will undergo a parallel displacement along the curve C when and only
when
de
(11) y+ - =0.
If f is tangent to C (or, more generally, when 0 remains constant), f will
undergo a parallel displacement along C only when C is a geodesic.
THEOREM. If a tangential surface vector is given a parallel displacement about
a smooth closed curve C, it will revolve through an angle,

(12) 27r - if y ds = f K dS.

Proof. From (11) we see that f revolves through the angle,

-
if ds fy ds

relative to T, the unit tangent vector of C. Since T itself revolves through 2r


in passing around C, the total rotation of f is 27r - fy ds, orJK dS by
Bonnet's Integral Formula (3). JJJJJ
§ 141 NORMAL SYSTEMS 313

141. Normal Systems. At every point of the surface S, r =


r(u, v), a straight line is defined by the unit vector m(u, v). The
points of this two-parameter family of lines are given by
(1) rl = r + Am.
Under what circumstances do these lines admit an orthogonal sur-
face? That is, when do they form the system of normals to a
surface?
When A is a definite continuous function A(u, v), r1 is the posi-
tion vector of a surface S1. If V denotes a surface gradient rela-
tive to S, we have, from (1),
Vr1 = Or + (VX)m + AOm,
+VA,
since m m = 1. In Vr1 = arlu + brl,,, the postfactors are tan-
gent to S1; hence, in order that m be normal to S1, it is neces-
sary and sufficient that (Or1) m = 0, that is,
(2) - VA = (Or) m = m - (m n)n (132.5).

But (2) implies that


(3) n rot m = 0;
conversely, from § 103, theorem 2, (3) implies that

(4) mg = (Or) m = -VA where X= - f m dr,


ro

the integral being taken over any path from ro to r(u, v) on S. We


have thus proved
THEOREM 1. In order that a two-parameter family of lines
r(u, v) + Xm(u, v) form a normal system, it is necessary and suffi-
cient that n rot m = 0, where n is the unit normal to the surface
r = r(u, v).
When m fulfils condition (3), we can find a one-parameter family
of surfaces normal to the lines r + Am. We need only determine
A(u, v) from (4) with some definite choice of ro. Since ro may be
chosen at pleasure, all possible values of A are then given by
314 GEOMETRY ON A SURFACE § 142

X + C, where C is an arbitrary constant. We thus obtain a one-


parameter family of surfaces,
r1 = r + (X + C)m,
normal to the lines.
Consider now a geodesic field over S with the unit tangent vec-
tor e. Since -y = n rot e = 0 at every point of S, we have
THEOREM 2 (Bertrand). The tangents to the geodesics of a field
form a normal system of lines.
We conclude with an important theorem in geometrical optics.
THEOREM 3 (Malus and Dupin). If a normal system of lines is
reflected or refracted at any surface, it still remains a normal system.

FIG. 141

Proof. Let el denote the unit vectors along the incident rays
and e2 the unit vectors along the refracted (or reflected) ray (Fig.
141). If µi denotes an absolute refractive index, µ1 sin 01 =
µ2 sin 02 (Snell's Law) and e1, e2, n are coplanar; hence,
(µ2e2 - µ1e1) x n = 0 and n-rot (µ2e2 - µ1e1) = 0,
by the theorem following (97.10). For the reflected ray (§ 75),
e2 = (I - 2nn) e1 and n - rot (e2 - e1) = 0.
In either case: n rot e1 = 0 implies n - rot e2 = 0-
142. Developable Surfaces. A developable surface (§ 138) is the
envelope of one-parameter family of planes. The limiting line of
intersection of two neighboring tangent planes is a line, or ruling,
on the surface. A developable is therefore a ruled surface. The
equation of a ruled surface may be written
(1) r = p(u) + ve(u),
§ 142 DEVELOPABLE SURFACES 315

where p = p(u) is a curve C crossing the rulings and e(u) is a


unit vector along the rulings. The surface normal is parallel to
ru x rv = (Pu + veu) x e.
When the normal maintains the same direction along a ruling, the
surface is developable. For the same plane is tangent to the sur-
face along an entire ruling; and, as each value of u corresponds to
a ruling and hence to a tangent plane, the surface is enveloped by
a one-parameter family of planes. Consequently the surface (1)
is developable when and only when pu x e and eu x e are parallel;
that is, when pu, e and eu are coplanar:
(2) pu e x eu = 0.
Case 1: e x eu = 0. Since e e = 1, we have e . eu = 0 and
hence eu = 0. Therefore e is constant, and (1) represents a gen-
eral cylinder.
Case 2: e x eu 34 0. We then may write (§ 5)

(3) Pu = a(u)e + 5(u)eu.


With a function A(u), as yet undetermined, let us write (1) as
(4) r= (P+Xe)+(v-X)e=q+(v-X)e,
where
(5) q(u) = p(u) + X(u)e(u).

Then
qu = pu + Xue + Xeu = (a + Xu)e + (S + x)eu;
and, if we choose X = -,6,
(6) qu = (a - {3u)e.
If a = l3u along C, qu = 0 and q is constant. Then (4) repre-
sents a cone of vertex q if the vectors e(u) are not coplanar, a plane
if they are coplanar.
In general, however, a 0 $u; then (6) shows that q(u) traces a
curve having e as tangent vector. The surface (4) then is gener-
ated by the tangents of the curve q = q(u); it is the tangent surface
of this curve.
Developable surfaces are planes, cylinders, cones or tangent surfaces.
316 GEOMETRY ON A SURFACE §143

143. Minimal Surfaces. In § 131 the mean curvature of a sur-


face was defined as J = - Div n = kl + k2.
With f = 1, the integral theorem (101.2) becomes

(1) fJnds=Jmde.
Next put f = r in (101.4) ; then, since Or = I - nn, Rot r = 0,
and we have
(2) fr x Jn dS =Jrxmds.
These equations admit of a simple mechanical interpretation:
THEOREM 1. The normal pressures Jn (per unit of area) over any
simply connected portion of surface S bounded by a closed curve C
are statically equivalent to a system of unit forces (per unit of length)
along the external normals to C and tangent to S.
Consider, now, a soap film with a constant surface tension q per
unit of length and subjected to an unbalanced normal force pn
per unit of area. Since any portion of the film bounded by a
closed curve C is in equilibrium, we must have

fpn dS + 2q Jmds= f (p + 2Jq)n dS = 0;


the surface tension q is doubled, because it is exerted on both sides
of the film. We therefore conclude that
(3) p = -2Jq.
Thus, for a soap bubble of radius r, the normal curvature is every-
where
k = «x - n = rc cos Tr =- 11r and J = -2/r;
thus the pressure inside of a soap bubble exceeds the pressure out-
side by an amount p = 4q/r.
In particular, if the film is exposed to the same pressure on both
sides, p = 0; then J = 0 at all points of the film. A surface whose
mean curvature J is everywhere zero is called a minimal surface.
Thus a soap film spanned over a wire loop of any shape and with
atmospheric pressure on both sides materializes a minimal surface.
Let So be a minimal surface bounded by a closed curve C. Its
existence is guaranteed by its physical counterpart, the soap film
§ 143 MINIMAL SURFACES 317

spanned over C. Imagine now that So is embedded in a field of


minimal surfaces, that is, a one-parameter family of surfaces in a
certain region R enclosing So, such that through every point of R
there passes one and only one surface of the family. Such a field
may be generated, for example, by the parallel translation of So.
At each point of R, the unit normal no to the minimal surfaces
may be so chosen that no is a continuous vector-point function;
and, from (97.2),
dno
div no = Div no + no = -J + 0 = 0.
do
Now let S be any other surface in R spanning the curve C and
enclosing with So a volume V of unit external normal n. From
the divergence theorem,

n=
f n no over

n no cos dS < fdS.


We have thus proved
THEOREM 2. A minimal surface spanning a closed curve and be-
longing to a field of minimal surfaces has a smaller area than any
other surface spanning the same curve and lying entirely in the field.
Example 1. In order to find a surface of revolution which is also minimal,
we set J = 0 in (138.5). The resulting differential equation,
uz" + z'(1 + z'2) = 0
may be written
z'z" z'z" 1
z72
+z'2+u=0

and has the first integral uz'/1/1 + z'2 = C. Solving for z', we have z' _
C/1/u2 - C2; hence
=Ccoshz Ck.
z + k =Ccosh-r C , u

This represents a catenary in the uz-plane; when revolved about the z-axis
it generates a surface known as the catenoid. Catenoids are the only minimal
surfaces of revolution.
318 GEOMETRY ON A SURFACE § 143

Example 2. The right helicoid is a surface generated by a line which always


cuts a fixed axis at right angles while revolving about and sliding along the
axis at uniform rates. In other words, the right helicoid is a spiral ramp.
A right helicoid about the z-axis has the parametric equations,
(4) x = u cos v, y = u sin v, z = as,
where u, v are plane polar coordinates p, p in the xy-plane. The lines v =
const are the horizontal rulings on the surface; the lines u = const are circular
helices on the cylinder x2 + y2 = u2.
Equations (4) give the vector equation,
(5) r = u R(v) + av k,
where R is a unit radial vector (R . k = 0). Now
ru = R, r = uP + ak;
n- ruxr uk - aP
=
_ I ru x rv I a2

a a
nu = (u2 + a2)2 rv, nv = (u2 + a2)2 ru.

Since ru r, = 0, the parametric lines are orthogonal; and, since ru . nu = 0,


r n = 0, they are also asymptotic lines. From nu x n = K ru x r (132.20),
we have
- a2
K= (u2 + a2)2

and, since nu x r + ru x n = 0, J = 0 (132.19). The right helicoid is a mini-


mal surface. Its negative total curvature is constant along any helix u - const.
Along the lines of curvature,
(rudu +rvdv) x (nudu ruxnudu2 +rvxnndv2 = 0,
or du2 - (u2 + a2) dv2 = 0. The two families of these lines therefore satisfy
the differential equations:
du/ u2 + a2 = fdv.
On integration, these give
u
sinh-1- = c f v or u = a sinh (c f v)
a
as the finite equations of the lines of curvature.
Example 3. On a minimal surface, the asymptotic lines form an orthogonal
system (§ 131, ex. 2). If we choose them as our field curves, k1 = k2 = 0,
and the Codazzi Equations (137.13), (137.14) become

27211 + i = 0, 2y1t1 - -
ds2
= 0.
144 SUMMARY: SURFACE GEOMETRY 319

Since K = -ti,
dK = -2t1-
dt1
_ -472K,
dK = -2t1-
dt1
= 4y1K;
dsl ds1 ds2 ds2
dK dK
VK = el - + e2 - = 4K(y1e2 - 72e1) = 4K n x R (134.3).
dsl ds2

144. Summary: Surface Geometry. On a surface r = r(u, v) of


unit normal n, the angular velocity of the trihedral Tnp (p =
T x n) along a curve is w = tT + yn + kp. If (p = angle (N, n),
positive in the sense of T,
t = r+ dv/ds, y= K sin cp, k = K cos V
are the geodesic torsion, geodesic curvature, and normal curvature, re-
spectively. From
do
(1) =wxn= -kT - tnxT,
ds
we conclude that k and t are the same for all surface curves having
a common tangent at a point.
Important surface curves and their differential equations:
dr do
Lines of curvature (t = 0) : -x-
ds ds
= 0;
dr do
Asymptotic lines (k = 0) : -ds-ds= 0;
d2r d2r
Geodesics (y = 0) : nx = 0 or Vr = 0.
ds2 ds2
The equations for geodesics follow from
dT
-xn=KNxn=yT.
ds
If e and e' = n x e are perpendicular directions in the tangent
plane,
-Vn = k ee + k' e'e' + t ee' + t' e'e.
Since Rot n = 0, Vn is symmetric, and t + t' = 0. The first and
second scalar invariants of - Vn,
J = - Div n = k + k', the mean curvature;
K = kk' + tt' = kk' - t2, the total curvature,
320 GEOMETRY ON A SURFACE § 144

have the same value for any pair of perpendicular directions. If


the point is not an umbilic (k constant for all directions), k attains
its extreme values in the orthogonal principal directions e1, e2 for
which t1 = 0, t2 = 0. In terms of the principal curvatures k1, k2,
J = k1+'k2iK=k1k2.
If a, b, n are reciprocal to ru, r,,, n, the first and second funda-
mental forms are
Or = E as + F(ab + ba) + G bb,
-Vn = Las+M(ab+ba) -}-Nbb;
and the fundamental quantities are given by

L = -ru nu, M = -ru nv = -rv nu, N = -rv nv.


If e1i e2 = n x e1 are the unit tangent vectors along a field of
surface curves and their orthogonal trajectories, we have
Vet = Re2 - Pn, Ve2 =-Re1 - Qn, Vn = Pet + Qe2,
where
P = do/ds1, Q = dn/ds2, R = y1e1 + y2e2.
Codazzi Equations: n V x (Vn) = 0; or

Gauss Equation: n V x (De1) = 0; or


n rot R = -K.
For any field of curves of tangent vector e,
y = n- rot e.
The last two equations show that y and K may be expressed in
terms of E, F, G and their partial derivatives.
When embedded in a geodesic field, the arc of a geodesic is
shorter than any other curve lying within the field and having the
same end points.
If a field of curves of tangent vector e cuts the field e1 at the
angle 0 = (eli e), positive in sense of n,
R = y1e1 + y2e2 = ye + y'e' - V0.
PROBLEMS 321

Bonnet's Integral Formula,

fK d S= fdo - ds,

is the integral ("Stokian") equivalent of the Gauss Equation,


n - Rot R = -K. If the simple closed curve C has a continu-
ously turning tangent fdo = 27r, but if the path has angles (in-
terior angles ai),
fdo = 2ir - - as).

If C is a geodesic triangle, fK dS = al + a2 + a3 - ir; and, on


a surface of constant K 3-1 0, the area of a geodesic triangle is
(al + a2 + a3 - ir)/K. A sphere of radius a has the constant
positive curvature 1/a2; a tractrix of revolution for which the
tangential distance to the asymptote is a has the constant negative
curvature -1/a2.
A minimal surface (J = 0) embedded in a field of minimal sur-
faces and spanning a closed curve C has a smaller area than any
other surface lying within the field and spanning C. Soap films
spanned over wire framework materialize minimal surfaces. The
right helicoid is a ruled minimal surface. The catenoid is the only
minimal surface of revolution.

PROBLEMS
1. If u, v, are plane polar coordinates r, 0, prove that the surface obtained by
revolving the curve z = f(x) about the z-axis has the parametric equations:
x = u cos v, y = u sin v, z = z(u).
2. A straight line, which always cuts the z-axis at right angles, is revolved
about and moved along this axis. The surface thus generated is called a conoid.
If u and v are plane polar coordinates in the xy-plane, show that the conoid
has the parametric equations:
x = u cos v, y = u sin v, z = z(v).

In particular when dz/dv is constant the conoid is a right helicoid:


x = U cos v, y = u sin v, z = av.
3. The central quadric surface,
x2 y2 z2 _
a2 ± b2 ± 1,
322 GEOMETRY ON A SURFACE
is an ellipsoid, a hyperboloid of one sheet, or a hyperboloid of two sheets according
as the terms on the left have the signs (+, +, +), (+, +, -), (+, -, -).
Show that the corresponding parametric equations are
x = a sin u cos v, y = b sin u sin v, z = c cos u;
x = a cosh u cos v, y = b cosh u sin v, z = c sinh u;
x = a cosh u cosh v, y = b cosh u sinh v, z = c sinh u.
4. The paraboloid,
x2

a2
t -=-,
y2

b2
2z

is elliptic or hyperbolic, according as the terms on the left have the signs (+, +)
or (+, -). Show that the corresponding parametric equations are
x = au cos v, y = bu sin v, z = Zcu2;
x = au cosh v, y = bu sinh v, z =. 4cu2.

5. The position vector,


r = f(u) + g(v),
traces a surface of translation. Show that any curve u = a (const.) may be
obtained by giving the curve rl = g(v) a translation f(a); and that any curve
v = b may be obtained by giving the curve r2 = f(u) a translation g(b).
What are the surfaces,
r=f(u)+bv, r=au+bv+c?
6. Show that a right circular cylinder of radius a has the constant mean
curvature J = 1/a.
7. For the surface xyz = a3 show that
3a6
Jr = 2a3(x2 + y2 + z2)
(X2y2 + y2z2 + 22x2)2 ' (x2y2 + y2z2 + 22x2)7

[Cf. § 132, ex. 2.]


8. For the elliptic paraboloid,
x2/a2 + y2/b2 = 2z/c
prove that the total curvature K 5 c2/a2b2.
9. For the surface of revolution about the z-axis,
r = iu cos v + ju sin v + kz(u),
show that
E=1+z'2 F=O, G=u2;
L= 2'2)1, M = 0, N = uz'/(1 + 2'2)1. [Cf. § 132.1

From (132.21) and (132.22) deduce the values of K and J and from (132.25)
find the principal curvatures. Check with (138.4) and (138.5).
PROBLEMS 323

10. For the conoid,


r = iu cos v + ju sin v + kz(v),
show that
E=1, F=0, G=u2+z'2=H2;
L = 0, M = -z'/H, N = uz"/H;
K = -z'2/H4, J = uz"/H3. [Cf. § 132.1

When z = av, show that the resulting right helicoid is minimal (§ 143);
and that its principal curvatures are fa/(u2 + a2).
11. Show that the curvatures J and K have the dimensions of (length) -1
and (length) -22. Verify this in Problems 6, 7, 8, 9, 10.
12. The position vector of a twisted curve r is given as a function ro(s) of the
are. The surface generated by its tangents is
r=rc(v)+uT(v)
where v = s and u is the distance along a tangent measured from r. Show that
this tangent surface has the curvatures K = 0, J = z/i u Jx. What are the
principal curvatures?
13. If a parabola is revolved about its directrix, show that the principal
curvatures of the surface of revolution satisfy the relation 2k1 + k2 = 0.
[With the notation of § 138, ex., the parabola has the equation z2 _
4a(u - a) when the directrix is the z-axis and u = 2a at the focus.]
14. The vectors a, b, n and ru, r,,, n form reciprocal sets of vectors over the
surface r = r(u, v) [Cf. § 95]. Hence show that any vector f(u, v), defined over
the surface, may be written
f
15. Prove that

H2a = Gr - Fr,,, H2b = -Fru + Er,,.


[Use Prob. 14 and (132.18).]
16. Prove that
FM - GL FL - EM
H2 ' 'b= H2 ;

FN - GM FM - EN
II, a I
H2 H2
[Use Prob. 15.]

17. Prove the derivative formulas of Weingarten:

HZnu = (FM - GL)ru + (FL - EM)r,,,


H2n _ (FN - GM)ru + (FM - EN)r,,.
[Use Prob. 14 and Prob. 16.]
324 GEOMETRY ON A SURFACE
18. Show that the asymptotic lines of the surface r = r(u, v) have the dif-
ferential equation,
(ru du + r dv) (nu du + n dv) = 0
[Cf. (130.10)]; or, in terms of fundamental quantities,
Ldu2+2Mdudv+Ndv2 = 0.
19. Prove that the asymptotic lines on a right helicoid (Prob. 10) are the
parametric curves and form an orthogonal net.
20. For the ruled surface,
r = p(u) + ve(u), (142.1).
prove that M = pu e x eu/H, N = 0.
21. Prove that a ruled surface is developable when, and only when, M = 0.
[Cf. (142.2).]
22. Prove that the asymptotic lines of a developable surface, not a plane, are
the generating lines (counted twice). [Differential equation: du 2 = 0.]
23. Prove that, along a curved asymptotic line,
(a) The osculating plane is tangent to the surface.
(b) The geodesic torsion equals the torsion: t = r.
(c) The geodesic curvature is numerically equal to the curvature: y = ±K.
(d) The square of the torsion is equal to the negative of the total curvature:
2 = -K.
24. Prove that
(a) If an asymptotic line is a plane curve other than a straight line, it is
also a line of curvature.
(b) An asymptotic line of curvature is plane.
25. Prove that the following conditions are necessary and sufficient in order
that the surface be
(i) A plane: L = M = N = 0.
(ii) A sphere: E/L = F/M = GIN.
26. Show that the lines of curvature of the surface r = r(u, v) have the
differential equation,
n (r, du + r dv) x (nu du + n,, dv) = 0
[Cf. (138.1)]; or, in terms of fundamental quantities,
I dv2 -du dv du2 I

E F G = 0.

L M N
27. Show that the parametric curves are lines of curvature when, and only
when, F = 0, M = 0.
28. In the notation of § 132, ex. 2, prove that the lines of curvature on the
surface z = z(x, y) have the differential equation,
dy2 -dx dy dx2

1 + p2 pq 1 + q2 = 0.
PROBLEMS 325

29. Show that the hyperbolic paraboloid of Prob. 4 has the parametric
equations:
x = 2a(u + v), y = 2b(u - v), z = cuv.
Prove that the differential equations of its asymptotic lines and lines of
curvature are, respectively,
dvl2 a2 + b2 + c2y2
du dv = 0,
\du / a2 + b2 + c2u2
and find the uv-equations of the asymptotic lines and lines of curvature.
30. If the parametric curves on a surface are its lines of curvature, show that
nu = -kiru, nv = -k2rv.
31. Two surfaces that have the same normal lines are called parallel. Sur-
face points on the same normal are said to correspond; thus
r(u, v) = r(u, v) + An(u, v)
determines corresponding points on the parallel surfaces S and S. Prove that
the distance X between the parallel surfaces is constant.
[If we take the lines of curvature of S as parametric curves,
ru = (1 - Aki)ru + Xun, rv = (1 - Xk2)rv + Xn
(Prob. 30); and, on cross multiplication,
ffn- = (1 - Xki)(1 - Ak2)Hn + Xu(1 - Xk2)n x rv + X ,(l - Xkl)ru x n.
Now ii = En where e = ±1; and as n, ru, r are mutually orthogonal,
H = E(1 - Xki)(1 - Xk2)H, Xu = Xv = 0.1
32. Prove that the lines of curvature on parallel surfaces correspond and
that the corresponding principal curvatures satisfy the equations,
_ Ekl Eke
kl T2 =
1 - Xkl ' 1 - xk2
Hence the mean and total curvature of S are given by
_
J _ 1rJ+X2K'
E(J - 2XK) K
K
- 1-xJ+x2K
33. Show that surfaces parallel to a surface of revolution are also surfaces
of revolution.
34. Prove the "Theorema egregium" of Gauss by establishing Baltzer's
formula for the total curvature K:
Fuv - UGuu -Evv 2Eu Fu - 2Ev 0 U Ei, Gu
H4K=I Fv--2Gu E F ZEv E F
F G -'Gu F G
326 GEOMETRY ON A SURFACE
Method. From (132.10) and (132.22) show that
H4K = [ruurx][rv,;r r,;] - [ruvrurv]2.
Using (24.14), the right member becomes
ruu rvv ruu ru ruu rv ruv ruv ruv ru ruv rv

ru rvv E F ruv ru E F
rv rvv F G ruv rv F G
In both determinants the upper left element has the cofactor EG - F2; hence
we may replace these elements by ruu rvv - ruv ruv and 0, respectively.
Now, by differentiating

we may show that


rut,, ru
i
2F!u, ruv ru = 2iEv, rvv ru = F,, i2Gu,

rvv rv = zGv, ruv rv = 2Gu> ruu r,, - Fu - 2Ev;

ruu rvv - ruv ruv - Fuv 1Guu - 2!Evv.


vv
35. On the surface r = r(u, v) let us write

u1 = u, u2 = v; D1 = aau , D2 = ava ;

el = ru, e2 = rv; el = a, e2 = b; then e" ep = SQ,

where the Greek indices have the range 1, 2. Moreover let


911 912
gap = ea ep, 9ap = ea e$, g = det gap =
921 922

hap = n DaDpr = hpa; h = det ha p.


Prove that
(a) 9u= E,g12=F,922=G;g=EG-F2.
(b) 911 =
922 g12 - -912 922 = 922

9 9 9
that is, g"p is the reduced cofactor of gap in det gap.
(c) h11=L,h12=M,h22=N;h=LN-M2;
(d) Daep - Dpea;
(e) Dagpy + Dog-y. - D.rgap = 2ey Daep.
(f) Daep = ey(Dagp , + Dpgya - Dygap) + nhap (summed over y = 1, 2).
(g) e" = g"pep (summed over (3 = 1, 2). [Cf. § 145.]
36. The Christoffel symbols Papa, g, y = 1, 2) for a surface r = r(ul, u2)
having gap du" dup as first fundamental form (§ 132) are computed from the
equations:
Pap = ey . Daep.
PROBLEMS 327

With reference to Prob. 35 prove that


(a) ra0 = r'a
(b) rP# = Diggya - Dyga#) (summed over p = 1, 2).
(c) 29r1i = 922 D1911 - 2g12 D1912 + 912 D29uu,
29ri2 = 922 D2911 - 912 D1922,
29r22 = -912 D2922 + 2g22 D2912 - 922 D1922
[Use the formula of part b; for example,
21'1ii = 911(D1911 + D1g11 - D1911) + 91Y(D1912 + D1921 - D2911),
1
29t1 = 922 Dig,, - 912(2D1912 -
(d) From the formulas for r,# obtain the formulas for rap.
(e) From Prob. 35 (f) prove the derivative formulas of Gauss:
X
Daep = raryea + heron (summed over a = 1, 2).
37. From n efl = 0 show that eo Dan = -han. Hence prove the deriva-
tive formulas of Weingarten;
Dan = -hafieO (summed over S = 1, 2).
Check these results with Prob. 17.
38. With the notations of Prob. 35 show that equations (132.21) and (132.22)
may be written
J = g0h,,#, K = h/g (a, 0 summed over 1, 2).
39. Show that the tangential projection of the vector curvature dT/ds = KNi
of a surface curve is
d2r dua duDr

= ey (--
ds2 + raA ds ds
summed over a, 0, y = 1, 2. Hence prove that the equations of a geodesic
on a surface are
y dua B
+ rad =0 (Y = 1, 2).
d3 ds ds
a 8
1 T = ep ds , = ep ds2 + Daeg d , Vr = eyes. Cf. Prob. 36. ]
CHAPTER IX

TENSOR ANALYSIS

145. The Summation Convention. Expressions which consist


of a sum of similar terms may be condensed greatly, without any
essential loss of clarity, by indicating summation by means of re-
peated indices. This usage, originally due to Einstein, is stated
precisely in the following
SUMMATION CONVENTION. Any term, in which the same index
(subscript or superscript) appears twice, shall stand for the sum of
all such terms obtained by giving this index its complete range of
values. This range of values, if not understood, must be specified
in advance.
By way of illustration, we repeat some of the formulas of § 24,
using the summation convention. The index range is 1, 2, 3.
The two forms of a vector (24.1) and (24.2) become
(1) U = uiei, u = u;e'.
In these equations i and j are summation (or dummy) indices. But
any other letter will do as well; thus
u = carer = ule1+ u2e2 + u3e3.

In order to compute u - v, we first recall the defining equation


for reciprocal sets,
(2) ei e' = Si (23.3).
Now
(3) u . v = (uiei) . (v,e') = uiv; 'Vi' = uiei;
we here use different indices in expressing u and v in order to get
the 32 = 9 terms in the expanded product (six are zero). Simi-
larly,
(4) u v = (uiei) . (vie;) = uiv' S = u1vi.
328
§ 146 DETERMINANTS 329

Note that the summation indices in the preceding examples ap-


pear once as subscript and once as superscript. The significance
of this arrangement (not required by the summation convention)
soon will be apparent.
146. Determinants. A permutation of the first n natural num-
bers is said to be even or odd, according as it can be formed from
123 n by an even or odd number of interchanges of adjacent
numbers. The total number of permutations of n different num-
bers is n!; one half of these are even and one half odd. For ex-
ample, when n = 3, 123, 231, 312 are even permutations, 213, 132,
321 are odd.
Consider now a permutation ijk r of the numbers 123 n.
We then define the permutation symbols ijk...r and as eijk""r

equal to 1 or -1, according as ijlc . r is an even or an odd per-


mutation of 123 . . . n; but, if any index is repeated, the epsilon is
zero. For example, when n = 3,
E123 = 231 = 312 = 1 E213 = 132 = 321 = - 1

112 = 122 = 222 = 0.

The epsilons just defined are useful in dealing with determi-


nants. To be specific, we shall take n = 3; but all the formulas
apply without change of form for any value of n.
From the definition of a determinant :
1 2 3
a1 a1 a1
ei jka 12a3
2
(1) a = det a _ as a2
3
a2
1
a3
2
a3
3
a3
e'k ai1 aj2ak3

The implied summations on i, j, k produce 33 = 27 terms; of these,


21 involve repeated indices and therefore give a zero epsilon,
whereas 6 involve permutations of 123 and give precisely the
3! = 6 terms of the determinant. (Write them out!)
The theorems relative to an interchange of rows and columns
are given by i '

(2) eijkarasat
aer8t = ijkarL
a8
atk

Proof. When r, s, t are 1, 2, 3, equations (2) reduce to (1).


Since rst changes sign when two adjacent indices are interchanged,
330 TENSOR ANALYSIS § 146

(2) will beestablished when the right members are shown to have
this same property. Now
Eijkarasat = eijkasarat = - ejikasa r aki

and, if we interchange the summation indices i, j, the last expres-


sion becomes - eijkasajat .
If we multiply the first equation of (2) by e" and sum on rst,
we obtain
eijkerstarasae.
(3) 3! a =
On the left er8tergt, summed over the 3! permutations of 123, equals
3!; on the right, the summation extends over all six indices and
produces 36 = 729 terms. Many with zero epsilons vanish; and
each non-vanishing term such as a1la22a3 3 appears six times, corre-
sponding to the 3! permutations of its factors. Equation (3) thus
is not useful for computation; its importance rests on the informa-
tion it gives about the determinant when its elements are trans-
formed.
The cofactor of an element az in the determinant (1) is defined
as its coefficient in the expansion of the determinant. The co-
factor of a; is denoted by A'. To find A',* strike out the row and
column in which ai stands; then A equals the resulting minor
taken with the sign (-1) i+j. If the elements of any row or col-
umn are multiplied by their respective cofactors and added, we
obtain the determinant-its Laplace expansion; but, if the ele-
ments of any row (column) are multiplied by the corresponding
cofactors of another row (column) and added, the sum is zero.
These important properties both are included in
(4) aiA; = a'Ai = a S .
Here r is the summation index. If in (4) we put j = i, both r and
i are summation indices, and we get
(5) a7AT = a; A, = a Si,
where S = Si + S2 + S3 = 3 (n = 3). But, if we put j = i in
(4) and suspend summation on i, we obtain the Laplace expansions,
(6) a7,Ar = arAi = a (i fixed),
for S = 1 for a fixed i.
§ 146 DETERMINANTS 331

A cofactor divided by the value of the determinant is called a


reduced cofactor. The reduced cofactor of a2 is therefore
(7) a; = A,/a (a F6 0).
In terms of reduced cofactors, equations (4) become
(8) aTc = 8a.
When the elements of a determinant are written aid, its defini-
tion becomes
(9) a = E'.'kalia2Ja3k = E1Jkailaj2ak3-
The reduced cofactor of ai3 then is written ai and equations (8)
become
(10) aira'r = aria'' = SL.

If the elements ai; are functions of a variable x, we have, from


(9),
da _ ilk (dali dal; da3k\
E a2ja3k + a1i a3k + aiia2i
dx = x x dx )
dal-Ali+da2jA21+da3kA3k
=
dx dx dx '
where A 'j denotes the cofactor of aij. Summing on two indices,
we may write

(11) -da= daii


d.r, dx
Ai' (i, j = 1, 2, 3).

The derivative of a determinant is the sum of the products formed by


differentiating each clement and multiplying by the cofactor of the
element.
The properties (8) of reduced cofactors enable us to solve a
system of linear equations when the determinant of the coefficients
is not zero. Consider, for example, the equations:
(12) ax' = yt (i, j = 1, 2, 3; a 0).
To solve these for x', multiply (12) by the reduced cofactor ak
and sum on i : the left member becomes S; x' = xk, and we obtain
xk = a;yi, or, on replacing k by j,
(13) x' = aiyi (i, j = 1, 2, 3).
332 TENSOR ANALYSIS 5 147

To find the product of the determinants a = det a, b = det b;,


we have, from (1) and (2),
(14) ab = EiJkaia2a3b

s t
ia2 ja3 kEr8tbsrbjbk
= a1

= Erst(albi)(aabj)(a3kbk)

r 8
=
ErstC1C2Ct3,

where
(15) c% = aabj = aib1 -F a2b2 -f- a2b33.

From (15), we see that the element in the ith row and jth column
of the product ab is given by the sum of the products of the corre-
sponding terms in the ith row of a and the jth column of b-the
so-called "row-column" rule.
Since the value of a determinant is not altered by an interchange
of rows and columns, we also can compute ab by "row-row,"
"column-row," and "column-column" rules. For example, the
product
1 2 3 1 1 1
a1 al al a1 a2 a3 1 0 0
2
(16) a2 a2 a2 ai a2 a3 0 1 0 = 1.
1 z 3 3 3 3
a3 a3 a3 a1 a2 a3 0 0 1

by use of the row-row rule.


THEOREM. If a determinant a 0, the determinant formed by re-
placing each element by its reduced cofactor is 1/a.
If we solve equations (13) for yi by using the reduced cofactors
in det a, we will obtain (12); in other words, the reduced cofactor
of a in its determinant is a. The two determinants in (16) thus
are reciprocally related: each is formed from the reduced cofactors
of the other. .'.
147. Contragredient Transformations. Let us now introduce a
new basis e1i e2, e3 by means of the linear transformation,
61 = cie1 + ciez + Cie3,

(1) e2 = c2e1 + c2e2 + 4e3,

e3 = c3e1 + c3e2 + c3e3,


§ 147 CONTRAGREDIENT TRANSFORMATIONS 333

where the coefficients ci are real constants whose determinant


c 0. In brief,
(1) ei = c'ej, c = det ca /- 0.
The condition c 0 ensures that e1, e2, e3 are linearly independ-
ent. For, if there were three constants At such that
Aiei = 0, then Aicej = 0;
the linear independence of the vectors ej requires that c'Ai = 0,
and, since c P 0, these equations only admit the solution Ai = 0.
This argument applies without change to space of n dimensions,
in which a basis consists of n linearly independent vectors.
In the present case (n = 3), we also may argue as follows.
From (1), we have
161
. e2 x e3 = cic2c3 ei ej x ek = C'iC2iC3Eijk e1 ' e2 x e3;

hence, on writing E = el e2 x e3, we have


(2) E _ (det c')E = cE.
Since the vectors ei form a basis E 96 0, hence c 0 implies
E 0 and the linear independence of the vectors e1.
For any vector u,
(3) u = 'uiei = ujej;

hence, on substitution from (1),


ujej,
or, since the vectors ej are linearly independent,
(4) u.l = Ci,u

If y is the reduced cofactor of c in the determinant c, we have


on solving equations (4),
(5) ut =
or, written out in full,
ul = riu1 + 'Y2u2 + Y3u3,
(5) u2 = 'Ylui + Y2u2 + Y3 u3,
u3 = 'iu1 + Y32 u2 + 3
'Y3u3.
334 TENSOR ANALYSIS § 148

The linear transformations (1) and (5) are said to be contra-


gredient. Their matrices,
Cl C1 c1 71 72 73
3 2
C= 1
C2
2
C2 C2 , r = 71 72
2
73
2 3 3 3 3
C31 C3 C3 71 72 73
are so related that any element of r is the reduced cofactor of the
corresponding element of C in its determinant c; and any element
of C is the reduced cofactor of the corresponding element of r in
its determinant y.
In order to state this definition analytically, we remind the
reader of the following definitions from matrix algebra. The prod-
uct AB of two square matrices A and B is the matrix whose element
in the ith row and jth column is the sum of the corresponding
products of the elements in the ith row of A and jth column of
(row-column rule). The transpose of a matrix A is a matrix A'
obtained from A by interchanging rows and columns. The unit
matrix I has the elements Si (ones in the principal diagonal, zeros
elsewhere).
If we compute Cr' or rC' and make use of the equations,
(6) cz7r = S, 7rc1 = bj,
we find that
(7) Cr'=rC'=I.
These equations characterize contragredient transformations.
Two matrices whose product is I are said to be reciprocal. Thus
C and r' or C' and r are reciprocal matrices. From (7), we see
that contragredient matrices are so related that the transpose of either
is the reciprocal of the other.
148. Covariance and Contravariance. When new base vectors
are introduced by means of the transformation,
(1) ei = c e5, det c 0.

the components of any vector u are subjected to the contra-


gredient transformation,
(2) ui = 7;u'.
In view of (24.3), this may be written
u-ei=7jue';
§ 148 COVARIANCE AND CONTRAVARIANCE 335

and, as this holds for every vector u,


(3)

The basis ei, reciprocal to ei, thus is transformed in the same way
as ui. To find the transformation for the components ui, multiply
(1) by u ; then, from (24.4),
(4) ui = cu
a transformation cogredient with (1). Thus quantities written
with subscripts transform in the same way (cogrediently); and
quantities written with superscripts also transform in the same
way-but the latter transformations are contragredient to the former.
Thus the position of the index indicates the character of the
transformation.
A vector u is said to have the contravariant components ui, the
covariant components ui. These terms suggest variation unlike and
like that of the base vectors ei; in other words, the transformation
(1) is regarded as a standard. Thus [u', u2, u3] and [U1, u2i u3]
are two ways of representing the same vector u; in the first it is
referred to the basis ei, in the second to the reciprocal basis e'
[u1, u2, u3J often is called a contravariant vector, [u1, u2, u3] a co-'
variant vector. Actually, the vector u is neither contravariant or
covariant, but invariant.
By using the properties of the reduced cofactors,
eir'Yr
i r i
(5) = Yi = ai,
we may solve equations (1) to (4) for the original base vectors and
components (§ 146). Thus the equations,
(6) ei = c,ei, ui = ciuj, et = 'Yie', uz - yiu',
have the solutions,
(7) ei = -?j-6j, ui = Yiui; e' = c ,%J, ut = c ui.
Note that the matrices c, y in (7) are the matrices cz, -y' of (6)
transposed. To be quite clear on this point the reader should
write out equations ei = (.'e; and et = c;ei in full.
From (1) and (3), we find that
ei ' e' = cier yie8 = ciYaSr = c 'Yr = 31j,
which shows that the new bases have the fundamental property
of reciprocal sets.
336 TENSOR ANALYSIS § 149

An expression, such as uiv i, uie i, ei e i, summed over the same


upper and lower indices, maintains its form under the transforma-
tion. Take, for example, the scalar product uivi: making use of
(2) and (4), we have
uiv2 = Ciur78vs = bryurv8 = u'1T.

The index notation thus automatically indicates quantities that


are invariant to affine transformations of the base vectors.
149. Orthogonal Transformations. When the basis ei and the
new basis ei both consist of mutually orthogonal triples of unit
vectors, the transformation,
(1) ei = ciei,
is called orthogonal. Since both bases are self-reciprocal (§ 23),
the corresponding transformation (148.3) between the reciprocal
bases,
ei = 7e', now becomes ei = 7e;.
Since this transformation must be the same as (1),
C11 ci ci 7i 72 73
7sq =r
2
(2) C= c2 c2 c3 = 71 72
1 3 3 3 3
C3 C3 C3 71 72 73

The matrix of an orthogonal transformation is called an orthogonal


matrix. From (2), we have the
THEOREM. Every element of an orthogonal matrix is its own re-
duced cofactor in the determinant of the matrix.
In view of the properties of reduced cofactors, the coefficients
of an orthogonal transformation satisfy the equations:
(3) crc; = Cic; = 8L.
The matric equation (147.7) characterizing contragredient trans-
formations becomes
(4) CC' = I
for orthogonal transformations. Thus a real matrix is orthogonal
when its transpose equals its reciprocal.
If c denotes the determinant of C, we have, from (4),
(5) c2 = 1, c = f 1.
§ 150 QUADRATIC FORMS 337

If the bases ei and ei are both dextral or both sinistral, we can


bring the trihedral e1e2e3 into coincidence with e1e2e3 by continu-
ous motion-in fact, by a rotation about an axis through the
origin. At any stage of the motion, the base vectors are related
by equations of the form (1); and, as the motion progresses, the
coefficients c change continuously from their initial to their final
values S , and c = det ci becomes det S = 1. But, since this de-
terminant equals =E1 at all stages of the motion and must change
continuously if at all, c = 1 when the bases have the same orien-
tation.
If one basis is dextral, the other sinistral, e1e2e3 and ele2(-e3)
have the same orientation. Hence the determinant of the trans-
formation,
el = c1er, -e3 = -caer,
e2 = C3er,
namely, -c, must equal 1; thus c = -1 when the bases have dif-
ferent orientations.
150. Quadratic Forms. A real quadratic polynomial,
(1) A(x, x) = aijxixj (i, j = 1, 2, , n),
for which aij = aji is called a real quadratic form in the variables
x1, x2, , x"`. The symmetry requirement aij = aji entails no
loss in generality; for, if aij 0 aji, the form is not altered if we
replace aij and aji by (aij + aji).
The determinant of2 the coefficients, a = det aij, is called the
discriminant of the form. The form is said to be singular if det aij
= 0, non-singular if det aij 0 0.
Associated with the quadratic form A (x, x) is the bilinear form,
(2) A(x, y) = aijx'y' = A(y, x),
known as its polar form.
The expansion of aij(xi + Xyi) (x' + Xyj) shows that a quad-
ratic form and its polar are related by the identity,
(3) A (x + X y, x + Xy) = A (x, x) + 2XA (x, y) + X2A (y, y).
If we make the linear transformation,
(4) xi=c yr, c=detc;F-- 0,
the form (1) becomes
(5) B(y, y) = br4yry8 where bra = aijcTc'.
338 TENSOR ANALYSIS §150

From the rule for multiplying determinants,


det bra = det (ai;cT) det c8 = det ai; det cr det c8;
the discriminant of B(y, y) is therefore
(6) b = det bra = c2a.
Linear transformations of non-zero determinant do not alter the singe
lar or non-singular character of a quadratic form.
If the form A (x, x) is singular, it can be expressed in terms of
fewer than n variables which are linear functions of x. For since
det ai, = 0, the system of n linear equations ai;xi = 0 has a solu-
tion (xa, xo, , xo) which does not consist entirely of zeros; thus
we may assume that xo 5-4- 0. Now A(xo, y) = ai;xoy' = 0, and,
from (3),
A (x + Axo, x + Axo) = A (x, x),
irrespective of the value of A. If we now choose A = -xl/xo
and write
y i = x, i + Axp = x i - x4x1 /x0,
we have A (y, y) = A (x, x). Since y' = 0, A (y, y) is expressed in
, y".
terms of the n - 1 variables y2, y3,
Henceforth we shall consider only non-singular forms. A non-
singular form is said to be definite when it vanishes only for
xl = x2 = = x" = 0. For all other sets of values, the sign of
a definite quadratic form is always the same. To prove this, sup-
pose that A(x, x) > 0, for the set x1, x2, , x" and A(y, y) < 0

for the set y', y2, , Then, from (3),


y".

(7) A (x + Ay, x + Ay) = 0


is a quadratic equation in A having two distinct real roots A1i A2; for
{A(x, y) 12 - A(x, x)A(y, y) > 0.
Consequently, the form vanishes for two different sets of values,
xi + Alyi and xi + A2yi and therefore cannot be definite.
For a definite quadratic form, the quadratic equation (7) in A
must have a pair of either complex roots or equal roots; hence, for
a definite form,
(8) JA (x, y) 12 - A(x, x)A(y, y) < 0,
the equal sign corresponding to the case xi + Ayi = 0, (i = 1.
, n), in which the sets xi and yi are proportional.
§ 152 RELATIONS BETWEEN RECIPROCAL BASES 339

A definite quadratic form is called positive definite or negative


definite according as its sign (for non-zero sets) is positive or nega-
tive. A non-singular form which is not definite is called indefinite:
such a form may vanish for values x` other than zero.
When A (x, x) is positive definite, it is well known that we can
find a real linear transformation (4) which will reduce A (x, x) to
a sum of squares:
(9) B(x, x) = Siyiyj = ylyl + y2y2 + ... + ynyn
The discriminant of a positive definite form is therefore positive;
for, from (6), b = 1 = c2a, and a = 1/c2 > 0.
151. The Metric. Using the notation,
(1) 9i; = ei ej = gji,
for the nine scalar products of the base vectors, we have
(2) u . v = (uiei) . (viej) = giptivj, u u = gi;uiu'.
Thus u u is a real quadratic form,
giiU'u1 + g22u2u2 + 933u3u3 + 2g12u'u2 + + 2931u3u1r
2923u2u2

in the variables u1, u2, u3; and, since u u > 0 when u 0, this
form is positive definite. Moreover, u v is given by the associated
bilinear (polar) form in ui, vj.
From (2), we have
(3) I
UI= gijl(itljr
UV _ gijuw'
(4) Cos (u, V) =
lUl VI VgijuiujVVgijvivi
Thus, when the six constants gij are known, we can find the
lengths of vectors and the angles between them when a fixed unit
of length is adopted. The quantities gjj, since they make measure-
ments possible, are said to determine the metric of our 3-space; and
gijuiuj is called the metric (or fundamental) quadratic form.
152. Relations between Reciprocal Bases. The discriminant c'
the metric form is
911 912 913

(1) 9= 921 922 923

931 932 933


:" Bother, Introduction to Higher Algebra, New York, 1907, p. 150.
340 TENSOR ANALYSIS § 153

The reduced cofactor of gij in this determinant is denoted by gij;


and we have, from (146.10), the relations:
= 9rigr' =
girg.r

(2) Si.

Since any vector may be written

u= ei,
(3) ei = gije'
We may solve these equations for ej by multiplying by gik and
summing on i; thus
k j = e k,
9ikei = 9ikgijej = 3je
or, if we interchange i and k,
(4) ei = 9kiek-
From (4), we have also
ei ej = gkiek ej = gkiak = gji

(5) gij = ez
e' = 9'i
Making use of (3) and (4), we now have, for any vector u,
(6)

(7) ui = ei u = gijuj
The equations enable us to convert contravariant components of a
vector to covariant and vice versa.
Finally, from (24.14), we have
(8) g = det gij = det (e1 ej) = [eie2e312 = E2,

(9) det gij = det (ei ej) = [ele2e312 = E-2 = 1


9
153. The Affine Group. When an origin 0 is given, the con-
stant basis el, e2, e3 defines a Cartesian coordinate system x1, x2,
x3; for any point P is determined by the components of its position
vector :
OP = x'el + x2e2 + x3e3.
The components xi of OP are called the Cartesian coordinates of P
relative to the basis ei.
§ 153 THE AFFINE GROUP 341

When the base vectors ei are subjected to the transformation,


(1) et=ce;,
the invariance of OP, namely,
(2) i''es = xZei,
induces a contragredient transformation on the coordinates. For,
if we substitute from (1) in (2), we find that x' = ci"xi; that is,
(3) xi = c;x'.
The matrix c? in (1) is transposed in (3); and, while (1) expresses
the new base vectors in terms of the old, (3) expresses the old
coordinates in terms of the new. We now can solve (3) for the
new coordinates by multiplying by the reduced cofactors ti; and
summing; thus we find xiy, = xk, or
(4) x' = y2xi, -y = det yj = 1/c.
The transformation (3) is called affine, or, more specifically,
centered affine, since the origin 0 has not been altered. The
centered-affine transformations with non-zero determinant form a
group, that is,
(a) the set includes the identity transformation:
ri = S ; a ; ' = x`, det 6 = 1;
(b) each transformation of the set has an inverse;
(c) the succession of two transformations of the set is equivalent
to a single transformation of the set.
Thus the transformations,
a x', x' = bj;xk, give xa = Ckx1,
where
ck = a bk and det ck = (det a) (det bk) F4- 0.
Our present point of view is that a transformation of the base
vectors induces a definite transformation of coordinates. But the
reverse point of view is adopted when transformation groups more
general than the affine are under consideration : the coordinates are
transformed, and we then inquire as to corresponding transformation
of the base vectors.
342 TENSOR ANALYSIS a 151

In the case of the affine group the transformation (3) of coordi-


nates obviously entails the transformation (1) of base vector.,.
For, if we put x' = x'cy in (2), we have
xi(czej - ei) = 0
for all xi, that is, (1) must hold.
154. Dyadics. Under the affine transformation,
(1) ei = c ei, xi = y;xi (c 0),
we have also
(2) ei = y;ei, ti = e xi (148.6).

Contravariant and covariant components of a vector, ut and ui,


transform like x i and xi, respectively. The sets of components u i
and ui often are called contravariant and covariant vectors. Actu-
ally they are different representations of the same invariant vector,
(3) U = uiei = uiei.
Nine (3 X 3) numbers Tii which transform like the nine prod-
ucts of vector components uivi, namely,
(4) Tii = yryBT'8

are called contravariant components of a dyadic. Similarly, nine


numbers Ti; which transform like uiei, namely,
(5) T,i = ('icjTra,
are called covariant components of a dyadic. Finally, nine num-
bers Ti; or T;i which transform like u ivi or uivi, respectively,
namely,
(6) 7,'i =
yrc;TB, T:' =
are called mixed components of a dyadic. Note that in Tti the
upper index comes first, in Tii the lower index.
Just as ui and ui are two representations of a single invariant
vector u, the components Tii, Tii, Ti, T;i are four representations
of one and the same dyadic,
(7) T = Tiieiei = Tiieiei = Ttieiei = Ti eiei.
Note that the indices on components and base vectors always are
placed in the same order from left to right. All forms of T given
§ 155 ABSOLUTE TENSORS 343

in (7) are invariant; for example, on using (1) and (4), we have
T'16x.e i ckek = 3,h6kTr8e1,ek
= YrYsjTrsc'eh
i r = Thkehe b,
from the properties of reduced cofactors (146.8).
The coefficients gij of the metric quadratic form are tensor com-
ponents; for, from (151.1),
(8) 9ij = ei ' ej = (tier) ' (cjes) = c'- jug,.
in agreement with (5). The same is true of gij; for, from (152.5),
9i7
(9) = e ' e' = (Yrer) - (Yses) = Y:Y39r8,
in agreement with (4). Again, if we define b, S transforms
as a mixed tensor,
(10) ei ej = (Yrer) - (c;es) = Yrc;a8.
Indeed, gij, gij and b are all components of the same tensor,
(11) gijeiej = gijeiej = beiej = 8jeiej = eiei = eiei,
namely, the idemfactor (§ 66). In fact, if we use the formulas,
(12) ei = girer, ei = girer (§ 152),

and remember the properties of gij, g" as reduced cofactors, any


two members of (11) can be shown to be equal; for example,
gijeie' = 9ij9irere' = b;ere-i = eje',
gije ej = gc,e gjrer = Sie er = e ei.
i i r ti i

Since eiei and eiei represent the same tensor, the order of the
indices in its mixed components, the Kronecker deltas, is im-
material.
Any component of T may be expressed in terms of components
of another type by means of the metric form. For example, we
have, from (7) and (12),
Tit = ei,T.ej = 9trer'T'ej = 9irT,J,
= 9'r ei'T -er =
girgjs er - T - e8 =
girgj8Trs
155. Absolute Tensors. We next define absolute tensors with
respect to the centered-affine group of transformations: xi = cxj
344 TENSOR ANALYSIS
Scalars p(x', x2, x3) which have one component in each coor(li-
nate system given by
(1)
gxl, x2, x3) = P(x', x2, x3)
are said to be absolute tensors of valence zero.
Vectors have three components, and these may be of two types
u,, ui. The laws of transformation,
(2) 2li = ciuj, ui = Yu
characterize absolute tensors of valence one. The vector itself,
u = uiei = uiei,
is invariant to the transformation: u = u.
Dyadics have 32 = 9 components, and these may be of 22 = 4
types: Ti;, Tij, T'j, Tij. If these components transform, respec-
tively, like the products ujv;, uivj, uiv;, uivj of components of two
absolute vectors, they are said to form absolute tensors of valence
two. The dyadic itself,

T = Tijeie' = Ti'eiej = TZjeie' = T'jetej,


is invariant to the transformation: T = T.
In general, an absolute tensor of valence m is a set of 3' compo-
nents that transform like the product of m absolute-vector com-
ponents; and, since each of these may be covariant or contra-
variant, the components are of 2' types. One of these types is
purely contravariant (T1 .k), another purely covariant (Tij...k);
the remaining types are mixed and have both upper and lower
indices.
Consider, for example, a tensor T of valence three (a triadic);
its 33 = 27 components may be of 23 = 8 types. Its covariant
components Ti;k will transform like uiv;wk, namely,

(3) Tijk = c2cjckT,.t,


and the mixed components Tijk like uivjwk, namely,
jk rjk at
(4) Ti = ea'Ya t Tr

The tensor itself,


T = T.;keiejek = Tijkeiejek = ,
§ 156 RELATIVE TENSORS 345

is invariant to the transformation; for example,


T = Tijkeiejek
(CiciCkTrst) (7ea) (' beb) (} cec)

SaSbS,'Trsteaebec

= Trstereset = T.
Note that the indices on the base vectors in T have the same order
as the indices on the component but occupy opposed positions.
We may solve the 27 equations (3) for the original components
Trst by multiplying by yayby, and summing; we thus find
j k (5)- r s t
yaybycTijk = SaSbScTrst = Tabc
In similar fashion we find, from (4),
(6) yaCbCkijk = brbbb -'r = Tab`.
156. Relative Tensors. When the coordinates are subjected to
an affine transformation,
(1) xi = Cr, ;xti
= yrxr (c = det c s 0),
the transformation equations, for all tensor components thus far
considered contain only the coefficients c, y. More generally, we
may have equations of transformation, such as
(2) T.jk = CNCr j
i t r ,
which involve the Nth power of the determinant c. In this case
the quantities in question are said to be components of a relative
tensor of weight N. When N = 0, equations (2) reduce to (155.4)
for an absolute tensor. As a consequence of (2), the relative tensor
T = Trseere8et becomes
(3) T = cNT
when the transformation (1) is effected.
The law of transformation of a tensor component is determined
by
(i) its valence: the number of its indices;
(ii) its type: the position of its indices from left to right;
(iii) its weight: the power of c that enters into the transforma-
tion equations.
346 TENSOR ANALYSIS § 157

If the weight of a tensor is not specified, it is assumed to be zero;


the tensor is then absolute.
A relative scalar so of weight N (valence zero) is transformed
according to
(4) gxl, x2, x3) = CNc,(x1) x2, x3).
The box product of the base vectors, E = el e2 X e3, is a relative
scalar of weight 1; for, from (147.2),
(5) E = cE.
The box product of the reciprocal base vectors, el . e2. e3 = E-1,
is a relative scalar of weight - 1; for E-1 = c-1E-1, from (5).
The determinant,
(6) g = det gij = E2 (152.8),
is a relative scalar of weight 2; for E2 = c2E2.
THEOREM. The permutation symbols Eijk and eijk, regarded as the
same set of numbers in all coordinate systems, are components of rela-
tive tensors of weight - 1 and 1, respectively.
Proof. If the formulas (146.2) are applied to the determinants,
c = det c2j y = det -?j- = 1/c ,
we have
CEijk =
iCjCkerat'

YEzlk = yr"Ys7 Brat.

Since Eijk = Eijk, Eijk = Eijk by definition, these equations assume


the form,
j kerat
e = C8Ctk
jarat,
z Eijk _ Y iryayt
required by relative tensors of weight -1 and 1.
157. General Transformations. Three equations,
(1) x' = f z(xls x2 x3) f
in which the functions f i are single-valued for all points of a region
R and which can be solved reciprocally to give the three equa-
tions,
(2) xt = gi(xl 22, 3)

in which the functions gi also are single valued, determine a one-


to-one correspondence between the sets of numbers xi and xi. If
we regard both sets of numbers as coordinates of the same point,
§ 157 GENERAL TRANSFORMATIONS 347

(1) defines a transformation of coordinates. The affine transforma-


tion is the particular case of (1) in which the functions f i are
linear and homogeneous in x', x2, x3.
Consider now the totality of such transformations in which the
functions f i(x', x2, x3) are analytic functions having a non-vanishing
Jacobian in R:
ax azi
(3) =det- 3,6 0.
OX ax'
The implicit-function theorem ensures the existence of the solu-
tions (2) of equations (1) in a sufficiently restricted neighborhood
of any point. In order to have a transformation of coordinates
in R, we must assume that the solution (2) exists and is single
valued throughout R.
The coordinate transformations (1) thus defined form a group;
for
(a) they include the identity transformation xi = xi whose
Jacobian is 1;
(b) each transformation (1) has an inverse (2) whose Jacobian
ax/ax I is the reciprocal of (3);
(c) the succession of two transformations of the set is equiva-
lent to another transformation of the set.
As to (c), the two transformations,
xL = fi(xl x2 x3) xz = h'(x' x2 x3)
are equivalent to
xz = h'[f`(x),f2(x),f3(x)] = ji(xl, x2, x3),
for which the Jacobian,
axi /0xi axr
detax'- = det\axr(-ax'-/
/
= (det
\ ax'/
ax
-
axi
axr
/
( det -)
\
Moreover, since f i and hi are analytic functions, the functions ji
L 0.

are likewise analytic.


In the affine transformation,
(4) xi = yrxt,
i
xi i-r ,
= crx
we have
(5)
axi

axr
yr -
axi

axr
=r z;
348 TENSOR ANALYSIS § 157

the Jacobians,
axi axi
(6) det - = y, det - = c, and yc = 1.
axr axr
A tensor of weight N, with respect to the affine group, trans-
forms according to the pattern,
CNyTy3CtkTrst.
(7) TU'.k =
In view of (5) and (6), this equation may be written
ax axi ax' axt
(8) Ti'k- at
IN axr axs axk
Now, by definition, a set of 33 functions Tr. $t is said to form a tensor
of valence 3 and weight N with respect to the general transforma-
tions (1) provided the components transform according to the
pattern (8). This equation becomes (7) when the transformation
is affine and constitutes a natural generalization of (7). The cor-
responding equation for a tensor of any valence or type is now
obvious. In particular, for contravariant and covariant compo-
nents of an absolute vector,
axi _ axr
(9), (10) uy =axr- ur, ii = ax- ur.
i
Since the new tensor components are linear and homogeneous
in terms of the old components, we have the important result: If
the components of a tensor vanish in one coordinate system, they
vanish in all coordinate systems.
More generally, tensor equations maintain their form in all co-
ordinate systems. If any geometrical or physical property is ex-
pressed by means of an equation between tensors, this equation
in an arbitrary coordinate system expresses the same property.
Although the coordinates themselves are not vector components
in the general group (1), their differentials dxi are the components
of a contravariant vector; for, from (1),

(11) dxi = - -
axi
ax'
dxI +
axi
ax2
axi
axi
axi
dx2 + - dx3 = - dxr.
axr
We regard the differentials dxr as the prototype of contravariant
vectors; and the rule of total differentiation gives the correct pat-
tern for their transformation.
§ 157 GENERAL TRANSFORMATIONS 349

The partial derivatives app/axi of an absolute scalar p(x', x2, x3)


are the components of a covariant vector; for, since
0(.' x2, x3) =
P(xl, x2,
x3),
when we replace xi on the right by the values (2),
a P ax2 a,p ax3 _ axr aV
(12) + ax2 ax i
ax ax 1
ax i ax3 at i ax ax r
We regard the derivatives ap/axr as the prototype of covariant
vectors; and the chain rule for partial differentiation gives the
correct pattern for their transformation.
In the affine transformation, the position vector r = xiei, and
(13) ei = ar/axi.
We now adopt (13) as the definition of ei in any coordinate system
x'; then
(14) ei=-=--=-er.
ar ar axr
axi axraxt
axr
axz

The base vectors ei thus transform after the pattern in (10) and
therefore merit a subscript.
The base vectors ei must transform after the pattern in (9),
namely,
a. t
(15) ey = - er;
axr
for, since er e8 = S8,
axr axj axr axj axj
ei
49V ax8 a.2'i ax' axi
and the sets ei, e' are also reciprocal.
We now can show that
(16) gij = gij = S

transform as tensor components; we need only make the replace-


ments indicated by (5) in the proofs of § 154. In fact gij, gi', S;
are all components of the idemfactor ere' = erer; thus S; some-
times is written gg. Moreover,
(17) ei = ei ere' = gire', ei = ei. erer = girer.
350 TENSOR ANALYSIS

158. Permutation Tensor. The box product E = el e2 x e3 is


a relative scalar of weight 1:
ax

ax
E
for the proof of (147.2) applies when we replace c; by axe/a.V.
Moreover, E-' = e' e2 x e3 is a relative scalar of weight -1.
The permutation tensor is defined by
(2) e1 - ej x ek eiejek = ei ej x ek eiejek.
These triadics are equal absolute tensors. For, if we put
ei = gireT) ej = 9ise3,
ek = gkter,

the left member becomes er es x e` ereset. Moreover, (157.14) and


(157.15) show that
(3) C-
= ei ej x ek eiejek.
From (2), we see that the covariant and contravariant compo-
nents of the permutation tensor are
(4) e1 - ej x ek = EijkE, ei ej x ek = EijkE-1

Since E and E-' are relative scalars of weight 1 and -1, Eijk and
Eijk are components of relative tensors of weight - 1 and 1. This
is the theorem of § 156; the former proof still holds when obvious
changes are made.
159. Operations with Tensors. The three basic operations on
tensors are addition, multiplication, and contraction.
1. Addition of tensor components of the same valence, weight, and
type generates a tensor component of precisely the same char-
acters.
Example. If Pij, Qij are both of weight N,
ax N axr ax"
Fij + Qij _ ax
-- (Pre + Qr.)-
a.ri azj

Hence Tit = Pij + Qij transforms in exactly the same manner as Pij and Qij.
2. Multiplication of tensor components of valence m1i m2, of
weight N1, N2, and of arbitrary type generates a tensor component
of valence m1 + m2i of weight N1 + N2, and of a type which is
defined by the position of the indices in the factors.
§ 159 OPERATIONS WITH TF''\SORS 351

Example. Let Pi' and Uk have the weights 2 and 1; then


ax ' ati a-'
at'.

ax axt Ox 3 az' axt


P'uk = I ai ly axr axe Pr8 az
axkut =
as ax'' ax8 D.
PT8 ut.

Hence T'jk = P''uk transforms as a tensor of valence 3, of weight 3, and of


the type indicated by its indices. Even though P''uk = ukPi', the product
tensors,
Pu = Pi'ukeie?ek, uP = ukPz'ekeie
are not the same.
3. Contraction. In any mixed tensor component an upper and
lower index may be set equal and summed over the index range
this generates a tensor component of the same weight and of va-
lence two less. Its type is determined by the remaining indices,
not involved in the summation.
Example 1. In the absolute component T'!.k set j = k; then
u=_VV jT1.11'+'V.22+T133
is a contravariant vector, for
Tt = a ax ax T3 = az ax
i i t i t

.. - ax"" ax8 all


T ..t -
Of ax8 Trt

all
=-38T'st = al'

T ur.
ax,* axr "` = ax"
A tensor component may also be contracted with respect to two
indices on the same level; we need only raise or lower one index
and then contract as previously.
Example 2. To contract the absolute component V% on the indices i, j
we first lower the index j
4 !r
T.jk = T..k9i*;
setting i = j now gives the covariant vector
!
Vk = T ::..kgir = T ..kgij.
We may also obtain this vector by lowering the index i and setting i = j:
Tick = Tt"k = V" k9ri
Contraction on a pair of indices is equivalent to the scalar multi-
plication of the corresponding base vectors in the complete tensor.
In the preceding examples, T = T!.ke,e ek,
u = Ti'.keiej ek = r'.ks;ei = T .,ei,
v = T:'.kei ejek = Tt'.kgi,ek
352 TENSOR ANALYSIS § 160

The proof that contraction produces tensors follows from the equa-
tion :
T'.'.keie;ek = TY?.keie,ek.

When the new base vectors on the left are expressed in terms of
the old, this becomes an identity; and, since scalar multiplication
is distributive with respect to addition, it remains an identity when
base vectors in the same position are multiplied on both sides.
Thus, in ex. 2,
V = T'.kgiiek = v.

As long as a contracted tensor has two or more indices, the fore-


going process may be repeated, each contraction reducing the va-
lence by two but leaving the weight unaltered. Thus the tensor
T!.kh may be contracted twice to yield two different scalars 7".'.i;
and T'!.;i, each consisting of nine terms. These are obtained from
the invariant tensor TT'.kheiejekeh by the formation of ei - ek,
e; eh and ei - eh, e; ek, respectively.
Contraction often is combined with multiplication. For ex-
ample, if we multiply the vectors u, v.; and then contract, we
obtain their scalar product u'vi. Again the product A B of two
dyadics defined in § 65 is equivalent to tensor multiplication fol-
lowed by contraction on the two inner indices. Thus, if A =
Ati,eV , B = Bkhekeh,
A - B = Ai;Bkhe'e' . ekeh = Aj;B'he'eh.
The product of the idemfactor I = ere? and a tensor T of valence
m is a tensor IT of valence m + 2. In general IT differs from
TI; but the contracted products reproduce T:

160. Symmetry and Antisymmetry. A tensor is said to be sym-


metric in two indices of the same type (both covariant or both
contravariant) if the value of any component is not changed by
permuting them. It is antisymmetric or alternating in two indices
of the same type if permuting them in any component merely
changes its sign. Thus, if
Tabc = Tbac, Tabc = -Tcba,
the tensor T is symmetric in a, b, alternating in a, c.
§ 161 KRONECKER DELTAS 353

Symmetry or antisymmetry are properties that subsist after a


general transformation of coordinates. Thus, for the preceding
example, we have
axa axb axe axb axa a.T.e
Tijk = Tabc = Tbac = Tjik
a
ax ax axk ax ax £ axk
axe axb axa
(-Tcba)
axk axj axi
A tensor cannot be symmetric or alternating in two indices of
different types; for a property such as Ti j = Vi i does not subsist
after a transformation of coordinates.
A tensor is said to be symmetric in any set of upper or lower
indices if its components are not altered in value by any permuta-
tion of the set. The subsistence of this property in one coordinate
system ensures it in all.
A tensor is said to be alternating (or antisymmetric) in any set
of upper or lower indices if its components are not altered in value
by any even permutation of the indices and are merely changed in
sign by an odd permutation of these indices.
A tensor Tij or Tijk (T ij, Tijk) which is alternating in all indices
is called a bivector or trivector, respectively. In such alternating
tensors, all components having two equal indices are zero. In a
trivector Tijk, the non-zero components can have but two values,
±T123; moreover, the contracted product,
EijkTijk = 3!T123

In this connection we remind the reader that a given permuta-


tion of n indices from some standard order can be accomplished
by a succession of transpositions of two adjacent indices and that
the total number of transpositions required to bring about a defi-
nite permutation is always even or always odd. The permutation
in question is said to be even or odd in the respective cases.
161. Kronecker Deltas. We now generalize the simple Kro-
necker delta S by introducing two others, defined as follows:
(1)
aiajbkc ijk
- Eabr E
ij ijr = aabr
ijr
(2) aab = EabrE
Since the epsilons have weights of 1 and -1, ba"bk, is an absolute
tensor of valence six; hence b b, formed by contracting S, is an
absolute tensor of valence four.
354 TENSOR ANALYSIS § 162

Those definitions show that S b and bay can only assume the
values 0, 1, -1; they are evidently alternating in both upper and
lower indices. Their precise values in any case are as follows:
If both upper and lower indices of a generalized delta consist of
the same distinct numbers chosen from 1, 2, 3, the delta is 1 or -1
according as the upper indices form an even or odd permutation
of the lower; in all other cases the delta is zero.
This rule is a direct consequence of the properties of the epsilons.
We have, for example,
012 32 23 13 o.
12 = 1, 023 - - 1, 011 = 021 ;

123 231 6213 v321


x123 = 0123 - 1, = 0123 8123 = 0

162. Vector Algebra in Index Notation. The three operations


on tensors enable us to give a succinct account of vector algebra.
Vectors are to be regarded as absolute unless stated to be relative;
and we shall often speak of components as vectors.
If w = u + v, we have
vi =ui+vi, or wi=ui+vi;
obviously vector addition is commutative and associative.
The tensor product of two vectors u, v is the dyad uv. On con-
traction, uv yields the scalar product,
(1) u . v = uivi = uivi = gijuivj = giju1vj.
From (1), we have
U. (v+w) =
The antisymmetric dyadic (bivector) P = uv - vu is called the
outer product of u and v; its covariant components are
(2) Pjk = UjVk - UkVj = 8jkuavb.

The dual of Pjk (cf. § 170) is the contravariant vector of weight 1:


e ijk Pik = E''ikujvk;
(3) p = 2
Z
its components,
U2V3 - u02, U3V1 - u1V3, u1V2 - u2v1Y

are the same as the non-zero components Pik. Since E-1 is a


scalar of weight -1, the components E-'p' are absolute. The
corresponding vector,
§ 162 VECTOR ALGEBRA IN INDEX NOTATION 355

(4) E-lpiei = E-1EijkeiujVk = E-1 = uXv (24.9).


v1 V2 V3

Similarly, from the outer product,


(2)' P'k = bab26avb = u'vk - ukvJ,
we obtain as dual the covariant vector of weight - 1:
(3)' pi =_ZEijkPJk
i = eijklljvk.
The absolute components Epi again give u X v :
el e2 e3
(4)' Epiei = Eeijkeiujvk = E ul u2 u3 = uxv (24.10).
vl v2 v3
From (4) or (4)', we have
uXv = - V X u, U- (v + w) . u x v + U X W.
The components of the triple product u x (v X w) are
(u X (v X w))' = E-1EijkUj(V X w)k (4)
= E-hE'fkujEEkabvawb (4)'

= EijkEabkujvawb

= SabUjva'Uib (161.2)
= uj(viwj - vjwi)
= (u w)vi - (u - V)wi;
hence
u X (v X w) = (u w)v - (u v)w.
The box product u x v w is given by either of the absolute
scalars,
ul u2 U3

(5) (U X V)'wi = E-'Cijk,ujvkwi = E-1 Vi V2 V3

W1 W2 W3
U1 u2 u3

(5)' (u X v)z-w = Eeijku'vkw' =E V1 V


2
V3

w1 w2 w3
in agreement with (24.12).
356 TENSOR ANALYSIS § m3

163. The Afne Connection. Any vector can be expressed as a


linear combination of the base vectors. When the 32 derivatives
of the base vectors ae;/axi = Die; are thus expressed,
(1) Die; = ref + r e2 + r 'e3 = ri,er,
the 33 coefficients r, are called components of the affine connection.
If we multiply (1) by ek , the right-hand member becomes rzJj r
= r ; hence
(2) r = ek Die;.
The law of transformation for r is given by
axk ec \ / -- DQ I I axb eb ),
(3) r = ek Die; =
ax c / l ax i / \ax '

where
(4) Di =axia - _axa
- -a
axi axa
_ -axb
Da;
axi
The differential operator Di transforms like a covariant vector (hence
the subscript). In (3), Da acts on both scalar and vector factors
following; hence, from (1),
axk
e ( a2xb eb+--raber
axa axb

ax` axiax' axi axi


Since e° eb = bb, ec er = 6'r, this gives
a2xb axk axa axb axk
(5)
axiaxj axb
+ - - - rab
axi ax' ax`
for the desired law of transformation. Therefore r is a tensor
component when and only when
a2xb axb
= 0, - = CJ (const),
49-Pi axi ax'
and hence xb = c;x', if the coordinates xi and xi have a common
origin.
The components r of the affine connection (between the deriva-
tives of the base vectors and the vectors themselves) are tensor
components only with respect to affine transformations.
§ 163 THE AFFINE CONNECTION 357

Although the Irk are not tensor components for general trans-
formations, we shall see that the index notation still serves a use-
ful purpose. j
Since e1 = D;r (157.13),

(6) P = ek DiD;r = ek D;Dir = Pk;


and, from (5), we see that the symmetry of P in the subscripts
persists after a transformation of coordinates.
If we transform coordinates from x to IF, we have, from (3),

(7)
P4 - \axk k/ \ax1' \4924

or, on making the replacements,


axk (9x, axb
e = - es,
k
Di = - Da, e; eb,
ax, axi ax'
_r _ axr axa (axb

(8) l l
PPQ
- ax, ec/ \a2p Da/ \ai7 eb/
a relation of the same form as (3). Consequently, the succession
of transformations r --p I` -> f produces a transformation r -- r
of the same form as r -+ 1. We express this property by saying
that the transformation (3), or its equivalent (5), is transitive.
If, in particular, x i = x', we may delete all the tildas (-) in (7) ;
this equation then gives, on expansion,
a2xJ air a. a. axr k
(9) P,, I'2J.
axP axq ax' axP a:rq axk

These equations constitute the transformation inverse to (5)


and also may be derived by solving these equations for P'ab.
This solution may be effected by the multiplication of (5) by
axi ax' axr
axp ax4 axk

t The coordinates xi present a similar situation. For afflne transformations


they are components of a vector; for general transformations this is not the
case, but the indices still serve to indicate that their differentials dxi are vector
components.
358 T1E;NSOR ANALYSIS §163

On writing y = x, the equation of transformation (5) becomes:


a2-' ax, ayk
(10)
i
+ - - rab axb 1

lay, ayj ayi ay' axc

If we now make the change of coordinates,

(11) xr = xo + yT - 2(rPq)oypy4,

where the gammas are computed for x = xo, the point x" = xo
corresponds to yr = 0. Since ay'r/ayi =
ax, a2xr
ayi = Si - ayi a y.i - - (r)o;
hence, at the point x' = xo (yr = 0), the brace in (10) becomes
-(r )0 +aj(rab)o = 0.
Consequently, all the gammas I'- = 0 vanish at the origin y'' = 0
of the new coordinates. Such a system of coordinates is termed
geodesic. Since the gammas are not tensors, the equations riI; = 0
(y = 0) do not imply r = 0 (x = x0).
Example 1. For cylindrical coordinates p, p, z, we have (§ 89, ex. 1)
r = [x, y, z] = [P cos ip, P sin ,p, z].
If we put x1 = p, x2 ='P,x3 =z,
el = Dlr = [cos p, sin p, 0],
e2 = D2r = p[- sin p, cos {p, 0],
e3=D3r=[0,0,1];
Die, = 0, D2ei = [- sin gyp, cos (a, 0] = 1 e2, D3ei = 0,
P

D2e2 = -p[cos rp, sin p, 0] = -pei, D3e2 = 0,


D3e3 = 0.

Hence all gammas are zero except

rig = r2, = 1/P, r22 = -P-


Example 2. For spherical coordinates r, sp, B, we have (§ 89, ex. 2)
r = [x, y, z] = r[sin 0 cos p, sin 0 sin p, cos 0].
§ 164 KINEMATICS OF A PARTICLE 359

If we putxi =r,x2= V,x3=0,*


ei = Djr = [sin 0 cos gyp, sin 0 sin gyp, cos 01,
e2 = D2r = r sin 01 - sin gyp, cos p, 0],
e3 = D3r = r[cos 0 cos gyp, cos 6 sin gyp, - sin 0];
1 1
D1e1 = 0, D2e1 = 1 e2, D3e1 = I e3,
r r
D2e2 = -r sin2 6 e1 - sin 0 cos 0 e3, D2e3 = cot 0 e2,
D3e3 = -rel.
The non-zero gammas are therefore
r12=r21=1/r, r13=r31= 1/r, r22= -r sin2 B,
r22 = - sin 0 cos 0, r23 =
1132
= Cot 0, r33 = -r.
164. Kinematics of a Particle. We now may find velocity and
acceleration of a particle in general coordinates:
(1) a = dv/dt = akek.
V = vkek,
In rectangular coordinates x1 = x, x2 = y, x3 = z we have vk =
dxk/dt (52.7). Hence, in general coordinates 2i, these components
become
8xk dxi dxk
(2) vk = - - _ -
axi dt dt
The time derivatives of the coordinates, which we write ±k, thus
transform as contravariant vector components.
For the acceleration we have, from (1),
dvk aek dxi
a= ek + vk vtvkr:kei;
dt ax i dt = vkek +
and, on interchanging the summation indices j, k in the last term,
(3) a = (vk + viv'r ,)ek
Therefore, in any coordinate system the velocity and acceleration
components are
(4) vk = xk ak = xk + x°'x'I').
Since the base vectors ei are not, in general, unit vectors, we
must distinguish between the components vk and ak and the nu-
merical values of the terms in vkek, akek. Thus the terms v'e1,
a'e1 have the numerical values vllell, a'lell.
* The order r, p, 0 is sinistral and [ele2e3] < 0; cf. p. 197.
360 TENSOR ANALYSIS § 1&

Example 1. Cylindrical Coordinates p, ,p, z. From § 163, ex. 1, el, e2, e3


have the lengths 1, p, 1; hence the values of the velocity terms in (1) are
a, pp, z
From the non-zero gammas r22 = -p, rig = 1/p, we have
a1 = a - Psp2, a2 = sp + 2Pcp/P, a3 = z;
and the numerical values of the acceleration terms are
P - Pcp2, p' + 2p', z.
Example 2. Spherical Coordinates r, p, 0. From § 163, ex. 2, e1, e2, e3 have
the lengths 1, r sin 0, r; hence the velocity terms in (1) have the values,
r, rcc sin B, r9.
With the non-zero gammas r22 = -r sin 2 0, r33 = -r we have, from (4),
al = r - p2r22 + O2r33 = r - rrp2 sin2 B - r92;
with r12 = 1/r, r23 = cot 6,
a2 = + 2r,pr12 + 2,pOr23 = + 2r(p/r + 24 cot 0;
and, with r13 = 1/r, r22 = - sin 6 cos 0,
a3 = 9 + 2rOri3 + ,,2r22 = 9 + 2r9/r - ,p2 sin B cos 0.
Hence the values of the acceleration terms are
1' - r,p2 sin2 0 - r92, r,p sin 0 + 2i ,p sin 0 + 2r09 cos 0,
r9 + 2r9 - r02 sin B cos 0.
165. Derivatives of e i and E. From the relation e' er = S*,
we have, on differentiation,
(Die') er = - e' Deer = - rsr,
and hence
(1) Die' rrer
From Die; = r jer and (1), we have
(2) Di(e,e') = rz,ere' - rreer = 0,
on interchanging the summation indices r, j in the second term.
Since the product E = el e2 x e3 is distributive with respect to
addition, its partial derivatives may be found by the familiar rule
for differentiating a product,
DiE = (Diet) . e2 x e3 + el (Die2) x e3 + el e2 x (D:e3)
= r,r'1er e 2 x e3 + ri2e1 er x e3 + ri3e1 e2- er
1 2 3
_ (ri1 + ri2 + ri3)e1 . e2 x e3,
§ 166 AFFINE CONNECTION AND METRIC TENSOR 361

or, if we apply the summation convention,


(3) DiE = rirE.
The derivative of E-i = el e2 e3 is therefore
(4) DiE-i = -E-2DiE rirE-i
From (3), we have
DjDiE = (Djrir + rarr .,)E;
and, since DjDiE = DiDjE,
(5) Dj rir = Dir;r
166. Relation between Affine Connection and Metric Tensor.
On differentiating gij = ei ej, we have
(1) Dk9ij = rkier - ej + rkjei - er = rkjJrj + rkjgir
Since Dkgij and F are both symmetric in the indices ij, there are
3 X 6 = 18 quantities in each set. Equations (1), 18 in number
and linear in the 18 gammas, may be solved for the latter.
We first introduce the notation,
(2) F9rk = rij,k,
just as if r k were a tensor whose upper index was lowered. Then
we have also
(3) rij,rgrk = r ,
for the left member equals
ri,gar9rk = I't S = r
We note that Fij, k is also symmetric in the indices i, j. Moreover
(4) Diej = F jer = rij,se9,
if we put e,. = g,.ses.
We may now write (1) in the form:
(5) rki,j + rkj,i = Dkgij.
If we permute i, j, k cyclically in (5), we obtain the equations,
rij,k + rik,j = Di9jk,
rjk,i + rji,k = Dj9ki;
362 TENSOR ANALYSIS § 167

and, upon subtracting (5) from their sum, we obtain


(6) ri.i.k = (Digjk + D;gki - Dkgi1)
Z
We may now compute r from (3).
In the older literature, the components of the affine connection
are denoted by
[ii,k] = rij,k, [ } = r., .V Z)

These Christoffel symbols of the first and second kind, respectively,


therefore are defined by the equations:
J l
(7) [ij, k] = (Digik. + D7gki - Dkgii), [ } = gkr[ij, r]. 7

z
167. Covariant Derivative. The gradient of an absolute tensor
T is defined as
aT
(1) vT=eh
axh
Since T = T,

eh
aT
=axh_ ar__
aT ax8 aT
8 r--
= Sre =e 8
aT
axr ;
axh axr ax8 axh
hence VT is a tensor of valence one greater than T.
If the components of T are T ;.'.'.k (the order of the indices is
not specified), the components of VT are written
Tab
(2) Vh i;...k,
the index h on V corresponding to the differentiation a/axh. This
is a covariant index, for the operator Dk = a/axh transforms like
a covariant vector:
(3) Dh=----=-Dr.
axr a
axh axr
aaxr
axh axh
For this reason the components (2) are called covariant derivatives
of Tab
If T is a relative tensor of weight N and valence m, E-NT is an
absolute tensor, whose gradient,
ehDh(E-NT),
is an absolute tensor of valence m + 1. If this is multiplied by
EN, we again obtain a relative tensor of weight N; this tensor is
defined to be the gradient of T and written
(4) VT = ENehDh(E-NT).
§ 167 COVARIANT DERIVATIVE 363

When N = 0, (4) reduces to (1). The components of VT, denoted


by prefixing Vh to the components of T as in (2), are called co-
variant derivatives.
From (165.3), we have
= -NE-'V-1DhE = -NrhrE-N,
DhE_N

and hence
Dh(E-'YT) = E-v(DhT - NrhrT),

(5) VT = eh(Dh - Nrhr)T.


Thus the operator,
V = eh(Dh - Nrhr), r

applied to any invariant tensor T (with its complement of base


vectors) generates another tensor VT of the same weight and va-
lence one greater.
We next compute the components of VT explicitly, where
(6) T = Tq '...ceaeb ... e,eiej ... ek.
If T is a relative tensor of weight N, VT contains the term,
-N rrhr
'
k eheaab
Tab'

ij... e,e iej e k.


It remains to compute the part of VT due to the operator ehDh.
Now Dh acts on the "product" of T ;.'.'.k and a series of base vec-
tors. Since this product is distributive with respect to addition,
DhT may be computed by the usual rule for a product; hence
DhT = (DhTi;...k)eaeb
ab c
. . . eye iaj . . . ek

+ eceiej ... ek + .. .
ab i j
ec(Dhe )e ... ek +
In the second line put
Dhea = rhaer, ... , Dhe, =
and in the successive terms interchange the summation indices
r and a, r and b, , r and c. In the third line put
Dhei - -rhrei r , ... ,
Dhe k- k r
- - rhre ,
and in the successive terms interchange the summation indices
r and i, r and j, , r and k. When this is done, each term of
VT contains the same complex of base vectors,
eheaeb ... eceiej ... ek,
364 TENSOR ANALYSIS § 167

and the component VhTt6.'.k of VT equals the sum of their scalar


coefficients :
Tab . C
17 v Tab e = D
k- q- k
a
Tij...krhr + + Tij.k rchr
Trj...k rhi -
r
- Tij...r rhk r

- Nrr rTab c
hr

This is the general formula for the covariant derivative of any


tensor component, absolute or relative. The last term is absent
when T is absolute (N = 0). For every upper index
hik
v Ta*: contains a term T°::.r::
t k r*
hr y

and, for every lower index *,


OhTa' *:: °k contains a term - T" ' a rr.
We consider now some important special cases.
If p is a relative scalar of weight N,
(8) vhcp = Dh-p - Nrhr(c
Since E and E-1 are scalars of weight 1 and -1, we have, from
(165.3) and (165.4),
(9) vhE = DhE - rhrE = 0,
(10) vhE-1 = DhE-' + rhrE-1 = 0.
Again, since g = det gij = E2 is a relative scalar of weight 2,
(11) Vhg = DhE2 - 2r;,rE2 = 0.
If v = vie1 = vie' is an absolute vector,
(12) VhV = Dhvti + vrrhr,
r
(13) vhvi = Dhvi - vrrhi.
These expressions are the mixed and covariant components of one
and the same dyadic Vv.
For an absolute dyadic T, the components of VT may have the
forms :
(14) VhT2 = DhT'' + Tr1rhr + Tirrhr,
(15) VhT`j = DhVj + T'jrhr - Tirrhj,
(16) VhTij = DhTij - Trjrhi - T,rrhj.
§ 168 RULES OF COVARIANT DIFFERENTIATION 365

For the metric tensor,


(17) G = gijeiej = gijeiej = eiei = I,
we have, from (165.2),
(18) VG = e'DhG = ehDh(eiei) = 0.
The components of VG therefore vanish:
(19) Chgij = 0, Vhgii = 0, Vhbi = 0.
Note that (166.1) is equivalent to Vkgij = 0.
Example. We can write
i20) VT = eh(Dh - Nrhr)T = ehvhT,
if we regard Vh as an operator that acts only on scalars. Covariant differentia-
tion then is defined by this operational equation. With this convention, we
have also
VVT = ei(Di - Nrir)ej(Dj - Nrn)T = eiejVjVjT.
The second member may be written
eiej(Di - Nrit) (Dj - Nrjr)T - eirise'(Dj - Nrr )T
= eiej ( (Di - Nrit) (Dj - Nrir) - rij(D. Nr )1T
when indices j and s are interchanged; hence
viv jT = (Di - N rir) (Dj - Nrrr,)T - rijV,T,
ViViT = (D, - Nrjr) (Di - Nrsr)T - r-! v.T,
and, on subtraction,
(21) (CiV1 - Vjvi)T = (D1Dj - DjDi)T,
in view of (165.5). On the left the operators Vr act only upon the scalar com-
ponents of T. We note that (21) applies to relative as well as absolute tensors.
168. Rules of Covariant Differentiation.
1. The covariant derivative of the sum or product of two tensors may
be computed by the rules for ordinary differentiation.
If we introduce a system of geodesic coordinates yi 1163), the
corresponding gammas 1 , will all vanish at the point yi = 0;
hence, from (167.7),
OTab...r=DTab...r
(1) h ij...k h ij...k,

at the origin of geodesic coordinates, which moreover can be


chosen at pleasure. For example, we have the tensor equation,
Vh(Tt'uk) = (VhTz3)uk + T''Vhuk,
366 TENSOR ANALYSIS § 169

in geodesic coordinates and therefore in any coordinates (§ 157).


Consequently, the sum and product rules of ordinary differentia-
tion also apply in covariant differentiation.
2. The covariant derivatives of the epsilons and Kronecker deltas
are zero.
Since these tensors have constant components, their covariant de-
rivatives vanish at the origin of a system of geodesic coordinates;
hence they vanish in all coordinate systems.
3. The operation of contraction is commutative with covariant dif-
ferentiation. For example, if we contract T=jk on the indices i, j
to form
T?ik = SiT -jk,
we have
OhVik = BiZVO -jk.

4. The components of the metric tensor (gij, gij, S) may be treated


as constants in covariant differentiation.
This follows at once from (167.19). For example,
Civj = Vi(g7rvr) = gjiViv'.
Thus we may find Vivj by lowering the index j in Divj; in other
words, Divj and Vivj are components of one and the same tensor:
Vv = V vieiej = V1vjeiei.
Evidently the operation of raising or lowering an index is commu-
tative with covariant differentiation.
5. The relative scalars E, E-1 and g may be treated as constants
in covariant differentiation.
Since E, E-1 and g = E2 are relative scalars of weight 1, -1, 2,
respectively, we have, from (167.4),
VE = 0, V -'=0, Vg = 0;
for in each case Dh(E-NT) = DO = 0-
169. Riemannian Geometry. Any set of objects which can be
placed in one-to-one correspondence with the totality of ordered
.
sets of real numbers (.r.1, x2, , x") satisfying certain inequalities,

I xi - ai I < ki (a1 and ki > 0 are constants),


is said to form a region of space of n dimensions.$ We speak of
(x1) x2, . , x") as a point; but the actual objects may be very
$ Veblen, Invariants of Quadratic Differential Forms, Cambridge, 1933, p.
13. In some applications the numbers xi may be complex.
§ 169 RIEMANNIAN GEOMETRY 367

diverse. Thus an event in the space-time of relativity may be


pictured as a point in four-space; and the configurations of a dy-
namical system, determined by n generalized coordinates, often are
regarded as points in n-space.
If we associate the space (x', x2, , x") with an arbitrary
non-singular quadratic form,
(1) 9ii dxi dx',
(9ij = 9ii, g = det gij 0),
we have a Riemannian space with a definite system of measure-
ment prescribed by this form (§ 151) ; and the geometry of this
metric space is called Riemannian geometry. We assume that the
coefficients gij are continuous twice-differentiable functions of x.
The base vectors ei are not specified; but their lengths and the
angles between them may be found from the relations:
(2) 9ij = ei e,.
The reciprocal base vectors now are given by
(3) ei =9irer,
where gii is the reduced cofactor of gij in det gij; for
(4) ei e; = girgir = d (152.2).
Moreover
ei e' = girS* = gij
(5)
If we now transform coordinates from xi to xi, we assume that
the new base vectors are given by
axr axi
(6) ei = - er, ei = - es;
axb

ax8
then ei and e' still form reciprocal sets, for
axr a'V
8r = o.
a:C i ax8

In view of (2) and (5), equations (6) show that gjj and gi'
transform like absolute dyadics. These tensors often are called
the fundamental covariant and contravariant tensors of Riemann-
ian geometry. By use of the equations,
(7) e = g,rer, ei = girer,

they permit us to raise and lower indices (§ 154) and thus repre-
sent any tensor of valence m by 2' types of components.
368 TENSOR ANALYSIS § lti'9

At a given point, gij gives the orientation of the base vectors ei


relative to each other. In order to determine the relative orienta-
tion of sets of base vectors at different points, we must know the
components r of the affine connection, defined by
(8) Diej = rer.
Then, just as in § 165, we have also
(9) Die' Tier (165.1).
We now assume that the affine connection is symmetric (r = r ).
The gammas then are determined by the metric tensor (§ 166) :
(10) rkV = 129kr(Digjr + Djgri - Drgij) I
The epsilons in n-space, defined in § 146, have n subscripts or
superscripts. The equations,
(11) -
ax ..
ax
xi axj
(9.t axk Eb...
axa axb
. . .ax`- '
ax axa axb ax°
IEij ...k = ...axk Eab...ej,
(12) ax ax` 49V
t We shall call the base vectors constant if Die, = 0 (i, j = 1, 2,. , n);
then raf = 0 and the functions gij = ei e; are also constant. Conversely,
when gii are constants, (10) shows that r!J = 0, and hence Die, = 0.
In Riemannian space it is not, in general, possible to introduce coordinates
xi for which the base vectors ei are constant. Only flat space (§ 175) is com-
patible with such Cartesian coordinates; then each point has the position vector
r = xiei and ei = ar/axi. Moreover, for any coordinates 2i in flat space, we
have (cf. § 163, ex. 1, 2)
ax' or ax'ar
(i) ei = = = .
021 e' ax' &V at i
The geometry on a surface with the metric tensor gi; is Riemannian (n = 2).
Unless the surface is flat (a plane, for example), constant base vectors cannot
be introduced. If we regard the surface as immersed in Euclidean 3-space
each surface point has the position vector r = xi + yj + zk, where
x = x(u, v), y = y(u, v), z = z(u, v)
are the Cartesian equations of the surface. If we write xl = u, x2 = v, the
base vectors on the surface may be taken as ei = ar/axi; for equation (i)
shows that these vectors transform in the manner prescribed in (6). Here
r is a position vector in Euclidean 3-space; but, in general, the surface points
have no position vector in the Riemannian 2-space they define.
Any Riemannian space of n dimensions may be regarded as immersed in
a Euclidean space (§ 178) of n(n + 1)/2 dimensions. This theorem has not
as yet been rigorously proved; see Veblen, op. cit., p. 69.
§ 169 RIEMANNIAN GEOMETRY 369

generalized from (146.2), show that eij"'k and eij...k are relative
tensors of weight 1 and -1; for these are the powers of the Jacobian
I ax/ax I when it is transferred to the right-hand side.
There are n types of Kronecker deltas in n-space: S,, S b, S bc,
up to Sa'bc'...f with 2n indices. As in § 161, they are defined in
terms of the epsilons. In the case n = 4, for example,
i 1 ibcd 1 ijcd
Sa = - eabcde Sab - - eabcde ,

3! 2!

ijk 1 ijkd ijkh ijkh


abc eabcde , aabcd = eabcde

The rule given in § 161 for the value (0, 1, or -1) of a generalized
delta still applies. Moreover the preceding definitions show that
all the deltas are absolute tensors, alternating in both upper and
lower indices. See Prob. 42.
The n-rowed determinant,
(13) g = 1t etii...ke,s...tgir9js ... gkt [cf. (146.3)],
n!
generalized from (152.1), is the contracted product of two epsilons
of weight 1 and n absolute dyadics. Hence the discriminant g of
the fundamental quadratic form is a relative scalar of weight 2.
The cofactor of gij in g is ggij; hence, from (146.11) and (166.5),
we have
Dhg = gg"Dhgij = ggi (rhi, j + rhj,
or, in view of (166.3),
(14) Dhg = grhi + gr,. = 2grhr-
From (14), we have also
(15) DhV_ -\'g-r,.,,,
a result of the same form as (165.3) with E replaced by -/g.
In defining the gradient of a tensor, we replace E by -\/g in
(167.4); thus, in Riemannian n-space,
N N
(16) VT = g2ehDh(g 2T).
The components of VT again are given by (167.7). Hence this
formula for the covariant derivative is still valid in Riemannian
geometry.
370 TENSOR ANALYSIS § 170

Since the metric tensor G = gjkejek = ekek still has the prop-
erty DhG = 0, VG and its components vanish as before:
(17) Ohgij = 0, Vhb = 0.
Ohg J = 0,
Moreover, from (16), vq = gehDh1 = 0, and
(18) Ohg = 0.

170. Dual of a Tensor. An m-vector is a tensor of valence m


which is alternating in all indices (cf. § 160). For convenience in
wording, we also regard scalars (m = 0) and vectors (m = 1)
as m-vectors. In n-space we can associate with any m-vector
(0 < m < n) an (n - m)-vector, its dual, defined as follows:
If and Qij...k are m-vectors of weight N, their duals are
the (n - m)-vectors,
(1)
1 ab...c ij...k
Pj...k,
m.

(2) gab...c - li Qij...kEij...k


ab...c,
m.
of weights N + 1 and N - 1, respectively. Note that the epsilons
have n indices in n-space; and, in the contracted products, the
contravariant tensor is written first, and the summation indices are
adjacent.
A scalar Sp has two duals, Eab .htp, Eab . hV, according as we use
(1) or (2); they have the same numerical components but different
weights. The dual of an n-vector Tijk...h is the scalar:
1 ijk...h
E Tijk...h = T123...n.
n!
THEOREM. m < n),
If T is an m-vector (0
(3) dual dual T = T.
Proof. Write T = P, dual T = p. Using (2) and (1), we have
1
(dual pab c
rs.
(n -- m) !
1
= Eab...c
ra...t
(n - m) !m!
1 i;...k
_ -m!brs tPi;...k [Cf. Prob. 42.]

= Pra...t)
§ 171 DIVERGENCE 371

since Pip -.k is alternating in all indices. Hence dual p = P; and,


similarly, dual q = Q.
When T is an n-vector, say Tijk...h, (1) gives the dual T123 . . n
now (2) gives
dual dual T = Eijk...hT123...n =
Thus the theorem holds in this case also, provided both dualizing
equations (1), (2) are used.
171. Divergence. The gradient of v = viei is
Vv = vhviehei.
The first scalar of this dyadic is
V11)
(1) div v = %ib =

If vi is an absolute vector, the divergence is the absolute scalar


Viv'; this definition applies in n-space and for any coordinate
system.
When vi is a relative vector of weight 1,
VhV' = Dhvi + 17rrhr - vtrhr (167.7).
On contracting with h = i, the second and third terms cancel; for
vrrsr = virir, since r and i are summation indices. Hence
(2) div v = Divi (wt. v = 1).
If vi is absolute, vi has the weight 1 imparted by the scalar
/; hence
viva = vi(9-i91vi) = 9-lvd(91v=) = 9-1Di(91vi),
in view of (2). Thus
(3) div v = Di( / vi) (wt. v = 0).
9
The Laplacian V2p of the scalar p is defined as div Vp. If rp is
absolute, v = V p is an absolute vector; then
Vr = Drop, Vi = 9irDrcp,
and, from (3),
(4) v2(p Di(V -J 9irDrcp)

The divergence of any tensor T is defined as the gradient VT


contracted on the first and last indices. Thus if T has the com-
372 TENSOR ANALYSIS 1 172

porients T 'j- kh of valence m and weight N,


(divT)ij...k VhTij...kh
(5) =
is a tensor of valence m - 1 and weight N.
THEOREM. When T i' ., kh is an m-vector of weight 1, div T is the
(m - 1)-vector,
(6) (div T) ij...k = DhTij...kh,
where V in the defining equation is replaced by D. Moreover,
(7) div div T = 0 (m > 1).
Proof. From (167.7),
VhTij...kh = DhTij...kh + 1,hrTrj...kh +
... + rk Tij...rh

ij" . kr - rhrT ij... kh.


+ rhrT
The two final terms cancel, as we see on" interchanging the sum-
mation indices h, r in the last term. The remaining terms con-
taining 1'hr vanish separately on summing over h and r; for rh,. is
symmetric and T antisymmetric in these indices. Thus (6) is
proved.
From (6) and the alternating character of T,
(div div T) 'j... = DkDhT'j... kh = 0.

172. Stokes Tensor. The gradient of the covariant vector vk is


the dyadic Vhvk. From this we form the antisymmetric dyadic,
(1) Pij = Sf V hvk,

known as the Stokes tensor. t When Vk is albsolute,


rrhkvr,
Ohvk = hVk - .

and, since az4r = 0 owing to the symmetry of rr , we have


(2) Pij = SfDhvk = Divj - Djvi (wt. v = 0).
In 2-space we can form from Pij the relative scalar of weight 1,
(3) 2eijPij = P12 = Dlv2 - D2v1 (wt. V = 0),
t Veblen, op. cit., p. 64.
§ 172 STOKES TENSOR 373

and from this the absolute scalar,


1
(4) (Dlv2 - D2v1) (wt. V = 0).
-Vg
This is the absolute invariant on the surface with the fundamental
form gig dxi dx', written n rot v or n V X v $ in § 97.
In 3-space we can form from Pii the relative vector of weight 1,
(5) -2
wi = 1EijkP7k = Ei'kD7'v k,
having the components,
D2v3 - D3v2i D3v1 - D1v3, D1v2 - D2v2.
The absolute vector,

(6) rot v = 1 wiei = 1 Ei'keiDjvk,


1/9
may be written as a symbolic determinant (cf. § 146):
el e2 e3
1
(7) rot v = D1 D2 D3
V9 V1 V2 V3

Comparing this with (88.19) now shows that rot v is in fact the
rotation of v previously defined.*
In § 91 we proved that a vector v is the gradient of a scalar in
3-space when and only when rot v = 0. In n-space we have the
corresponding
THEOREM. Let v be a continuously differentiable vector in a region
R. Then, in order that v be a gradient vector,
(8) V = OAP, vi = Dip,
it is necessary and sufficient that the Stokes tensor vanish in R:
(9) Pi,=Diva-D;vi=0.
Proof. The condition is necessary; for, if vi = Dip, Pi; vanishes
identically, owing to the continuity of the second derivatives of cp.
In (97.9), H=-v19, u=x1, v=x2, ru=el, rv=e2, r,.f
= f2 in our present notation.
* In (88.19), J u = x1, ru = el, ru f = fl, etc., in our present
notation.
374 TENSOR ANALYSIS § 173

The condition is sufficient. For the method of § 91, extended to


case of n coordinates, leads to the function,
(10) (P = fal
z1

vj (xl,
I. x2
x2)
... , xn) dxl

+ v2 (a', x2, ... xn) dx2


a2

" f"
f zn
3

v3 (ale a21 x3,

v (ai a2 ..
xn) dx3 ...

an-1 xn) dxn


n
where all a2, , an are the coordinates of a fixed but arbitrary
point. In the ith integral xi is the variable of integration while
xi+1' ... xn are regarded as constant parameters. We now can
,
show from (9) and (10) that Dig = vi. Let us compute, for ex-
ample, D3V. Only the first three integrals in (10) contain x3; their
derivatives with respect to x3 are, respectively,
v3(xl, x2) ... , xn) _ v3(al) x2, ... , x"),
v3(a1 , x2, ... , xn) - v3(al, a2, ... 'X n), v3(al, a2, ... , xn),
when we make use of D3vi = Dlv3, D3v2 = D2v3 in the first and
second; hence D34P = v3(xl, x2, , xn).
173. Curl. We define the curl of a covariant tensor Tb...d of
valence m (m < n) as the (m + 1)-vector:
1
(1) (curl T)hij...k = bhij .k VaTbc. d
m.-
When m = 0, T = P (a scalar), 0! = 1, and
curl cp = ahV aco = V MP
is the gradient of gyp.
When m = 1, T = v (a vector), curl v is the Stokes tensor
(172.1).
In general, we have the
THEOREM. When Tab.. d is an absolute tensor of valence m < n,
Shaijc
b ......kD..Tbc...d,
(2) (curl T)hij ... k =
M.
§ 173 CURL 375

where V in the defining equation is replaced by D. Moreover,


(3) curl curl T = 0 (m < n - 1).
Proof. From (167.7), we have
VaTbc...d = DaTbc...d - rabTrc...d 1'adTbc...r.

Hence in the right member of (1) there are m terms of the type,
1 Sabc...dl,r 7,
ab rc...d,
m!

these all vanish separately when we sum over the subscripts of I'a*.
Thus (2) is proved.
From (2) we have
(curl curl T) = 1 Sahi kD (-1-
1)'(,'_ Sab d D T 1

g a
(m + . .

1
Sfet...a DgDaTb...d
m!
which vanishes when we sum over g and a.
When Tbc...d is an m-vector, the summation in (1) or (2) may
be carried out in m + 1 stages by setting a = h, i, j, k in ,
turn and summing over the other m indices. Thus, from (1), we
have
(4) (curl T)hij...k = VhTij...k + (-1)mViTj...kh + ... ,
taking the m + 1 cyclical permutations of the subscripts hij .. k
in order and placing (- 1)' before the second, fourth, terms.
In the cases m = 1, 2, 3, we thus obtain
(curl T)ij = VjTj - ojTi,
(curl T)ijk = ViTjk + VJTki + VkTij,
(curl T)hijk = VhTijk - ViTjkh + VVTkhi - VkThij
When Tbc...d is an absolute m-vector, we obtain, from (2), an equa-
tion of the form (4) with V replaced by D.
When S is an absolute tensor of valence m - 1, T = curl S is
an absolute m-vector and curl T = 0. Then (4) becomes
(8) 0= (-1)'DiTjk...h + .
376 TENSOR ANALYSIS § 175

Thus if vi is absolute, T i j = S Davb is an absolute bivector, and


(9) DiTjk + DjTki + DkTij = 0.
174. Relation between Divergence and Curl. For alternating
tensors, we have the
THEOREM. If T be d is an n2-vector (n2 < n),
(1) dual curl T = div dual T,
provided dual T is taken contravariant when T is a scalar.
Proof. From § 170,
1
(dual curl T)pq"'r ii ... k(CUrlT) ij...k

Epq...r ij...kaab ..kdOaTb...d


(m + 1)!m!

pq... r ab... dVaT


b...d
M!

Va (1. E pq "r Tb ...d

= Va(dual T)P . ra
= (div dual T)pq*'*.
On taking duals of both members of (1), we have
(2) curl T = dual div dual T.
Moreover, if we replace T in (1) by dual T, we have also
(3) div T = dual curl dual T.
175. Parallel Displacement. A vector p is said to undergo a
parallel displacement along a curve xi = pi(t) when dp/dt = 0
along this curve. If p = pkek, we have
dp
-dpkek + pk aek dxi

axi dt
= dpk dxi
- ek + pk rikej,
dt dt
dt dt
or, on interchanging summation indices j, k in the last term,
dp
dt (+pr)ek.
dt
§ 175 PARALLEL DISPLACEMENT 377

Hence, if dp/dt = 0, the components pk satisfy the differential


equations:
dpk dx'
pj
r". = 0 (k = 1, 2, ... , ii).
(1) dt + dt
A solution pk(t) of this system, satisfying the arbitrary initial con-
ditions pk(0) = ak, defines a vector at each point t of the curve.
The vector ak at the point PO (t = 0) is said to undergo a parallel
displacement along the curve into the vector pk(t) at the point P.
In (1) dxi/dt depends upon the functions pi(t) defining the curve;
hence, in general, the solutions pk(t) will change when the curve
is altered. In other words, the vector pk obtained by a parallel
displacement of ak from PO to P depends, in general, upon the path
connecting these points.
The length of a vector p and the angle between two vectors p, q
remain constant during a parallel displacement; for, if dp/dt and
dql dt vanish along a curve, we have also
d d
dt(P.P) =0, dt(P - q) = 0.
We shall say that a vector remains constant during a parallel dis-
placement.
If s is the are along the curve,
ds2 = gi; dxi dx1 = g;xix' dt2.
If we choose the are as parameter (t = s), we have gjjziz' = 1, an
equation which states that the tangent vector dxi/ds to the curve
is of unit length. If a curve has the property that its unit tangent
vectors dxi/ds are parallel with respect to the curve, it is said to
be a path curve for the parallel displacement. With t = s, pk =
dxk/ds, (1) gives the differential equations of the path curves,
d2xk dxi dx'
0.
(2) ds2 t' ds ds -

The path curves are the "straightest" curves in our Riemannian


space-the analogues of straight lines in Euclidean geometry.
When equations (1) can be satisfied by functions pk(x', , x")
of the coordinates alone, the parallel displacement is independent
of the path, and the space is said to be flat. Then
dpk apk dxi
dt axe dt
378 TENSOR ANALYSIS § 175

and the ordinary differential equations (1) are replaced by the


partial-differential equations:
ap''
(3) axi + p'1' = V pk = 0.
Since Vipk are the components of Vp = e'Dip (§ 167), we see that
a flat space contains vectors p(x', , X'), such that Vp = 0, or

(4) Dip = 0 (i = 1, 2, ... , n).


Since pk = p ek, we see that (4) is equivalent to
(5) D,Pk = p Dzek.
For any fixed value of k, the n functions p D,ek are components
of a gradient vector, and for this it is necessary and sufficient that
Di(p D,ek) - D,(p Diek) = 0 (§ 172, theorem) ;
or, since Dip = 0,
p (DiD, - D,Di)ek = 0.
These equations must hold independently of the initial conditions
imposed upon p and are therefore equivalent to
(6) (DiD; - D,Di)ek = 0.
Since Di transforms like a covariant vector (167.3), the operator
(7) Di, = DiD; - D,Di = a bDaDb,
transforms like a covariant dyadic; for
ax, axb a2xb axa axb
D0Db,
DiD' -
(at.,

Da) \ax1 Dbl axiaxI Db + azs a


ax, axb
Di, _ DiD, - D,Di = -. Dab-
at, ax'
Moreover,
axa axb \1 (axc axa axb axc
Di;ek Dabec,

Caxi at-, Dab/ \axk ec/ at - 4921 axk

since D,,Dkxc = 0; hence


axa axb axc axh
eh DiJek = - -axk ed Dabec
at t at, axd
§ 175 PARALLEL DISPLACEMENT 379

This equation shows that e' Di;ek is an absolute mixed tensor


of valence four, say
(8) Rijk-h = eh Di;ek.
The components of this curvature tensor R are thus the coefficients
in the equation,
Rijkheh;
(9) Dijek =
the condition (6), necessary for flat space, now assumes the tensor
form,
(10) Rtijk = 0.
h

If it holds in one coordinate system, it holds in all.


When the space is flat, we can choose n linearly independent
vectors ai at a point P and, by giving them parallel displacements
to neighboring points obtain a set of constant base vectors ei = ai
in a region about P. For these base vectors, we have
gi; = ai a; = const, I'k = 0 (166.6),
and the corresponding coordinate system x is called Cartesian.
To determine the Cartesian coordinates y = x corresponding to
the base vectors ai, we have
ayr
ek =axk a,.,

aek c12yr
ar,
OX, ax' axk
and, on dot-multiplying (11), member for member, by
ay8
- er = a8 ,
axr
ays a2y3
r t
(12)
ax, rA axi axk
From (11), we have the necessary conditions for the integrability
of equations (12):
D;ek = Dke;, Di,jek = 0;
t This equation also follows from (163.9) with Tj = 0.
380 TENSOR, ANALYSIS § 176

that is,
r r h
rjk = rkj, Rijk = 0.
These conditions are also sufficient for the complete integrability
of equations (12).$ When these conditions are fulfilled, (12) ad-
mits solutions y(x', x2, , x") which with ay/axi take on arbi-

trary values at a given point. If we place the origin of the Car-


tesian coordinate system at the point x0, we have y = 0 when
x = x0; and, if we choose n linearly independent sets of initial
values,

... , = pn (i = 1, 2, ... , n),


axi - pi, a
= p2,
ax"
when x = x0, we obtain n corresponding solutions y'(x) whose
Jacobian I ay/ax I = det p 0 when x = x0. In the region about
x = x0 for which I ay/ax 19 0, the n functions yj(x) thus obtained
define a Cartesian coordinate system. In brief, we have the im-
portant
THEOREM. A necessary and sufficient condition that a Riemann
space, with symmetric afne connection, be flat is that its curvature
tensor vanish identically.
We may readily compute the components Rijkh from (8):
*.k'; eh (DiDjek - D;Diek)
= eh {Di(r;ker) - Dj(riker)I
eh {(Dirjk)er - (Djrik)er} + eh {rjkrires - rikrjres},
and, on putting eh e, eh e8 = 887 we have

(13) Ri;kh = D2-r' - Djr k + r; r r -


176. Curvature Tensor. From the defining equation for the
curvature tensor,
Dijek = ji rer,
we obtain the covariant components,
(1) eh Dijek = Rijk r9rh = Rijkh
$ Cf. Veblen, op. cit., p. 70-1.
§ 176 CURVATURE TENSOR :381

We now can express the covariant curvature tensor Rijkh in terms


of the gammas :
Rijkh = eh (DIDjek - D;Diek)
= eh' {Di(rjk,re') - Dj(rik,rer)} (166.4)
Sh{Dirjk,r - Djrik,rI - 0h{ rjk,rria - rik,rrja};
(2) Rijkh = Dirjk,h - Djrik,h - rjk,rrik + rik,rrjh
Since
rjk,rrih = grerikrih = rjkrih,s,
we also may write (2) in the form:
(3) Rijkh = Dirjk,h - Djrik,h - rikrih,r + rr,krjh,r.
Rijkh has the following types of symmetry:
Rijkh + Rijkh = 0;
Rijkh + Rijhk = 0;
Rijkh + Rjkkh + Rkiih = 0;
Rijkh - Rkhij = 0.
Proofs. (I) follows from Dij = -Dji. Since the scalars ek eh
= gkh have continuous second derivatives (§ 169),

gives (II). Thus Rijkh is alternating in its first two and last two
indices.
Since the affine connection is symmetric (§ 169), we have
(4) Die; = r er = r;ier = Djei.
Hence, on adding the identities,
Dk(D1ej - Djei) = 0,
Di(Djek - Dkej) = 0,
Dj(Dkei - Diek) = 0,
we obtain
Dijek + Djkei + Dkjej = 0,
which, on multiplication by e h- , gives (III).
382 TENSOR ANALYSIS § 176

Now (IV) is a consequence of (I), (II), and (III). From (III),


we have
Rijkh + Rikih + Rkiih = 0,
Rjkhi + Rkhji + Rhjki = 0,
-Rkhij - Rhhkj - Rikhj = 0,
- Rhijk - Rijkk - Rjhik = 0.
When we add these equations and make use of (I) and (II), only
the underlined terms survive, namely, 2Rijkh - 2Rkhij, and we
obtain (IV).
The symmetry relations (I) to (IV) reduce the number of inde-
pendent components of Rijkh to i22n2(n2 - 1).
Proof. Rijkh = 0 when i = j or k = h (I, II); hence, in general,
the number of non-zero components is (nC2)2 = n2(n - 1)2/4.
But, when ij ; kh, these are paired, because Rijkh = Rkhij (III);
hence, if we add the number nC2 of unpaired components Rjjij to
the preceding total, we obtain double the number of components
with distinct values. The number of components with distinct
values thus is reduced to
1n2(n - 1)2 n(n - 1)
1=*n(n-1)(n2-n+2).
4 + 2

These are still further reduced by the ,,C4 relations (III); for
i, j, k, h must all be different in order to get a new relation. If,
for example, i = j,
Riikh + Rikih + Rkiih = Rikih + Rkiih = 0
is already included in (I). The number of linearly independent
components is therefore
n(n - 1)(n - 2)(n - 3)
n(n - 1)(n2 - n + 2) -
24
= n2(n2 - 1).
For n = 2, 3, 4 this gives 1, 6, 20 linearly independent compo-
nents Rijkh, respectively.
Example 1. When n = 2, the contracted product,
eijEkhR,,kh
= 4R1212,
§ 176 CURVATURE TENSOR 383

is a relative scalar of weight 2 (§ 169); hence R1212/g is an absolute scalar.


Now, from (2),
(5) R1212 = D1r21,2 - D2r11,2 - r21,.ri2 + rll,rr22
We shall compute this expression when the base vectors are orthogonal:
g12 = 0. From (166.6),
rll,1 = !Digll, r11,2 = -iD2g11,
r12,1 = ZD2911, r12,2 = 2D1g22,
r22,, = -2D1922, r22,2 = 3Dzg22.

Moreover, since g = 911922, 911 = 1/911, g2 = 1/g22; hence


r21,rr12 = r21,,ri2 + r21,2r12
= r21,,r,2,1 911 + r21,2r,2,2 g22
- (D2g11)2 + (D1922 )2

4911 4922

r11,rr22 = rll,lr22 + r11,2r22


= r11.1r22,1 911 + rll,2r22,2 g22

(Dig,,) (D1922) (D2911) (D2g22)


491, 4922
Substituting these results in (5) gives
81212 - DiD1922 + 2 D2D2911
Dlgll + D1922 D19zz + Dw11 + D2922 D2g11 l
j
St l\ 911 922 911 922 J

2
The absolute scalar, \ J

(6) K
_81212 1 (
Dl
1
D19zz +
\
D2 911 J ,
9 219 (.\/g- D2
is precisely the total curvature of the surface whose fundamental differential
form is
ds2
= gll dx1 dx1 + 922 dx2 dx2;
for, if we put x1 = u, x2 = v, 911 = E, g22 = G, H =, (6) agrees with
(139.6).
Example 2. We may contract the tensor,
(7) h
R41k = ghrRijkr,
in essentially two different ways.
384 TENSOR ANALYSIS § 177

With h = k, we have
k
(8) Rt1x = gkrRiikr = 0,
since gkr and Rijkr are, respectively, symmetric and antisymmetric in k, r.
From (175.13), we see that (8) is equivalent to the identity:
(9) Dirjr = Djr;,.
With h = i, we obtain the Ricci tensor,
(10) Rik = R0 ' = 9ihRijkhi
this is a symmetric dyadic; for, from (IV),
(11) Rkj = 9''`Rikih = 9'Rihik = ghiRhiki = Rik.
The first scalar of this dyadic,
(12) R = gikRik = gikgihRijkh,
is an absolute invariant. In the case n = 2, g12 = 0 considered in ex. 1, we
have
(13) R = 2g11g22R2112
= -2R1212/9 = 2K.
177. Identities of Ricci and Bianchi.
Ricci Identity. In analogy with
Dij = D,Dj - DjDj, we also write Vij = ViVj - V, Vi.
With this notation, (167.21) becomes
(1) VijT = DijT
for any tensor, absolute or relative, with its base vectors. On the
left Vii acts only on scalars (cf. § 167, ex.); on the right Dij acts
only on base vectors, for Dij<p = 0 when cp is a scalar. Equation
(1) yields the Ricci identity when we evaluate DijT by making use
of the formulas :
(2)
Dijek = Rij'heh (175.9),
Dijek = - Rijh
(3)
keh.

Equation (3) is proved as follows:


gkryh8Rijrhee
Dijek = Dijgkrer = gkrR,;r.aea =
_ -g gkrRijhrea
ha
= R;jhkeh .
For the vector vkek, we have
Dijvkek = vkDijek = vkRijk aea = vaRija kek,
(4) Vijvk = v8RtijB'k.
§ 178 EUCLIDEAN GEOMETRY 385

Similarly, for vkek,


ke3
Vijvkek = vkDijek = -vkRijs = -v3Rijk'ek,
(5) Vijvk = -v8Rijk8.
The general Ricci identity now is readily established. Thus, if
the components of T are Thk:::;,a, for every upper index * in T,
DijTh::...m contains a term, Th..' ..mRij,;
and, for every lower index * in T,
vijTh::;:: ,, contains a term, ,
Bianchi Identity. At the origin of geodesic coordinates (§ 163),
we have, from (174.13),
DiRjkhm = DiRjkhm = DiDjrkh - DiDkrjh.
By permuting ijk cyclically in this equation, we obtain two others.
On adding the three equations, we find that the right members
cancel; we thus obtain the Bianchi identity:
(6) DiRjkhm + OjRkihm + OkRijh-m - 0.
Since this tensor equation holds at any point, it is also true for
general coordinates.
178. Euclidean Geometry. When the space is Rat, we can de-
termine a Cartesian coordinate system xi (§ 175). The corre-
sponding metric tensor gij has then constant components.. If in
addition the metric form gijxixj is positive definite, the space and
its geometry are termed Euclidean. We can then always make a
real linear transformation to coordinates yi for which gij = b, the
Kronecker delta, and the metric form becomes a sum of squares
(§ 150) :
b yiyj = y'y' + y2y2 +.... + yny+l.
The corresponding base vectors ai then form an orthogonal set
(ai - aj = b), and the coordinates yi are said to form an orthog-
onal system.
Let xi denote a Cartesian coordinate system with the base vec-
tors ai. If we transform to another Cartesian system xi with the
base vectors Ai, we have
aj =
-
8xt
ai
386 TENSOR ANALYSIS § 178

and, since both a, and ai are constant,


aa; a2xi exi
ai = 0, = 0.
a2k aXk a2 82k a2'

On integrating this equation twice, we have


(1) xi=c;2'+Ci,
where c and Ci are sets of constants. The transformation between
any two Cartesian coordinate systems is therefore linear with con-
stant coefficients. As in the general transformations of § 157, we
require that the Jacobian,
ax'
(2) det - = det c 54 0.
a2'
The transformations (1) with non-vanishing determinant form a
group-the affine group.
When the Cartesian coordinate systems y and y are both or-
thogonal, the law of transformation,
aya
- gab, becomesaya
gi; = ayi ayb Ji- aya = S;.
09)
If we multiply this equation by ayi/ayk and sum with respect to i,
we obtain
ayk ay'
(3) = ayk
ay;
Since orthogonal coordinate systems are also Cartesian, the trans-
formation between two orthogonal systems has the form:
(4) yk = c;4j' + Ck.
The inverse transformation is
(5) yj = Yk(yk - Ck),
where Yk is the reduced cofactor of c; in det c;. Equation (3) thus
becomes
(6)

This is precisely the condition that the matrix c; be orthogonal


(§ 149, theorem); thus a transformation between orthogonal coordi-
§ 178 EUCLIDEAN GEOMETRY 387

nate systems is orthogonal (has an orthogonal matrix). Condition


(6), which characterizes an orthogonal matrix, also implies that
(7) c'iCj = cic = Sj,
in view of the relations (146.8) between reduced cofactors. Con-
ditions (7), in turn, imply (6); either (6) or (7) is a necessary and
sufficient condition that the matrix c; be orthogonal.
Two orthogonal transformations,
yz=ay'+Ay, yj =bkyk+Bj,
have an orthogonal resultant,
s i i i, j;
yi
we have, for example,
r r r a r s a s a
cic; = (aebi)(atbj)t
= Stbib; = bib;
t
= S;. i

Moreover (5), the inverse of (4), and the identity transformation


yi = S;yj are orthogonal. Consequently, the orthogonal transforma-
tions form a group.
In view of (3), the equations,
ayk ayj
Vi = ayj vk, U' =ayk tlk

show that covariant and contravariant vectors transform alike


under orthogonal transformations. Within the orthogonal group,
the distinction between covariance and contravariance vanishes, and
tensor components may be written indifferently with upper or
lower indices. For example, we may write Sij or Sij for the Kro-
necker delta.
The orthogonal group of transformations admits as a subgroup
those transformations for which
ayy
(8) det c = = 1.
a
If we regard (4) as a transformation between the points y and y
in the same Cartesian coordinate system, the transformation is
called a displacement or rigid motion. In fact in 3-space the trans-
formation (5) may be written
(9) s=r-a,
388 TENSOR ANALYSIS § 179

where the dyadic CF in non ion form is given by the matrix (ck).
Since this matrix is orthogonal and its determinant is 1, the trans-
formation (9) is a translation followed by a rotation (§ 75, theo-
rem), in brief, a displacement. The subgroup characterized by
det c = 1 is therefore called the displacement group. In view of
(8), we see that, within the displacement group, the distinction be-
tween absolute and relative tensors also vanishes.
A displacement which leaves the origin invariant is called a
rotation. Thus the transformation y' = c;y' is a rotation if the
matrix (c) is orthogonal and has the determinant 1. Rotations
form a subgroup of the displacement group.
179. Surface Geometry in Tensor Notation. The equations,
x' = xi(ul) u2) (i = 1, 2, 3),
define a surface embedded in Euclidean 3-space. The space co-
ordinates xi are rectangular Cartesian and are designated by italic
indices (range 1, 2, 3); the surface coordinates u" are curvilinear
and are designated by Greek indices (range 1, 2). If we write
al = i, a2 = j, a3 = k, the position vector to the surface is
r = xiai. The metric tensor in space is then
(1) Si; = ai - a;.
If we limit the coordinate transformations x - x to the displace-
ment group (§ 178) the distinction between covariance and contra-
variance as well as the distinction between absolute and relative
tensors does not exist in 3-space.
First fundamental form. The base vectors on the surface are
ar ar axi
(2) a"=-=--=x.ai,
au" axi au"
where x" = axi/au". Since
e"-e#= =x'xp'si; =xaxs,
the metric tensor for the surface is
(3) gap = e., ep = x;.xP.
This tensor defines the first fundamental form on the surface:
ds2 = g"g du"du#.
§ 179 SURFACE GEOMETRY IN TENSOR NOTATION 389

Note that x, is a covariant surface vector; for if we make the


transformation u -+ u, we have
axe axi auo
aic" aufl au"
If a vector v has the "surface components" v" (a = 1, 2),
v = vae« =
and vi = v"xa (i = 1, 2, 3) are the "space components" of v.
Unit surface normal. The space vector,
N = el x e2 = Eijkaixix2,
has the components Ni = Eijk xix2; moreover
N2 = EijkEist xix2 x1x2 = sit xix2 XIX2
= xix2 xix2 - xix2 xix2 = 911922 - g12

Thus N2 = det ga# = g t; hence the unit normal n to the surface


has the components Ni/\:
(4) ni = ni = eijk xix2.

Second fundamental form. On the surface with metric tensor


gang, the Christoffel symbols are given by (cf. § 166)
(5) is = ZgX7(D«goti + DRgy« - D7g«#)
Covariant derivatives are then computed from the formulas of
§ 167. In particular we have, from (167.13),
(6) Vaxkp = DaDftxk - VPxQ.

Since the covariant derivative of the metric tensor gij is zero


(167.19),
V. xksxy = 0.
(7) Vagoti =

If the product rule (§ 168, 1) is used, this equation and its cyclical
permutations give
(7a) xyV 4 + x0Vaxy = 0,
(7b) xgVVxy + xyVflxx = 0,
(7c) xkovyxa + xxVyx1 = 0.
t This also follows from the expansion of (el x e2) (e1 x e2) given in (20.1).
390 TENSOR, ANALYSIS § 179

Subtract the third equation from the sum of the first two; then,
in view of (6), we obtain
xyDax3 = 0.
This equation states that the space vector Vax is perpendicular
to both el and e2 (whose space components are 4, x2); that is
Vax, is a multiple of the unit normal nk, say
(8) oax3 = ha#nk.
The symmetric covariant tensor,
k
(9) ha0 = nkVaXQ = hha,
defines the second fundamental form on the surface: ha#duadus.
Derivative formulas. From (6) and (8) we have
(10) Daxkg =F xk\ + haonk;
these equations are the derivative formulas of Gauss. If we adjoin
the (constant) base vector ak to each term they become
(10)' Dae# = raaea + hafln. f
On multiplying (10) by xry and summing on k, we have also
(11) 'xk Da X = raggay = rap,7.
Since nk is a space vector with no components along el or e2,
Vank = Dank. On differentiating nknk = 1, we obtain
(12) nkVank = nkDank = 0;
hence Dank (1 nk) is a tangential surface vector. Similarly, from
nkx' = 0, we obtain
nkvaxp i xpvank = 0;
or, in view of (9) and the symmetry of hao,
(13) hp = -xxVank = -xaV nk.
Hence
hapxa = -9aavpnk,
90XhaflxX o Vpnk = - Yank,
(14) Dank = Dank -
-ham xa.
t Note that the term hapn in this equation is in apparent disagreement with
(163.1); this is due to the fact that our 2-dimensional geometry is not intrinsic
but that of a 2-space embedded in a 3-space.
§ 179 SURFACE GEOMETRY IN TENSOR NOTATION 391

These equations are the derivative formulas of Weingarten. If we


adjoin the (constant) base vector ak to each term, they become
(14)' Dan =
Equations of Codazzi and Gauss. From (8) we have
Osxy = hsynk,
VaV#xky = (Vahs.y)nk + hsyVank,
= (Vahsy)nk - hsyhaIxk
in view of (14). Now form V$Vaxk, and subtract it from the last
equation; writing Vas for the operator VaVs - we thus
obtain
(15) VasXY = (Vahsy - V$hay)nk + (hayha"' - hsyha")xx.
If we replace Vasx1' by the value,
X xkx,
Vasxak = -Raay
given by Ricci's Identity (177.5) and adjoin ak to each term, (15)
becomes
(16) -Rasyx ex = (Vahsy - oshay)n + (hayhft" - hsyhal')ex.
This vector equation is equivalent to the scalar equations :
(17) 0 = Vahsy - vshay,
(18) ROT" = hsyha)` -
Equations (17) are the Equations of Codazzi. When a = /3 the
right member vanishes identically; and an interchange of a and 0
repeats the same equation. Hence there are but two independent
Codazzi equations; these may be written with a = 1, S = 2, y =
1, 2;
(19) V1h21 - V2h11 = 0, V1h22 - V2h12 = 0.
On multiplying (18) by gas and summing on X, we obtain the
covariant curvature tensor:
(20) Rasys = hsyhas - hayhss.
From the symmetry has = hsa we may verify at once the four
symmetry relations of Rasys :
Rasys = - Rsays, Rasys = - Rassy,
Rasys + Rsyas + Ryass = 0, Rasys = R,ysas.
392 TENSOR ANALYSIS §180

As Greek indices range over 1, 2, these relations show that there


is but one independent equation (20). This Equation of Gauss
may be taken as
22
(21) - 81212 = h11h22 - = h,

where h = det han.


Total and mean curvature. The contracted product,
E"S Erya R,,,#,3 = 4 R1212 and g = 811922 - 912
are both relative scalars of weight 2; hence,
(22) K = -R1212/9 = h/9
is an absolute scalar, namely the total curvature of the surface
(§ 176, Ex. 1).
The mean curvature of the surface is defined as the absolute
scalar J = g"0hg.
Since g"R is the cofactor of gag in det ga#,
(23) 911 = 922/9, 922 = 911/9,
912 = 921 = -912/9;
hence
9a,6haQ
(24) J = = (922h11 - 2g12h12 + 911h22)/9

180. Summary: Tensor Analysis. Under general transforma-


tions,
ax
.ti ={i(x1,
J
x2, . . , xn)
, 001
ax
the component of a relative tensor of weight N transforms accord-
ing to the pattern:
ax N axi ax' axc b
Ti'.k =
ax axa axb axk

When N = 0, the tensor is absolute. For brevity, components of


tensors often are called tensors.
The number of indices on a tensor component is called its
valence. In n-space a tensor of valence m has nm components.
A tensor of valence zero is a scalar. A scalar p(x) has one com-
ponent in each coordinate system given by
ax N
,7p (.t)
= I ax I
'P(x)
§ 180 SUMMARY: TENSOR ANALYSIS 393

A tensor of valence one is a vector. The differentials of the co-


ordinates and the gradients of an absolute scalar are the prototypes
of absolute contravariant and covariant vectors :
axz
dx' _ - dzr,
axr
-- _ a.-axr-
axi axraxi

Measurement is introduced into Riemannian geometry by the


non-singular quadratic form:
ds2 = gi; dxi dx.' (gi; = gji, g = det gi; F6 0).
The character of the geometry depends upon the choice of the
n(n + 1)/2 functions gij(x) of the coordinates. The relations
ei e; = gij determine the lengths of the base vectors ei and the
angles between them.
Any vector in n-space at the point x is linearly dependent upon
the n base vectors ei at this point. The reciprocal base vectors
e' are defined by e' . e; = S . If g = det gij and g'' is the reduced
cofactor of gij in g,
(1} ei = girer, ei = girer;
(2) ei e; = gi;, ei e' = S, e' e = gi'.
In passing from coordinates x to x, the transformation of the
base vectors is prescribed by
axi
(3) ei=-axr
er,
axi
ei=-er.
axr
These equations show that gjj, M, g2' are components of an abso-
lute dyadic, the metric tensor G = gi,eiel = e;e'.
Use of equations (1) permits indices to be raised or lowered on
tensor components:
9irT.r., T.i. = girT*r.
If T is a tensor of weight N (say T = T?'.ke,ejek),
IN
ax
T = - T.
at l

Addition of tensor components of the same valence, weight, and


type produces a component of this same character.
394 TENSOR ANALYSIS §180

Multiplication of tensor components of valence ml, m2i of weight


N1, N2, and of arbitrary type produces a component of valence
m1 + m2 and of weight N1 + N2.
Contraction of a tensor of valence m > 1 results on forming the
dot product of any two of its base vectors. If the vectors in ques-
tion are ei e' = s, the components of the contracted tensor are
obtained from the original components by putting i = j and per-
forming the implied summation.
The components of the affine connection rk are functions of the
coordinates defined by
(4) Die; = I'er (rk = ek Diei).
Then also
(5) Die' = - rrer;
(6) DiG = 0 (G = eye');
(7) Dig = 2grir
By definition,
ri,,k = gkrr, then rk. = gkrri;,r.
When r k is symmetric in ij,
13

ri7,k = 2(Digik + Digki - Dkgi.l)


The gradient of a tensor of weight N and valence m is defined
as the tensor,
N N
(9) VT = g2ehDh(g 2T)
of weight N and valence m + 1. When T is absolute, VT =
ehDhT.
The components of VT, denoted by prefixing Vh to the compo-
nents of T, are called covariant derivatives. For any tensor
we have
(10) VhT : = DhT; - NrhrT:
+ (T -rhr + ...) - (TK:rr +
the first parenthesis contains one term for every upper index, the
second contains one term for every lower index.
The metric tensor G is absolute and VG = 0, from (6); hence
Vhgii = 0, Vhba = 0, Vhgii = 0.
§ 180 SUMMARY: TENSOR ANALYSIS 395

Since g is a relative scalar of weight 2, Vg = 0 from (9), and


Ohg=0.
The covariant derivatives of the epsilons and Kronecker deltas
are zero.
The divergence of a tensor T is defined as the gradient VT con-
tracted on the first and last indices: thus
VhTij...kh;
=
(divT)ij...k

when T is an m-vector of weight 1, we may replace Oh by Dh.


The curl of a covariant tensor of valence m < n is de-
fined as the (m + 1)-vector:
1 abc...d
(curl T)hii...k = ' Shij...kVaTbc ...d.
m.
When T is absolute, we may replace Va by Da.
A Riemannian n-space xi with the metric tensor gij and base
vectors ei has the associated curvature tensor,
RiJk h -eh (DiDj - DjDi)ek.
Its covariant components,
Rijkh = eh (DZDj - D;Di)ek,
have four types of symmetry:
Rijkh + Rjikh = 0, Rijkh + Riihi = 0,
Rijkh + Rikkh + Rki h = 0, - Rkhij = 0.
Rijkh
These relations reduce the number of linearly independent com-
ponents Rijkh to n2(n2 - 1)/12. When n = 2, there is but one
independent component, say R1212; and the absolute scalar
-R1212/g = K, the total curvature of a surface whose funda-
mental form is gij dxi dxj.
A Cartesian coordinate system yi is one in which the components
gij of the metric tensor are constants; then all t = 0, and the
base vectors ai remain invariable in space (aaj/aye = 0).
The Riemannian space xi with metric tensor gjj is said to be
flat if it is possible to transform to a Cartesian coordinate system.
When the affine connection is symmetric, a necessary and suffi-
cient condition for a flat space is that the curvature tensor vanish.
If the space is flat and the metric form gijxixj is positive definite,
the space and its geometry are termed Euclidean. We then can
396 TENSOR ANALYSIS

make a real linear transformation to orthogonal coordinates yi for


which gij = S; the metric form then becomes a sum of squares.
A transformation y i = c)y' + C' between orthogonal coordinate
systems is characterized by the relations:
ayi ay'
or c; = y= (orthogonal matrix);
ay' ay,
then
crx; = cTcT = S, and det c; _ ± 1.
Within this orthogonal group of transformations, the distinction be-
tween covariance and contravariance vanishes. Orthogonal trans-
formations for which det c) = 1 form the displacement subgroup in
which the distinction between absolute and relative tensors van-
ishes.
When yi and y' are regarded as points in the same coordinate
system, the transformation yi = c)y' is a rotation when its matrix
is orthogonal and its determinant +1.

PROBLEMS
Summation Convention. Index range is 1, 2, 3 unless otherwise stated.
1. Prove the following:
(a) Etjk ETjk = 2! Si;

(b) Eijk Srst = 3! Erst;

(c) Si = 3, S =3 2, Stjk = 3! [§ 161.1

2. Show that
u1 u2 143

V1 v2 v3 = Eijk UiVjWk.
,W3
W1 1U2

3. Show that the two-rowed determinant formed by columns i, j of the


matrix,
xi x2 ... xn
y2 ... yn) is ST8 xrys.
Y

4. For the dyadic vi, in (77.1) show that the scalar invariants are

'P1 = 'Pit, 'P2 = fit, 'P3 = 1 Eijk Erat 'Pzr'Pjs'Pkt;


3!
and that the vector invariant has the components z fi k'Pik.
5. Show that the general solution of the equations:
aix' = 0, bix' = 0 is x' = X ajbk.
PROBLEMS 397

6. Show that the cofactor Arl of the element a= in a = det a is

Atr = 1 Eiik Erat


2
[ At ali -=2l E
rat Eiik aPaa
: f k
' = 'll
= E rat E pat a = b r a.]
7. If yl, y2, y3 are functions of x1, x2, x3, det (ay'/axi), is called their Java
bian and written
a(yl, y2, y3) ,
or, more briefly, ay
a(xl, x2, x3) x

If z1, z2, z3 are functions of yl, y2, y3, prove that


Oz az
ay
ax ay ax

If, in particular, the functions


z` (yl, y2, y3) = x`,
show that
ay ax
= 1.
ax ay
8. Prove that
a(y', y', yk) = Stik a(yl, Y2, y3)
x3) .
x,' xt) rat a(xl'
a( xr' x2,
9. Prove that the
t
Cofactor of ay. in
axi
ay
Ox
I
isax'
-. I ay
ay' ax r
ay' axi ay i = t
= ak
ax' a yk a yk -
10. If the elements of a = det aii are functions of x1, x2, x3, prove that
as _ aa,t Aat
axr axr
11. Prove that
a ay axi
02y'

ay
axr ax axr axi ayi ax

[Apply Probs. 9 and 10.]


12. Prove that the bordered determinant:
vl V2 V3 0
all a12 an ul
= uiviAii.
a21 a22 a23 U2

a31 a32 a33 u3

This determinant is formed by bordering det aii with the vectors vi and ui.
If aii is symmetric it also equals viuiA `i.
398 TENSOR ANALYSIS
13. When the index range is from 1 to n show that the bordered determi-
nant (written compactly),
V, 0
_ 1)n+lusy7iAi7 (i, = 1, 2,
aii ui
14. If Aij is the cofactor of aij in a = det aii, and A = det Aii, show that
aA = a3 (and hence A = a2 when a 96 0). State the corresponding theorem
when the index range is from 1 to n.
15. Show that the n-rowed determinant,
a = det apq = ij"'k aii a2i ... ank

= j8 ... E
ii...k
ari aaj ... aek,
n!
and that the cofactor of ari is
Ari = 1 Erd... Eij...k a8. . . . atk
(n - 1)! '

16. If det aii is symmetric (aii = aii), show that det Aii is also (A ii = Aii).
17. If the n-rowed determinant a = det aii is antisymmetric (aii = -aii),
show that a = 0 when n is odd.
18. Show that the linear equations,
aij xi = 0, (i, j = 1, 2, , n),
for which det aij = 0 and not all the cofactors Aij vanish has a non-zero solu-
tion of the form xi = Aki for some value of k.
19. If aij and gii are symmetric dyadics and gijx`xi is a positive definite
quadratic form (§ 150), prove that the roots of the cubic equation,
(1) det (aii - X gii) = 0, (i, j = 1, 2, ... , n),
are all real.
[The system of n linear equations,
(aij - X gii)zj = 0,
must have a solution zi = xi + iyi other than (0, 0, , 0) when X = a + 1$
is a root of (1). Hence
[aii - (a + l(x1 + iy') = 0;
i$)gii11

and, on equating real and imaginary parts to zero, we have


(2) aii xi - a gii x' +,6 gii y' = 0,
(3) aij yi - a gii yj - 0 gii xi = 0.
On multiplying (2) by yi and (3) by xi, and subtracting, we find that
0(gii xx' + gii yy') = 0, and hence 0 = 0;
for if gii xx' = gij y'y' = 0, xi = yi = 0 and consequently zi = 0, contrary to
hypothesis.]
PROBLEMS 399

Tensor Character.
20. Prove the theorem: If aii, bii, cj are absolute dyadics,
a = det aii, b = det bii, c = det cj
are scalars of weight 2, -2, 0; the cofactors Aii, Bii, Ct in these determinants
are dyadics of weight 2, -2, 0; and the reduced cofactors A"/a, Bii/b, Cs/c
are all absolute dyadics. [Cf. Prob. 15.1
21. Show that, if u, v, w are absolute vectors,
ui u2 u3 u1 U2 u3
Vi V2 V3 and vl V2 V3

W1 W2 W3 W1 W2 w3

are relative scalars of weight 1 and -1.


22. If aii and b" are absolute antisymmetric dyadics in 3-space, show that
(a) a23, a31, a12 are components of a contravariant vector ui of weight 1;
.(b)b23, b31, b12 are components of a covariant vector vi of weight -1:
[ux. = 1 Eijk aik; Vi = 2 Eiik bik].
Z

23. If T1. ..,,, Sz1"'h are absolute n-vectors (§ 170) in n-space, prove that
T12...n and S12 ' are relative scalars of weight 1 and -1.
ti ...h
[n! T12... , = e T"...n.]
24. If vi, ui, aii are absolute tensors, prove that the bordered determinant
in Prob. 13 is a relative scalar of weight 2.
25. Prove the "quotient law": If the set of functions viTi:: a are tensor
components of the type indicated by the indices for all absolute vectors vi,
then Tz:: c is a tensor of the same weight.
26. If uk is a covariant vector, prove that the total differential equation
uk dxk = 0 has the same form in all coordinate systems.
This equation is said to be integrable when there exists a function X such
that Auk = aw/axk; for the equation is then equivalent to dp = 0 and _
const is an integral. Show that
eilk uiDjuk = 0

is a necessary condition for the integrability of ilk dxk = 0 and that the form
of this condition is the same in all coordinate systems.
27. Let u, v be quantities (scalars, base vectors, tensors) whose "product"
uv is distributive with respect to addition but not necessarily commutative.
Show that the differential operator Dii = DiDi - DiDi has the property,
Dii(uv) = (Disu)v + u(Di v).
28. Show by direct calculation that the operator Dii transforms like a co-
variant dyadic:
ax, axb
Dii = . - Dab.
a2i a2i
400 TENSOR ANALYSIS
29. Deduce (177.3) from (177.2) by applying the operator Dii to eh ek
= Bt- [Cf. Prob. 27.1
30. If T is a tensor (complete with base vectors) of weight N and valence in,
prove that eiei DiiT is a tensor of weight N and valence m + 2.
[Since E-NT is an absolute tensor, R5'-NT = E-NT; hence, from Prob. 28,
eie_i N_ ,.-axa -axb
Di (R-,T) = .
e N eaeb N
axi ax'
Thus (Prob. 26),
eie1 Dii(E-NT) = E-N eiei DiiT
is an absolute tensor, which multiplied by EN yields the relative tensor
eiei DiiT of weight N.]
31. In Riemannian n-space with the metric tensor gii show that the gradient
of an absolute scalar p(xl, . , xn) is given by

Vv = eiDi,p = eigii Div = 1 (Div)e7Gi1,


9

or, in view of Prob. 13, by the bordered n-rowed determinant


(-1)n+i ei 0
(1) vw =
9 19ii Disc
Show that the scalar product of (1) by VV = elDkv gives
(2) (-1)n+1 Disc 0
9 19ii Div I

32. In Riemannian n-space with the metric tensor gii show that the diver-
gence of an absolute vector v is given by

div v Di(9 D' (-!'L G t')


- V9-
or by the bordered n-rowed determinant,
(-1)n+l Di 0
(1) dlv v =
9
1
1/9 9ii Vi

where the determinant is to be expanded according to the elements of the


first row and the operators Di applied to differentiate their cofactors.
In particular, if v = Vv, vi = Dip, we obtain the Laplacian,
(_1)n+1 Di 0
1
V9- 9ii Di'v
9

33. The equations of a surface in 3-space are xi = xi(ul, u2). Show that
au t
za = (i = 1, 2, 3; « = 1, 2)
au
PROBLEMS 401

is a contravariant vector in 3-space (1 = 1, 2, 3; a fixed) and a covariant vec-


tor in the 2-space (a = 1, 2; i fixed) formed by the surface.
Prove that
i x9k
pi = 21eao ziik xa
is an absolute covariant space vector normal to the surface. Write out its
three components in full.
34. In special relativity it is customary to use the independent variables,
xl=x, x2=y, x3=Z, x4=ct
where c is the speed of light. The interval ds between two events in space
time is defined by
(1) ds2 = CO dt2 - dx2 - dye - dz2 = gii dxi dxi.
Hence the metric tensor is given by
(2) 9i1 = g22 = 933 = -1, 941 = 1, gii = 0 (i j).
If v is the speed of a particle relative to a frame j5, we have (ds/dt)2 = c2 - v2
from (1), and hence

(3)
ds
dt=y
c
where y=(
/1-- v' \\
)

In a rest frame ao attached to the particle a clock registers the proper time r.
Putting v = 0, t = r in (3) we have
ds _ dt
(4) c and
dr = dr = y.
Corresponding to the position vector (x, y, z) in space we now have the
event vector xi = (x, y, z, ct) in space time. The velocity and acceleration four-
vectors are now defined as
(5) U =dxi- , W, =dui-
dr dr

Denoting t derivatives with dots, prove that


Eli _ _y (1, y, Z, c), ui = 'Y(-x, -y, -Z, c); utui = c2;
w' 'Y(7x + -Yx, 'Yy + 'Yy, 'YZ + ?'z, -ic); wtui = 0.
Thus the velocity four-vector has the constant magnitude c and is always
perpendicular to acceleration four-vector.
Show also that in the rest frame ao (t = r),
uo = (0, 0, 0, c), uo = (x, y, z, 0).
Covariant Differentiation.
35. Prove the product rule (§ 168, 1) for covariant differentiation without
resort to geodesic coordinates.
402 TENSOR ANALYSIS
36. If Ti iis an absolute bivector, show that
1
div T = vi 7'ij = - Dj(V'g Tij).

37. If u and v are absolute vectors, prove that


Dh(gij UV) = ui. Vhv2 + vi Vhuh.
If I u I is the length of u, prove that

DI I- Ui Vh U

38. Prove that


Vijvk + vjkvi + Vkivj = 0. [Cf. (177.5).
39. Prove that
Rijkh + RJkih T RkjJh --0-
By contracting this identity with h = j, show that the Ricci tensor Sik (§ 176,
ex. 2) is symmetric.
40. In the Bianchi identity (177.6) raise the index h and contract with
j = h, k = m to obtain
vi Rjk'k + Vi Rki + vk Ri1'k = 0.
On introducing the mixed Ricci tensor,
Rij = R = Rii = Rr4i
(its first scalar), show that this equation implies that the divergence of the
tensor R*j - 1SjR is zero.
41. Check the results of § 164, ex. 1, by differentiating r = pR(,p) + zk
twice with respect to the time. [Apply (44.2) and (44.3).]
42. The generalized Kronecker delta 6a'.".' d has k subscripts and k super-
scripts (1 5 k 5 n), each ranging from 1 to n. Its value is defined as follows:
If both upper and lower indices consist of the same set of distinct numbers,
chosen from 1, 2, , n, the delta is 1 or -1 according to the upper indices
from an even or an odd permutation of the lower; in all other cases the delta
is zero.
Prove that if Som.': d has 2k indices,
(n - k)! Sab...d - E

where the epsilons necessarily have n indices. Hence show that the delta is
an absolute tensor.
CHAPTER X

QUATERNIONS

181. Quaternion Algebra. The problem of extending 3-dimen-


sional vector algebra to include multiplication and division was
first solved by Sir William Rolvan Hamilton in 1843. He found
that it was necessary to invent an algebra for quadruples of num-
bers, or quaternions, before a serviceable algebra for number triples,
or vectors, was possible. Without attempting to motivate Ham-
ilton's invention, we proceed to a brief account of quaternion
algebra.
A real quaternion is a quadruple of real numbers written in a def-
inite order. We shall designate quaternions by single letters,
p, q, r; thus q = (d, a, b, c), q' = d', a', b', c'). The fundamental
definitions are the following.
Equality: q = q' when and only when d = d', a = a', b = b',
c=c'.
Addition:
(1) q+q' = (d+d',a+a',b+b') c+c').
Multiplication by a Scalar X:
(2) Xq = (ad, Xa, Xb, Xc).
Negative: -q = (-1)q.
Subtraction: q - q' = q + (- q'). Hence
q - q' = (d-d',a-a',b-b',c-c').
The zero quaternion (0, 0, 0, 0) is denoted simply by 0.
From these definitions it is obvious that, as far as addition, sub-
traction, and multiplication by scalars are concerned, quaternions
obey the rules of ordinary algebra:
(3) p + q = q+p, (p+q) +r = p+ (q+r);
(4) Xq = qX, (Xu)q = X(µq);
(5) (X + µ)q = Xq +µq, X(p + q) = Xp + Xq
403
404 QUATERNIONS § isi
In order to define the product qq' of two quaternions in a con-
venient manner, we shall denote the four quaternion units as fol-
lows:
1 = (1, 0, 0, 0), i = (0, 1, 0, 0), j = (0, 0, 1, 0), k = (0, 0, 0, 1).
Then, in view of the preceding definitions, we can write any qua-
ternion in the form:
(6) q = (d, a, b, c) = dl + ai + bj + ck.
Definition of Multiplication: The quaternion product,
qq' = (dl + ai + bj + ck) (d'l +a'i+b'j+c'k),
is obtained by distributing the terms on the right as in ordinary
algebra, except that the order of the units must be preserved, and
then replacing each product of units by the quantity given in the
following multiplication table:
1 i j k

(7) First
1 i j k

factor i -1 k -j
j -k -1 i
k j -i -1
Note that i2 = j2 = k2 = -1, and the cyclic symmetry of the
equations :
ij = k, jk = i, ki = j; ji = - k, kj = - i, ik = -j.
With this definition we find that
(8) qq'=dd'-aa'-bb'-cc'
+d(a'i+b'j+c'k)+d'(ai+bj+ck)
Ii j k
+ a b c

a' b' c'

If we form q'q by interchanging primed and unprimed letters, the


first two lines above remain unchanged, but the interchange of
rows in the determinant is equivalent to changing its sign; hence
§ 181 QUATERNION ALGEBRA 4U3

q'q = qq' only when the determinant is zero. That q'q and qq' differ
in general was to be expected; for, from the table, ij = k, ji = - k.
The table shows that multiplying a unit by 1 leaves it unchanged ;
hence (dl)q = q(dl) = dq, (dl)(d'l) = dd'l; and, from (1), dl +
d'1 = (d + d')1. Quaternions of the form dl therefore behave
exactly like real scalars and may be identified with them. Hence-
forth we shall write dl or (d, 0, 0, 0) simply as d; in particular,
1 or (1, 0, 0, 0) is regarded as the real unit 1.
We also may identify i, j, k with a dextral set of orthogonal unit
vectors. For, if we make the orthogonal transformation,
i = c11i + e12j + c13k,
c21i + c22j + c23k,
I6 = c3ii + c32i + c33k,
we have the relations cirri, = S,j (149.3) and
i2 = - Clr Clr = - 1,

li j k
3k = C2. C3r + C21 C22 C23 = c11i + c12j + C13k = i,
C31 C32 C33

since the elements of det cii (= 1) are their own cofactors (§ 149).
Thus every quaternion q = d + ai + bj + ck is the sum of a
scalar d and vector v = ai + bj + ck. With Hamilton we use the
symbols Sq and Vq to denote the scalar and vector parts of q; thus
(9) Sq = d, Vq = ai + bj + ck, q = Sq + V q.
The operations S and V are evidently distributive with respect to
addition.
We are now in position to prove the fundamental
THEOREM. Quaternion multiplication is associative and distribu-
tive with respect to addition; but the commutative law pq = qp holds
only when one factor is a scalar, or the vector parts of both factors are
proportional. In symbols:
(pq)r = p(qr);
p(q + r) = pq + pr, (p + q)r = pr + qr;
(12) pq = qp only when Vp = 0, or Vq = 0, or Vp = XVq.
406 QUATERNIONS § 182

Proof of (10). It will suffice to verify (10) for all possible com-
binations of the units i, j, k. Since the multiplication table is un-
changed under a cyclical permution of i, j, k, we need examine (10)
only for those products whose left factor is i; thus we find

Proof of (11). If we form p(q + r) and pq + pr by formal ex-


pansion and without use of the multiplication table, the two ex-
pressions will agree term for term; hence they will agree also after
the table is used.
Proof of (12). We have seen that q'q = qq' only when the de.
terminant in (8) vanishes. This occurs only in the three cases:

a=b=c=0; a'=b'=c'=0; a'a b'b c'c -_-_-;


that is, either q or q' must be scalar, or Vq and Vq' must be pro-
portional.
From (8), we have
(13) S(qq') = S(q'q),
for both scalars equal dd' - aa' - bb' - cc'. From (13), we may
prove that, the value of the scalar part of a quaternion product is not
changed by a cyclical permutation of its factors. We have, for ex-
ample,
S(p . qr) = S(qr . p) = S(q . rp) = S(rp - q),
and hence
(14) S(pqr) = S(qrp) = S(ppq).
182. Conjugate and Norm. The conjugate of a quaternion q,
written Kq, is defined as
(1) Kq = Sq - Vq.
§ 182 CONJUGATE AND NORM 407

The conjugate of a sum of quaternions is evidently the sum of


their conjugates:
(2) K(q + q') = Kq + Kq'.
Since the vector parts of q and Kq differ only in sign, q(Kq) _
(Kq) q. This product is known as the norm of q and is written Nq.
If
q=d+ai+bj+ck, Kq=d-ai-bj-ck,
we have, from (181.8),
(3) Nq = q(Kq) = (Kq)q = d2 + a2 + b2 + c2.
Therefore Nq is a scalar; and Nq = 0 implies that a = b = c =
d = 0, that is, q = 0. If Nq = 1, q is called a unit quaternion.
If v = ai + bj + ck, we have, from (181.8),
i j k
(4) vv' (aa' + bb' + cc') + a b c

a' b' c'

Since changing the sign of the determinant is equivalent to inter-


changing the second and third rows,
(5) K(vv') = v'v.
On taking the conjugate of every term in
qq' _ (d+v)(d'+v') = dd' + dv' + d'v + vv',
we see that
(6) K(qq') = dd' - dv' - d'v + vv = (d' - v') (d - v) = (Kq') (Kq)
Therefore the conjugate of the product of two quaternions is equal to
the product of their conjugates taken in reverse order. Since Kv =
- v, (5) is a special case of (6).
We now use this property to compute the norm of a product.
From (3), we have
N(pq) = pq - K(pq) = pq . Kq . Kp = p . Nq . Kp = pKp - Nq,
since Nq is a scalar and therefore commutes with Kp; hence
(7) N(pq) = Np Nq.
The norm of the product of two quaternions is equal to the product of
their norms.
408 QUATERNIONS § 182

By mathematical induction we immediately may extend (6) and


(7) to products of n quaternion factors:
(8) K(gig2 ... TO = Kgn . Kg.-1 ... Kq1,
(9) N(g1g2 ... qn) = Nq1 . Nq2 ... Nqn.'
From (6), we conclude that the product of two quaternions is zero
only when one factor is zero. Thus, if pq = 0, Np - Nq = 0; and,
since the norms are scalars, Np = 0 or Nq = 0, whence p = 0 or
q = 0.
We now can appreciate Hamilton's exceptionally happy choice
of a multiplication table for the quaternion units. t For quater-
nion algebra has a unique place among the algebras of hyper-
numbers,
x = x1e1 + x2e2 + ... + xnen,
linear in the units ei, with coefficients xi in the field of real num-
bers, and for which the associative law of multiplication holds
good. For it can be shown $ that the most general linear associa-
tive algebra over the field of reals, in which a product is zero only when
one factor is zero, is the algebra of real quaternions.
Quaternions include the real numbers (x, 0, 0, 0) with a single
unit 1, and the complex numbers (x, y, 0, 0) with two units 1, i.
Both real and complex numbers form a field, that is, a set of num-
bers in which the sum, difference, product, and quotient (the
divisor not being zero) of two numbers of the set must be definite
numbers belonging to the set. Moreover, quaternions include
vectors (0, x, y, z) in space of three dimensions. But (4) shows
that the product of two vectors is not in general a vector, but a quater-
* This theorem applied to qq' gives Euler's famous identity:
(d,2 + a,2 + b12 + c,2)
(d2 + a2 + b2 + c2)
(dd'-as'-bb'-cc')2+(ad'+da'+bc'-cb')2+
(bd' + db' + ca' - ac')2 + (cd' + dc' + ab' - ba')2.
t As to the genesis of quaternions, Hamilton himself has written:
"They started into life, or light, full grown, on the 16th of October, 1843,
as I was walking with Lady Hamilton to Dublin, and came up to Brougham
Bridge, which my boys have since called the Quaternion Bridge. That is to
say, I then and there felt the galvanic circuit of thought close; and the sparks
which fell from it were the fundamental equations in i, j, k; exactly such as I
have used them ever since."
t Dickson, Linear Algebras, London, 1914, p. 10.
§ 183 DIVISION OF QUATERNIONS 409

nion. Unlike addition-the sum of two vectors is always a vector


-the multiplication of vectors leads outside of their domain.
Vector multiplication is not closed, and, consequently, a pure vector
algebra having all the desirable properties of quaternion algebra
is not possible.
183. Division of Quaternions. If q is not zero, Nq is a non-zero
scalar; and we may write the defining equation for the norm (182.3)
as
Kq
q = 1, or qq = 1.

We therefore define Kq/Nq as the reciprocal of q and write

(1)
q-1
=
-
Kq
Nq
; then qq-1

The last equations show that Nq - Nq-1 = 1, or


= q-'q = 1.

(2) Nq-1 = -
Nq
In order to divide p by q ( 0), we must solve the equation
(3) (4) rq = p or qr = p
for r. This is easily done by multiplying by q-1 on the right in
(3), on the left in (4). We thus obtain the two solutions,
(5)(6) r1 = pq-1, r2 = q-1p,
which are in general different. For this reason the symmetrical
notation p/q will not be used. The notations (5), (6) are unam-
biguous: r1 satisfies (3) and may be called the left-hand quotient
of p by q; and r2, the right-hand quotient, satisfies (4). These
solutions are unique; if, for example,
rq = r1q,(r - r1)q = 0,
and, since q 0 0, r - r1 = 0, or r = r1.
On taking norms in (5) and (6), we have
Np
(7) Nrl = Nr2 =
Nq
The norm of either quotient of two quaternions is equal to the quotient
of their norms.
410 QUATERNIONS § 184

From (182.8) and (182.9),

(8) (gig2 ... qn.)-i = N(gig2 ... q) . qn 11 ... qi 1.


nn = qn i

The reciprocal of the product of n quaternions is equal to the product


of their reciprocals taken in reverse order.
The definition (1) shows that the reciprocal of a unit quaternion
is its conjugate; the reciprocal of a unit vector is its negative.
Example. Solve the equations (3) and (4) when
p=1+3i-j+k, q=2-i-2k.
From (1), we have q-' = (2 + i + 2k)/9; hence
ri=pq 1=1(1+3i-j+k)(2+i+2k) = 1(-3 + 5i - 7j + 5k);
r2 =q 1p = 1(2+i+2k)(1 +3i - j+k) = y(-3+9i+3j+3k).
Note that Sri = Sr2, Nri = Nr2 in conformity with (181.13) and (182.7).
.84. Product of Vectors. The product of two vectors vv' is the
,

quaternion (182.4) whose scalar and vector parts are


(1) S(vv') = - (aa' + bb' + cc'),
i j k
(2) V (vv') = a b c

a' b' c'


In order to find their geometric meaning we adopt a special basis
i, j, k. Choose i as the unit vector along
v' v, j as the unit vector perpendicular to i
v in the plane of v, v' and so directed that
° the angle (j, v') is not greater than 900
k (Fi g. 184 ). Th en k i s the un it vec t or
Fia. 184 which completes the right-handed orthogo-
nal basis i, j, k. If the angle (v, v') = 0, we have
v vli, v' = Iv'I(icos0+jsin0),
vv' = vIv'I(-cos0+ksin0);
(3) S(vv') = - I v I v' I cos 0,
I

(4) V(vv') = IvI v' sin 8 k.


Noting that k is the unit vector perpendicular to the plane of v
and v' and directed so that v, v', k form a right-handed set, we see
§ 184 PRODUCT OF VECTORS 411

that (3) and (4) are geometric expressions for the scalar and vec-
tor parts of vv', entirely independent of the basis i, j, k. They
form the cornerstones of the vector algebra of J. Willard Gibbs,
who defined the scalar and vector product of two vectors u, v as
(5)(6) u- v=-S(uv), u X v = V(uv),
using the dot and cross to distinguish between these two types of
"multiplication." Henceforth we shall use Suv, Vuv, Kuv to de-
note S(uv), V(uv), K(uv); similarly Suvw = S(uvw), etc.
A change in the order of the vectors in (3) has no effect, but in
(4) reverses the direction of k; hence
(7) (8) Svu = Suv, Vvu = - Vuv.
From u(v + w) = uv + uw and the distributive character of S
and V,
(9) (10) Su(v + w) = Suv + Suw, Vu(v + w) = Vuv + Vuw.
Since Kuv = (- v) (- u) = vu,
uv = Suv + Vuv, vu = Suv - V UV;
(11)(12) Suv = 2(uv + vu), Vuv = 1(uv - vu).
Turning now to products of three vectors, we have
(13) Suvw = Svwu = Swuv (181.14).
Since Kuvw = (-w)(-v)(--u) = -wvu (§ 182),
(14) Suvw = SKuvw = -Swvu
(15) Vuvw = - VKuvw = Vwvu.
Also from uvw = u(Svw + Vvw),
(16) Suvw = SuVvw,
(17) Vuvw = uSvw + VKVvw.
We now compute Vuvw in another way:
2Vuvw = uvw - Kuvw
= uvw + wvu
= uvw + vuw - vuw - vwu + vwu + wvu
= (uv + vu)w - v(uw + wu) + (vw + wv)u;
412 QUATERNIONS § 185

hence, from (11),


(18) V uvw = uSvw + wSuv - vSuw.
Comparison with (17) now gives the important formula:
(19) VuVvw = wSuv - vSuw.
Finally, we express these results in the dot and cross notation
of Gibbs:
(5)(6) Jul Ivl cos(u,v), uXv = Jul lvlsin(u,v)k;
(7)(8) v, vxu= -u-V;
(9)

(10) u.(v+w) =uXv+uxw;


(13)
(14)
(19) uX(vXw) =
From (5) and (6) we have also
(20) uv =
as the connecting link between quaternion and vector algebra.
185. Roots of a Quaternion.t Every quaternion
q = d + ai + bj + ck
with real coefficients may be written as a real multiple of a unit
quaternion :
(1) q = h(cos 0 + e sin B), 0 < 0 < 27r.
Here
h= d2+a2+b2+c2,
cos B = d/h, sin 9 = f \/a2 + b2 + c2/h
and, when a2 + b2 + c2 0, e is the unit vector:

(2) e = f ai+bj±ck
1/a2+b2+c2
t In this and following articles we again denote vectors with bold-face
letters; but ab denotes the quaternion (not dyadic) product.
§ I35 ROOTS OF A 413

When q is a real number, sin 0 = 0, and e may be chosen at pleas-


ure. Since e2 = - 1, we have, by De'Xloivre's Theorem,
(3) q' = h"(cos no + e sin no).
We now may find the nth roots of a real quaternion
(4) Q=H p -I- a sin gyp), 0 < (P < ,r;
the angle always may be taken in the interval from 0 to it by
choosing the appropriate e in (2). In solving q' = Q, we consider
two cases:
1. sin 0; we choose the e in q the same as in Q. Then
hn = H, cos no = cos cp, sin no = sin gyp,

and n nth roots of Q are given by (1), provided


(5) h = H' I', the positive root,
(6) 0 = (gyp + 2irm)/n (m = 0, 1, ,n- 1).

These n values of 0 comprise all values in the interval 0 < 0 < 2ir
which satisfy the preceding equations.
2. sin p = 0: the e in q is then an arbitrary unit vector.
IfQ>0: o =0, 0 =2mir/n (m=0,1, ,n-1).
When n = 2, the values 0 = 0, Tr give just two roots =LA/Q-, both
real. When n > 2, some values of 0 (0 0 or ir) give non-real roots
q with which any e may be associated.
IfQ<0: p =ir, 0=(2m+1)ir/n (n=0,1, ,n-1).
In every case some values of 0 (0 ir) give non-real roots q with
which any e may be associated.
We summarize these results in the
THEOREM. A quaternion with real coefficients, but not a real num-
ber, has exactly n nth roots. If Q is a positive real number, it has
just two square roots ± /; in all other cases a real number has
infinitely many quaternion roots with real coefficients.
In all cases the roots may be computed from (5) and (6). For
example, if Q = 1 + i + j + k, we write
Q = 2(cos 60° + e sin 60°), e = (i + j + k)/1'3.
414 QUATERNIONS §186

The cube roots of Q are then


q= (cos 0 + e sin 0), 0 = 20',140',260'.
186. Great Circle Arcs. Every unit quaternion
q=d+ai+bj+ck (Nq=1)
can be expressed in the form
(1) q = cos 0 + e sin 0.
Here e is given by (185.2); and 0 satisfies
(2) cos 0 = d, sine = a2 + b2 + c2.
If we choose the plus sign in these formulas, 0 < 0 < 7r. In par-
ticular, if q = 1, -1, e, the angle 0 = 0, zr, 2 7r, respectively.
THEOREM. The unit quaternion cos 0 + e sin 0 may be expressed
as the quotient ba-1 of any two vectors which satisfy the conditions:
(i)lal=lbl,
(ii) Angle (a, b) = 0,
(iii) Plane a, b is perpendicular to e,
(iv) a, b, e form a dextral set.
Proof. In view of (i), we may write
Ial bIcos0+Ial bIsin0e
cos9+esin9=
IaI2
hence, if we choose the vectors a and b so that conditions (ii),
(iii) and (iv) are fulfilled,
-Sba - Vba
cosh+esin0=
-Sab + Vab
Na
_
Na
= --=b-
ba
Na
Ka
Na
or, in view of (183.1),
(3) cos 0 + e sin 0 = ba-1.
From Fig. 186 we see that, when a, b, e form a right-handed
set, the angle (a, b), when less than 7r, is counterclockwise, viewed
from the tip of e. We shall describe the sense of (a, b) as positive
relative to e. When the angle (a, b) = 0 or 7r, q is 1 or -1, re-
spectively, and e is entirely arbitrary.
§.,% G'ItEAT CIRCLE ARCS 415

To every unit quaternion, q = cos 0 + e sin 0 corresponds to a


great circle are AB of a sphere centered at 0, provided OA = a
and OB = b satisfy the preceding conditions (i) through (iv).
Thus q corresponds to an are of a great circle whose plane is nor-
mal to e and whose central angle 0
has the positive sense relative to e.
All such arcs of this great circle are
equally valid representations. If the
are AB represents q, AB is free to
move about in its great circle, provided
its length and sense remain unaltered. q
The cases q = 1 (B = 0) and FIG. 186
q = -1 (0 = -7r), in which e is
arbitrary, are exceptional. Any point of the sphere represents
q = 1; and any great semicircle represents q = - 1. A unit vec-
tor q = e (0 = 2 r) corresponds to a quadrantal are in the plane
through 0 normal to e.
If q = cos 0 + e sin 0 corresponds to the are AB,
(4) qI = Kq = cos 6 - e sin B
corresponds to an are having the same plane and angle, but whose
sense is positive relative to -e. Hence q-' corresponds to the
are BA. Moreover,
(5) - q = - cos 0 - e sin 0 = cos (ir - 0) - e sin (7r - 0) ;
hence, if q corresponds to the are AB, and AOA' is a diameter,
-q corresponds to are A'B, the supplementary are reversed in
Sense.
Using the sign '' to denote correspondence, we sum up our find-
ings as follows:
1 ti point, -1 - semicircle, e quart.ercircle;
q - are AB, q-I - are BA, -q '' are A'B.
The utility of this representation is due to a simple analytical
method of adding great circle arcs "vectorially." To add two arcs,
shift them along their great circles until the terminal point of the first
(AB) coincides with the initial point of the second (BC); then the
great circle arc AC is defined as their vector sum. If a = OA,
416 QUATERNIONS §186

b = OB, c = OC, the preceding theorem shows that the arcs AB,
BC, AC are represented by the quaternions ba-1, cb-1, ca-1, re-
spectively. Write
p = ba-1, q = cb-1; then ca-1
= cb-lba-1
= qp;
and the equation,
are AB + are BC = arc AC,
may be written
(6) are p + arc q = are qp.

For three arcs,


(7) arc p + are q + are r = arc qp + are r = arc rqp;
and, in general, the vector sum of any number of great circle arcs is
given by the arc corresponding to the product of their representative
quaternions taken in reverse order.
In interpreting such arc-quaternion equations, remember that
are q = 0 means that q = 1; and, if are q is any great semicircle,
q = -1. For example, if arc p, are q, are r form the sides of a
spherical triangle taken in circuital order,
arc p + arc q + arc r = 0, arc rpq = 0, rpq = 1.
In general, the arcs representing the quaternions q1, q2, , qn,
taken in this circuital order, will form a closed spherical polygon
when and only when
(8) qnqn-1 ' g2g1 = 1.
Example. Spherical Trigonometry. Consider again the spherical triangle
ABC of § 22. With the notation of this article
are BC ' cb-1 = cos a + a' sin a,
are CA ' ac-1 = cos $ + b' sin Q,
are AB - ba 1 = cos y + c' sin y.
The "vector" equation
are CA + arc A B = arc CB
corresponds to the quaternion equation (bat) (ac 1) = be 1, or
(i) (cosy + c' sin y) (cos i3 + b' sin 3) = cos a - a' sin a.
§ 187 ROTATIONS 417

If we expand the left member, put


c'b' = - b' c' - b' x c' cos a' - a sin a';
and equate scalar parts in both members, we have
(ii) cos 0 Cos y - sin 3 sin y cos a' = cos a.
This is the cosine law (22.6) of spherical trigonometry. On equating the vector
parts in both members of (i) we have,
(iii) a' sin a + b' sing cos y + c' cos l; sin y = a sin a' sin 0 sin y
and hence, on multiplication by a ,

sin a' sin /i sin y = a a' sin a = a bxc,


or
sin a' a- bxc
sina sin a sin#sin y
Since the right member is unchanged by a cyclical permutation, we have
sin a' sin /3' sin y'
(iv)
sina sing sin y
the sine law (22.5) of spherical trigonometry.
187. Rotations. With the aid of quaternion algebra, finite rota-
tions in space may be dealt with in a simple and elegant manner.
This application depends upon the fundamental
THEOREM. If q and r are any non-scalar quaternions, then
(1) r' = qrq-1
is a quaternion whose norm and scalar are the same as for r. The
vector Vr' is obtained by revolving Vr conically about Vq through
twice the angle of q. Thus if
q = Nq(cos 0 + i sin 6),
Vr' is obtained by revolving Vr conically about i through an angle 26.
Proof. The norm and scalar of r' are
(2) N(qrq-1) = Nq Nr Nq-1
= Nr (182.9),
(3) S(qrq-1) = S(q 'qr) = Sr. (181.14).
Moreover on writing r = Sr + Vr,
qrq-1
= Sr + q(Vr)q-1;
418 QUATERNIONS § 187

Now, from (3), q(Vr)q-1 has


the same scalar as Vr and is there-
fore a vector; hence
(4) V (qrq-1) = q(Vr)q-1
Let us now write
r = Nr(cos c + e sin (p);
then from (4),
Vr' = Nr sin e' where e' = qeq-1.
If we choose j in the plane of e and i
(Fig. 187a) and k to complete the dextral
set i, j, k, then
e = icosA+ Jsin A
e' = (qiq-1) cos A + (qjq 1) sin A.
Since Vq is parallel to i, qi = iq and
Flo. 187a
qiq--1
= iqq-1
= i.
Moreover,
qjq-1
= (cos 0 + i sin 0)j(cos 0 - i sin 0)
= (j cos 0 + k sin 0) (cos 0 - i sin 0)
= j (cost 0 - sine 0) + k(2 sin 0 cos 0) ;
hence j goes into
j' = j cos 20 + k sin 20,
a vector obtained by revolving j about i through an angle 20 in
the positive sense. Consequently
e = icosA+ jsinX -*e' = icosA + j'sinA
and also Vr -. Vr' by a conical revolution about i of the same
amount. This completes the proof.
We note that vectors transform into vectors. In particular, if
q = a, a unit vector,. 0 = 90°, and
e' = aea-1 = -aea
is obtained by revolving e conically through 180° about a. The
transformation -a( )a thus gives vectors a half-turn about a.
§ 187 ROTATIONS 419

The transformation a( )a may be regarded as a half-turn fol-


lowed by a reversal; a vector thus transformed is simply reflected
in a plane normal to a (Fig. 187b). Thus aea is the reflection of e
in the plane normal to a.
A rotation through an angle a about a a
implies that the angle turned has the positive
sense relative to a. If
p = Cos. a+asin1a,
q= cos 2 S -{- b sin 3
z
are unit quaternions, the operators p( )p-1 Fia. 187b
and q( )q-1 effect rotations of a about a and
,3 about b; for brevity we call these the rotations p and q. The
succession of rotations p, q corresponds to the operator,
)p-'q-1
qp( = qp( ) (qp)-1;
since qp is also a unit quaternion, say
qp = cos zy + c sin 21
'Y,

the resultant is equivalent to the single rotation qp, that is, a ro-
tation through the angle y about c. Similarly, the resultant of
the rotations q and p corresponds to the operator,
)q-'p-1 = pq( )(pq)-1
pq(
Since pq = qp only when Vp and Vq are parallel (the values p, q
= -4-1 are excluded), the composition of rotations is non-commu-
tative except when they have the same axis.
The rotation p followed by the rotation q is equivalent to the single
rotation qp. More generally, the succession of rotations q1, q2,
qn is equivalent to the single rotation gnq,,-1 . g2g1.
Since (-q)-1
= -q-1, the rotations q( )q-1 and (-q)( )(-q)
are the same. If q = cos B + a sin B,
-q = cos(Tr-9)+(-e)sin(7r-8);
thus the rotation -q is a rotation through 27r - 20 about -e;
this produces the same result as the rotation q, namely 20
about e.
Since q-lq( )q-lq = 1( )1, the rotation q-1( )q is the reverse of
q( )q-1; this is also evident from q-1 = cos 0 - e sin 0.
420 QUATERNIONS § 187

Example 1. The rotation of 90° about j followed by a rotation of 90° about


i is represented by the quaternion product
(cos 45° + i sin 45°) (cos 45° + j sin 45°) = !(I + i + j + k);
that is, by
i+j+kV' i + j + k
1
+ - = cos 60° + - sin 60°.

2 V3 2 -\/3
The resultant rotation is therefore a rotation of 120° about an axis equally
inclined to the (positive) axes of x, y and z.
Example 2. The resultant of two reflections in planes normal to the unit
vectors a, b corresponds to the operator
ba( )ab = ba( ) (ba)-1.
But if a b = cos 0, a x b = e sin 0, we have
ba = -a b - axb = -(cosB+esin0) (184.20).
Now the rotation q( )q-' = (-q)( )(-q)-1; hence successive reflections in
two plane mirrors is equivalent to a rotation about their line of intersection
of double the angle between them.
Example 3. From (181.13) we know that S(qp) = S(pq); moreover from (4)
V (qp) = V (gpgq-1) = qV (pq)q-1
Hence V(qp) is obtained by revolving V(pq) about Vq through double the
angle of q. Thus, if u is a vector, V(up) is obtained by revolving V(pu) 180°
about u; that is, the vector u bisects the angle between V(pu) and V(up).
Now let a, b, c be three radial vectors from the center of a sphere to its
surface.Then a, b, c bisect the angles between Vabc, Vbca; Vbca, Vcab;
Vcab, Vabc respectively. In other words, if we form a spherical triangle
whose vertices are Vabc, Vbca, Vcab, the middle points of the sides opposite
lie on the vectors b, c, a respectively.
Example 4. If the sides of a spherical polygon are represented by the
quaternions qi, q2, , qn taken in this circuital order, gnqn-1 g2gt = 1
(186.8). Hence the succession of rotations,
gnqn-1 ... q2ql( )gig2 ... q.-1qn = 1( )1,
about axes through a point 0 will restore a body to its original position. We
state this result for the case of a triangle as follows:
THEOREM (Hamilton and 1)onkin). If ABC is any spherical triangle, three
successive rotations represented by the directed arcs 2 BC, 2 CA, 2 AB (about their
polar axes) will restore a body to its original position.
This same theorem applies to the polar triangle A'B'C'. Since the side
B'C' = a' = 7r - A (§ 22) and has OA as polar axis, successive rotations of
2ir - 2A, 27r - 2B, 2ir - 2C about OA, OB, OC will restore a body to its
original position. Since the rotations 2ir - 2A and -2A about OA give the
same displacement, we may state the
§ 188 PLANE VECTOR ANALYSIS 421

THEOREM (Hamilton). If ABC is any spherical triangle on a sphere centered


at 0, three successive rotations about OA, OB, OC through the angles 2A, 2B, 2C
in the sense of CBA will restore a body to its original position.

188. Plane Vector Analysis. The three-term quaternion c + ai


+ bj has given rise to two types of vector analysis in the plane.
The one interprets c + ai as a vector w in the complex plane; then
the product wlw2 is always a complex vector. The other inter-
prets ai + bj as a "real" vector w and decomposes the quaternion
product w1w2 into its scalar and vector parts, which are used
separately as "products."
To indicate the interpretation used, we denote complex and real
vectors by italic and bold-face letters, respectively. Thus the
plane vector whose components are u, v may be written as
w = u + iv, or w=ui+vj.
In the first case,
(1) w1w2 = (ulna - vlv2) + (ulv2 + u2v0i;
in the second,
(2) w1w2 = - (ulu2 + vlv2) + (u1v2 - u2v1)k.
In Gibbs's notation,
ulu2 + V1V2 = w1 w2i u1v2 - u2v1 = k w1 x W2.
Let w = u - vi denote the conjugate of w; then, from (1) and (2),
w1W2 = W1 W2.+ik.wlxw2,
W1W2 = W1 W2 - i k W1 x w2,
and hence

(3) W1 W2 = 2 (wlw2 + W1w2),


(4) 21i(w1w2 - wlw2)
The conditions (128.6) and (128.7) for perpendicular and parallel
vectors may be read from these equations.
We next consider corresponding differential invariants. The
operator,
V =i-+j--+i-
a
ax
a a
ay
a
ax ay
422 QUATERNIONS §188

(read - as corresponds to). If we introduce the conjugate variables,


z=x+iy, 2=x-iy,
-+i- _ -+i-ay az-+(ax-+i-ay a2-
a
ax
a
ay
az
ax
az a az az a

a a
(1 + i2) + (1 - i2)
az a2,
or
a a _ 0 a 8 _ a
(5)
ax + Z y a- 2 az ' ax - Z ay
2
az

follows in the same way.


Corresponding to the gradient V of a real function p(x, y), we
have 2 a(p/az in the complex plane, in which the variables x, y in
are replaced by the values,
y = 2i(2 - z).
x = 2 (2 + z),
For example, x2 + y2 = z2, and hence 0(x2 + y2) = 2z in the
complex plane.
The unit vector,
(6) e = icos0+jsin8NeiO = cos0+isin9.
In view of (3), the operator for differentiation in the direction e,
namely
(7)

Therefore
(8)
dw
- = e Vw -e t8 aw + e-i0aw
ds
-,
az
-
az

and hence dz/ds ti eie. If w is a complex function, dw/dz in the


direction 0 corresponds to the ratio of dw/ds to dz/ds; hence
dw aw _2. aw
e in the direction 0.
(9) dz az + az
When dw/dz is independent of 0, w is said to be an analytic func,
tion of z; for this, it is necessary and sufficient that
aw
az
=0, or -+i-
a

ax
(u+iv)=0,
ay
a
§ 188 PLANE VECTOR ANALYSIS 423

in view of (5). The last condition is equivalent to the familiar


Cauchy-Riemann Equations :

(10)
au
ax
av
ay
=0, -+-=0.
au
ay
av
ax

We next find the correspondents for div w = V w and k rot w


= k V x w by making use of (3), (4) and (5) :
3z'
divwti- +aw-, a2 az

(12) k rot w /aw


\az
aw
tii(---
az

Moreover, for the Laplacian V2 = V V, we have


a2 a2 a2
(13) V2=2
azaz+2azaz - 4azaz
Thus a real function cp(z, 2) is a harmonic when a2,p/3z ai = 0; for
example,
log I z I = 2 log zz = (log z + log z)
2
is harmonic.
If the plane vector w - w(z, 2), the condition,
aw
(14) rotw=ON---=0.
Ow

az az
When rot w = 0,
w = VX, where a= w dr;
:o
hence, when w(z, 2) is irrotational,

(15) w= z , where p = f (w dz + w dz)

isthe real function 2X(z, 2). The field lines cut the curves
const at right angles.
If the vector w is plane, the condition,
Ow aw
(16) div w = 0 - - + - = 0.
az a2
424 QUATERNIONS § 188

From (85.6), rot (k x w) = k div w; hence, when w is solenoidal,


k X w is irrotational. Since k X w is w revolved through +7r/2,
k x w - iw. Thus when w(z, 2) is solenoidal, (15) applies when w
is replaced by iw (and iv by -iw). From this result, we conclude
that
(17) w=i where = ifZ(w dz - w d2)
az zo

is a real function. The field lines are the curves = const.


From these results we have
THEOREM 1. If <p(z, 2) is a real function with continuous partial
derivatives, app/a2 and i app/a2 give, respectively, an irrotational field
with lines orthogonal to cp = const, and a solenoidal field with lines
= const.
The preceding results give a simple method for decomposing a
plane vector function into an irrotational and a solenoidal part.
For, if
w = W1 + w2i rot W1 = 0, div W2 = 0,
there exist real functions gyp,' such that

w=-+i-=-(
a,
app

a2
o +ii');
a2
a

az

' + i¢ = Jw(z, 2) d2 (z const)

is determined to an arbitrary additive f (z) and w1 = app/az, w2 =


i ay/az.
If the vector field w(z, 2) is both irrotational and solenoidal,
rot w = 0 implies that
az p
w=-; az
then =w,asp
az
divw=2 dz dz
=0.
Hence p is a real harmonic function, and w is an analytic function
of z. Conversely, if w is an analytic function of z, (9w/a2 = 0;
hence aw/az = 0, and (11) and (12) show that w is solenoidal and
irrotational.
THEOREM 2. In order that the complex vector w(z, 2) be irrotational
and solenoidal, it is necessary and sufficient that its conjugate be an
analytic function of z.
§ 188 PLANE VECTOR. ANALYSIS 425

Example 1. The vector


(x2 - y2) - 2xy i,
w = 22 =
is both irrotational and solenoidal; for its conjugate z2 is an analytic function
of Z.

Example 2. To decompose the vector,

w = z2 = (x2 - y2) + 2xy i,


into its irrotational and solenoidal parts, we may take

v + ik = fz2 d2 = z22.
Since ,P - i= 22Z,
= 12z2(2 + z), ¢ = Zi z2(2 - z);
av
wl _ = z2 + 2z2 = (3x2 + y2) + xy it
a2 2

ak
w2 = i = -z2 + 2z2 = - (x2 + 3y2) + xy i.
a2 i
Stokes' Theorem in the plane,

corresponds to

i - aw
492
fk . rot w dA =

- -aw
az
dA = a
fw . dr,

(w d2 + w dz),

by virtue of (12) and (3). If we replace w by -iw (and iv by i4 b),


we obtain, after canceling i,

-+
i f(afv
az
-aw
az
dA = 2 J(-w dz + w dz).

On adding these equations and then replacing w by w, we have

(18) fw dz = 2if -aw


az
dA.
When w is analytic in the region within the circuit, aw/a2 = 0,
and (18) reduces to Cauchy's Integral Theorem: Jw dz = 0.
426 QUATERNIONS § 189

Example 3.

2iA =
When w = z in (18),

12 dz = J(x - iy) (dx + i dy) = i f (x dy - y dx).

This gives the well-known circuit integral for a plane area.


With w = z2, we obtain the static moments of a plane area about the axes,
expressed as circuit integrals.

189. Summary : Quaternion Algebra is a linear, four-unit


(1, i, j, k) associative algebra over the field of reals. The unit 1
has the properties of the real one; and
i2=j2=k2= -1, ij=k, ji= -k,
the last equations admitting cyclical permutations. Quaternions
q = d + ai + bj + ck include real and complex numbers (d,
d + ai). Since i, j, k may be interpreted as dextral set of orthog-
onal unit vectors, quaternions also include vectors v = ai + bj
+ ck. Thus q = d + v, a scalar plus a vector.
Quaternion multiplication is associative and distributive, but
not in general commutative; in fact pq = qp holds only when p
or q is a scalar or when the vector parts of p and q are propor-
tional.
The product vv' of two vectors is the quaternion :

vv' = - (aa' + bb' + cc') +

= -v - v' + v x v' in Gibbs' notation.


The quaternion q = d + v has the conjugate Kq = d - v. The
conjugate of the product qq' is K(qq') = (Kq') (Kq).
The norm of a non-zero quaternion q is the positive real number,
Nq = q(Kq) = d2 + a2 + b2 + c2;
and N(qq') = (Nq) (Nq'). The equation q = 0 implies Nq = 0,
and conversely; hence, if qq' = 0, either q = 0 or q' = 0.
A unit quaternion q (Nq = 1) may be put in the form
q = cos0+esin9 (IeI = 1),
and associated with the great circle are of angle 0 and pole e on
the unit sphere. On a fixed great circle, all arcs of the same length
PROBLEMS 427

and sense correspond to the same q and are denoted by arc q. The
scalars 1 and - 1 correspond, respectively, to any point and to any
great semicircle of the sphere.
Great circle arcs may be added vectorially; and
are p + are q = arc qp.
If three arcs form a spherical triangle, say-
arcp+arcq+arcr = 0, then rpq= 1.
)q-1
If q = Nq(cos 0 + e sin 0), the operator q( effects a conical
revolution of 20 about e on the vector of the operand; thus if
r' = qrq-1, then
Nr' = Nr, Sr' = Sr, T Y = Vr revolved about e through 20.
When q = e, a unit vector, q-1 = -e; the operator -e( )e gives
vectors a half-turn about the axis e.
The operator e( )e reflects vectors in the plane normal to e.

PROBLEMS
1. Solve the quaternion equations, rq = p, qs = p, for r and s when
q=2-i-2k, p= 1+3i-j+k.
Verify that Vr = Ns.
2. Every quaternion q satisfies the quadratic,
q2 - 2qSq + Nq = 0,
known as its principal equation. The conjugate Kq satisfies the same equation.
3. Show that 2 + 5i and 2 + 3j + 4k have the same principal equation;
hence factor its left member in two different ways.
4. Show that if we identify the quaternion units with the 2 X 2 matrices
j
1 - (0 1 ), ( 0 L
0)'

\-1 0), - \-c0 0/'


where L (iota) is the complex unit, L2 = -1, these matrices satisfy the multi-
plication table (181.7).
5. If u, v, w are vectors prove the following identities:
(a) u22 _ -Nu; (u - v)(u + v) = u'- + 2Vuv - v2;
(b) 2 Suvw = uvw - wvu;
(c) 2 Vuvw = uvw + wvu;
(d) S(u + v) (v + w) (w + u) = 2 Suvw.
428 QUATERNIONS
6. Show that the multiplication table of the units i, j, k is completely given
by
i2=j2=k2=ijk= -1.
7. Interpret the equations, ij = k, jk = i, ki = j and kji = 1, in terms of
"arc vectors."
8. Solve the equation aq + qb = c for the quaternion q if a, b, c are known
quaternions and Na 5x, Nb. [q = (ac - cb)/(Nb - Na).)
9. Solve the equation qa + bq = q2. [Reduce to aq-1 + q -1b = 1.]
10. If a, b, c are vectors for which Vabc = 0, prove that a, b, c are mutually
orthogonal.
[Vabc = aSbc + VaVbc and VaVbc 1 a; hence Sbc = 0, VaVbc = 0.1
11. If a, b, c, d are vectors for which Vabcd = 0, prove that
(a) bed is a vector parallel to a;
(b) Vbcda = Vcdab = Vdabc = 0;
(c) the vectors a, b, c, d are coplanar;
(d) cda is a vector parallel to b; and thus cyclically.
12. A body is revolved through 90 degrees about two axes e1, e2 which
intersect at an angle 0 (cos 0 = e1 e2). Show that the equivalent single
rotation is through an angle 2 cos' z (1 - cos 0) and about an axis parallel
to e1 + e2 - e1 x e2.
13. If a and b are unit vectors along intersecting axes, show that the half-
turns -a( )a, -b( )b in this order are equivalent to a rotation of twice the
angle from a to b about their common perpendicular.
14. The quaternion q = a1a2 a,,_1 an is the product of it unit vectors.
Show that if Sq = 0, q is the vector
q = ±anan-1 ... a2a1 (+ when it is odd, - when n is even).
Hence show that the successive reflections ai( )ai in the order i = 1, 2, 3,
, n reduce to the single reflection q( )q when n is odd, and to a half-turn
-q( )q when n is even.
15. If the successive reflections ai( )ai in the order i = 1, 2, 3,. , it re-
duce to a single reflection or half-turn, show that this is true for any cyclical
permutation of this order.
16. Show that the succession of reflections in three coaxial planes reduces
to a single reflection.
17. Prove the famous theorem of Euler (1776): Any displacement of a rigid
body which leaves the point 0 fixed is equivalent to a rotation about an axis through
0. [Let the displacement move the trihedral ijk in the body to the new posi-
tion ijj1k1. This may be accomplished by two reflections: the first takes
i -+ i1, j -s j'; the second takes j' - ji, leaving i1 undisturbed.]
18. If q = cos 0 + e sin 0 (not scalar), show that the rotation q( )q-1 fol-
lowed by the reflection a( )a reduces to a single reflection b( )b when and
only when a e = 0; moreover,
b = aq = a cos 0 + a x e sin 0.
19. If we define qn for arbitrary real n by (185.3), show that any quaternion
may be expressed as the power of a vector. When q = cos 0 + e sin 0, prove
PROBLEMS 429

that q = e2e/r; and that the r( ti tion of angle about the axis e is represented
by the operator e ' n ( )e -" Oi*.
20. Show that successive h ili-turns about three mutually orthogonal and
intersecting axes will restore a body to its original position.
21. ABC is a spherical triangle and P, Q are the middle points of the sides
AB, BC, respectively. Prove .hat two successive rotations represented by the
arcs AB, BC are equivalent tc the rotation represented by twice the great cir-
cle are PQ.
22. Prove the theorem: Th,ee successive rotations represented by the arcs
AB, BC, CA of a spherical triangle ABC are equivalent to a rotation about OA
through an angle equal to the spherical excess (A + B + C - 7r) of ABC.
23. Prove Rodriques' Cons'ruction for the composition of two rotations
through the angles a, f3 about he axes OA, OB:
Draw great circle arcs AC m.nd BC such that
angle (AB, AC) = -!a, angle (BA, BC) = 2
determining the spherical triangle ABC; then rotation a about OA followed by
0 about OB is equivalent to the rotation
-y = 2(CA, CB) about OC.
117se Hamilton's Theorem given in §187, Ex. 4.]
Give the construction when rotation 0 about OB is followed by rotation a
about OA. Draw a figure to illustrate both constructions.
24. Prove that successive rotations through angles of gyp, ,r/2, o about the
axes of x, y and z respectively are equivalent to a rotation of 7r/2 about the
y-axis.
25. If r = xi + yj + zk and r' = x'i + y'j + z'k, are position vectors and
q is an arbitrary quaternion, show that the transformation r' = qr(Kq) repre-
sents the most general rotation and expansion of 3-dimensional space. What
is the ratio of expansion?
If q = d + ai + bj + ck, express x', y', z' in terms of x, y, Z.
INDEX
The numbers refer to pages. A starred number locates the definition of the
term in question. The letter f after a number means "and following pages."
Terms under a noun (key word) are to be read in before this word, unless
preceded by a preposition such as "of" or "for." Terms under an adjective
(key word) are to be read in after this word; the repeated adjective is indicated
by dash.
Absolute tensors, 343f Antecedents of a dyadic, 136*
Acceleration, 109* Antisymmetric dyadic, 141 *
angular, 122* -tensor, 352*
four-vector, 401 Archimedes, Principle of, 258
general components of, 359 Arc vector, 415f
normal component of, 109 Area, of a polygon, 39
of Coriolis, 124* of a spherical triangle, 310
of gravity, 124 of a triangle, 44
rectangular components of, 110 vector, 37*f
relative, 123 Areas, Law of, 133
tangential component of, 109 Astatic center, 62
transfer, 123* Asymptotic lines, 285*, 288, 303
uniform, 110 Axes, rectangular, 26
Addition, of dyadics, 137* principal, of a dyadic, 162
of forces, 75 of inertia, 160
of great circle arcs, 415f of strain, 164
of motors, 67* of stress, 256
of quaternions, 403* Axis, 25*
of tensors, 350* instantaneous, of velocity, 117*
of vectors, 3* of acceleration, 127*
statical, 4* of a motor, 66*
Addition theorems (sine, cosine), 28 radical, 62
Adjoint dyadic, 152*
Af ine connection, 356*f, 361, 368 Barbier's Theorem, 134
-group, 340f, 386 Base vectors, 24, 147, 349, 367, 368
-transformation, 138 dextral and sinistral, 24, 25, 41
Alternative airport, 113 Basis, 25
Analytic function, of (real) coordi- Bernoulli's Theorem, 279
nates, 347 Bertrand, J., 132, 314
of complex variable, 272, 422* Bianchi identity, 385
Angular acceleration, 122* Bilateral surface, 218
-excess, 310 Binormal, 92*
-speed, 114* Bivector, 353*
-velocity, 115* B6cher, M., 339
431
432 INDEX
Bonnet, 0., 286, 288, 296, 309 Components of a dyadic, 166, 342
Bonnet's Integral Formula, 309 of stress, 256
Box product, 42* of a tensor, 344
Bracket notation for vectors, 25 of a vector, 26, 48
of velocity, 164, 359
Carnot's Theorem, 60 Composition of velocities, 120f
Cartesian coordinate system, 379 *f, Conditions, necessary and sufficient,
385f viii*
Catenary, 103, 317 Cone, 304, 315
Catenoid, 317 Conjugate, of a complex number 274,
Cauchy-Riemann Equations, 272, 422
423 of a dyadic, 141
Cauchy's integral, 265 of a quaternion, 406*
-Integral Theorem, 276, 425 Cc njugate lines, 79, 80
Caustic, 104 Conoid, 321 *, 323
Center, astatic, 62 Consequents of a dyadic, 136
instantaneous, 118* Continuity, Equation of, 258f, 264
mean, 20* Contraction, 351 *f
of curvature, 99* Covariant derivative, 362*f, 365f, 401
of mass, 258*, 236 of dyadics, 364
of normal curvature, 285* of epsilons, 366
Central forces, 133, 134 of Kronecker deltas, 366
Centrifugal force, 124 of metric tensor, 366
Centroid, 9* of scalars, 364, 366
Ceva, Theorem of, 14 of tensor product, 365
Characteristic equation, 154*, 168f of tensor sum, 365
-numbers, 154*, 156 of vectors, 364
Characteristics, 98*, 99, 133 Covariant vector, 335*
Christoffel symbols, 326, 362*, 389 Cross product, 34
Circle, nine-point, 33 Curl, of a tensor, 374*, 376
of curvature, 99 of a vector, 183*
rolling, 120 Curvature (K), 92*, 96, 97
Circular motion, 110 center of, 99*, 285
Circulation, 266*f circle of, 99
Clifford, W. K., 64 geodesic (y), 284*
Codazzi, Equations of, 300f, 391 integral, 309*f
Cofactor, 168*, 330*f lines of, 285*, 302f
reduced, 331 * mean (J), 222, 286*, 287, 292, 392
Cogredient transformations, 335 normal (k), 284*, 287
Collinear points, 8f of plane curves, 101
-vectors, 2* of surface curves, 283f
Complete quadrangle, 16 principal, 287*
-quadrilateral, 18 radius of (p), 93*
Components, contravariant, 335, 342, total (K), 286*, 287, 292, 306f, 392
344 Curvature tensor, 379, 380 *f
covariant, 335, 342, 344 Contragredient transformations, 334*
mixed, 342, 344 Contravariant components, of a
of acceleration, 164, 359 dyadic, 342
of affine connection, 356 of a tensor, 344
I \ DEX 433

Contravariant component;;, c a ve:-- Derivative, of a dyadic, 171


tor, 48, 335 of a motor, 126
Contravariant vector, 334i of a vector, 84f
Coordinates, barycentric-, '23* Derivative formulas, of Gauss, 327,
Cartesian, 25*, 340*, 335 390
curvilinear, 191 of Weingarten, 323, 327, 391
cylindrical, 195 Desargues' Theorem, 15
geodesic, 358* Determinants, 329*f
homogeneous, 51f bordered, 397, 400
line, 52 Developable surface, 303, 304, 314*f
orthogonal, 89f Dickson, L. E., 408
plane, 51 Differential, total, 90*
Pliicker, 52, 55, 63 Direction cosines, 27
point, 51 Discriminant of a quadratic form, 337
spherical, 196 Displacement group, 388*
Coplanar line vectors, Distributive Laws, 30, 35, 137
-points, 12f Divergence, of a vector (div f, V f),
-vectors, 2 183*f, 260*, 371
Coulomb's Law, 241 of a tensor, 371 *f, 376
Coriolis, acceleration of, 124" surface (Div f, V f), 207*
Theorem of, 124 Divergence Theorem, 235*
Cosine law, for plane triangles, 44 Division of quaternions, 409*
for spherical triangles, 46 417 Dot product, 29
Covariant components, of a ,ayadie, Double-dot product, 175
342 Double layer, 244
of a tensor, 344 Doublet, 244
of a vector, 48, 335 Dual angle, 69*
Curves, Bertrand, 132* -number, 64 *
congruent, 97 -vector, 63*
field of, 231*, 293* Dual of a tensor, 354, 370*, 376
of constant curvature, '100 Dupin, Theorem of, 305
parallel, 94*, 131 Dyad, 136*, 166
parametric, 205* Dyadic, 136*, 342
plane, 100f adjoint, 152*
reducible, 223* antisymmetric, 141
space, 88f complete, 139*
Cycloid, 131 conjugate, 141*
Cylinder, 304, 315 field, 293f
inertia, 159*
D'Alembert's Principle, 260 invariants of, 147, 148f, 150*
Darboux, G., 289 linear, 140*
Darboux vector, 93*, 105, 1:!2, 232 planar, 140*
Del (V), 184* second of a, 151*
Deltas, Kronecker, 46*, 253:',366, singular, 140*
402 stress, 253f
Derivative, covariant, 36:2*f 365f symmetric, 140*, 156
directional, 178, 182, :187 unit, 144*
normal, 180 zero, 137*, 140
of a determinant, 331 Dyadic equality, 136*
434 INDEX
E = [eie2ea], 49* Form, polar, 337*
E, F, G, 205*, 290* quadratic, 337f
e,f,g,291* second fundamental, 290*, 389
Einstein, A., 328 singular, 337 *
Electric intensity, 241* third fundamental, 291 *
Ellipsoid, of inertia, 160* Franklin, P., 276
Energy, of a fluid, 282 Frenet's Formulas, 93*, 100, 122
(kinetic) of a top, 176 Functional dependence, 190f
Energy ellipsoid, 177 Fundamental quantities, 291
Envelopes, 103, 132 -quadratic form, 339
Epsilons 329*, 346,
366,369 g = det gij, 339*, 369*
Equation, exact, 200 y (geodesic curvature), 284*, 295,
integrable, 200, 231 297
intrinsic, 103* Gammas (rte), 326, 356*, 368
Equiangular spiral, 130 Gauss, C. F., 300, 301, 307, 310
Equilibrium, of a cord, 298 derivative formulas of, 327, 390
of a deformable body, 255f Equation of, 301, 302, 306, 320, 392
of a fluid, 257 Theorem of, 307
of a rigid body, 76 Geodesic, 285*, 297f
Euclidean geometry, 385f Geodesic coordinates, 358*
-space, 368, 385* -curvature (y), 284*, 295, 297
Euler, L., 176, 260, 288, 311, 408, 428 -field, 299
Eulerian Equations, for fluid motion, -line, 285*
260f -lune, 310
for a top, 176 -torsion (t), 284*, 286, 305
Evolute, of a plane curve, 101 * -triangle, 310
Gibbs, J. W., 29, 136, 421, 426
Falling body, 124 Gradient, of a scalar, 179*
Field dyadic, 293f of a tensor, 187*
Field, of curves, 231*, 293*, 299 of a vector, 181 *
of geodesics, 299 surface, 206 *
of numbers, 408* Green's identities, 237, 238
of surfaces, 317 -Theorem, 216f
Field lines, 226f Group (definition), 341*
Flat space, 377*f, 285 displacement, 387f
Floating body, 257 of affine transformations, 340f
Flow, 266* of general transformations, 347
Force, body, 253, 255 of orthogonal transformations, 387
centrifugal, 124* of rigid motions, 387f
conservative, 261
Coriolis, 124* H = [rrn] = E(, 205* 291
on rigid body, 75f Half-turn, 427*, 429
surface, 253, 255 Hamilton, W. R., 403, 405, 408, 420,
transmissibility of, 75 421
Form, definite, 338* Hamilton-Cayley Equation, 160f,
first fundamental, 204*, 290*, 388 176, 291
indefinite, 339* Harmonic function, 196, 197, 214,
metric, 339* 239*, 244f, 247, 423f, 424
INDEX 435

Heat conduction, 247 Kinematics, of it, particle, 108f, 359


Helix, 105*f of a rigid body, 114E
circular, 107 Kinetic energy, of a fluid, 281, 282
Helmholtz's Equation, 263 of a top, 176
Homogeneous coordinates, 51 Kronecker delta (Sji), 46*, 343, 366
-strain, 163 generalized, 353*, 366, 369, 402
Kutta-Joukowsky Formulas, 273f
i, j, k, 26*
i, j, k, 404 * L, M, N, 290*
Idemfactor (I), 144*, 151, 34i3 Lagrange's Theorems, 62
Inertia, dyadic, 159* Lagrangian Equations for fluid mo-
moment of, 159* tion, 265
product of, 160* Laguerre, E., 289
Integral, circuit, 216f Lamb, H., 273
line, 222f, 224f Laplace expansion of a determinant,
surface, 218f, 221f, 233f 330
volume, 233f Laplace's Equation, 196, 197, 214,
Integrability condition, 20), 231 238, 247 *, 248
Interception (of a plane), 11': Laplacian (V2), 185*, 400, 423
Intrinsic equation, 103* Laurent series, 276
Invariable plane, 177 Lift, 277
Invariant, differential, 181, :.83, 207, Linear dependence of vectors, 7*
209 -relation between vectors, 9, 12, 19,
first scalar, 147*, 149 1'13 21, 22
of a dyadic, 147*f, 149 -relation between motors, 73
second scalar, 149*, 173 -vector function, 135*
third scalar, 149*, 17.3 Line integral, 216f, 218f, 222f, 224f
vector, 147*, 149, 173 Liouville, Formula of, 307
Invariant directions, 153f, 156.'
-planes, 155 Macduffee, C. C., 161
Inverse square law, 134 Magnetic shell, 244
Involute of a catenary, 103 Males and Dupin, Theorem of, 314
of a circle, 103, 131 Matric algebra, 169f
of a plane curve, 102* Matrix, 167, 169f, 203, 334, 336
of a space curve, 131 * Mean center, 20*
Irrotational vector, 198*f -curvature (J), 222, 286*, 287, 292,
392
J= 192* -value theorem (harmonic func-
J (mean curvature), 286* tions), 240
Jacobian, 190*f, 264, 347 Menelaus, Theorem of, 13
Joachimsthal, Theorem of, ;305 Metric, 339*
Metric quadratic form, 339*
K (total curvature), 286* -tensor, 349, 366
k (normal curvature), 284* Meusnier's Theorem, 285
K (curvature), 92*, 96 Minimal surface, 316*f
Kelvin's Theorem, on circu' ation, Minimum equation (of a dyadic),
267 161*
on minimum energy, 281 Mises, von, R., 68, 70
Kepler's Laws, 134 Mobius strip, 218*
436 INDEX
Moment, of inertia, 159* Operators:
of momentum, 176* Dij = DID, - D,Di, 378, 384
of a motor, about an axis, 77*f V (grad), 184, 215, 363
about a point, 63*f V2, 185
of a vector, about an axis, 55*f Vh, 362, 365
about a point, 55 *f Dij = vivj Vjvi, 384
Motion, of a fluid, 260f, 263f V. (Grad), 206
irrotational, 267 q( )q-', 419
line of, 267, 268 Orbit, of a planet, 134
plane, 270f Orthogonal coordinates, 194*f
steady, 268f -group, 387, 396
of a particle, 108f -surfaces (triple system), 305
circular, 110 -trajectories, 231, 300
relative, 110f, 117, 123f -transformations, 336*, 386, 387
under gravity, 124 -triple of unit vectors, 26*, 48, 194,
uniformly accelerated, 110 336
of a rigid body, 114f, 127 Outer product, of vectors, 354*
Motor, 63, 65*
acceleration, 127* P, Q, R, 294*
axis of, 66 * Parallel displacement, 312, 376f
derivative of, 126* -curves, 94 *
force, 77* Parallelogram law, 3
pitch of, 66* Parametric equations:
proper, 66* of a circular helix, 108
velocity, 117*, 120, 128 of a conoid, 321, 323
Motor identities, 73f of an ellipsoid, 322
-product, 70* of a general curve, 88
-sum, 67 of a general surface, 203
Moving trihedral, 92*, 122, 128, 283, of a hyperboloid, 322
312 of a paraboloid, 322
m-Vector, 370* of a right helicoid, 318
of a surface, of revolution, 305, 322
Nabla (V), 184*, 363, 421 of translation, 322
Nonion form of a dyadic, 167*f Particle, falling, 124
Norm, of a quaternion, 407* Pascal's Theorem, 169
Normal, principal, 92* Path curves, 377*
Normal derivative, 180* Permutation symbols, 329, 346
-form of a dyadic, 162f -tensor, 350 *
-plane, 133 Peterson and Morley, Theorem of, 74
-system of lines, 232, 313*f Pfaff's Problem, 230f
-vector to a surface, 205*, 208, 222, Pitch, of a motor, 66*
286, 316, 389 Plane, diametral, 176
Null line, 78* equation of, 51
-plane, 78* normal, 133
-system, 78*f osculating, 98*, 133
polar, 176
Operators: radical, 62
ta()a,418,419 rectifying, 133
Di = a/axi, 356, 362 Planet, motion of, 134
INDEX 437

PI(icker coordinates, , '. Quadrangle, complete, 16f


Poinsot's Theorem, 1; Quadratic forms, 337*f, 367
Point of division, 8*i Quadric surfaces, 175, 321, 322
Poisson's Equation, 2.t"' Quadrilateral, complete, 18
Polar form, 337 Quaternion, conjugate of (Kq),
-triangles, 45, 420 406 *
Polygon, plane, 39 norm of (Nq), 407*
spherical, 416, 420 real, 403
Polyhedron, 38 reciprocal of, 409*
Polyhedron Formula )f l:, ler, 311 * roots of, 412f
Position vector, 5*, 8 scalar part of (Sq), 405
Postfactor, 136* unit, 407*, 412, 414
Potential, complex, 27'4' * vector part of (Vq), 405
scalar, 200*, 241f zero, 403
vector, 202*, 245 Quaternion algebra, 403f
velocity, 267* -division, 409*
Prefactor, 136* -multiplication, 404*f
Pressure, fluid, 255, '& 5i' -units (i, j, k), 404*, 408
Principal axes, of ine 1t.a, .60* Quotient law, 399
of strain, 164*
of stress, 256* R, 294 *
Principal curvatures, 28(1 R, 90*
Principal directions, of a 1 yadic, 162* Radical axis, 62
on a surface, 287* -plane, 62
Principal normal, 92' Radius, of curvature (p), 94*
-stresses, 256 * of torsion (v), 94*
Principles, fundamental c f statics, Rank of a matrix, 203
75f Rate of change, local, 259*, 261
Product, box, 42 of a vector, 121
cross, 34 substantial, 259*, 261
dot, 29 Reciprocal bases, 46*f, 335, 367
double-dot, 175* -dyadics, 145*
indeterminate, 13(. -quaternions, 409*
motor, 70* -sets of motors, 74*f
of dual numbers, 64'' -sets of vectors, 46*f, 144, 192, 207,
of dyadics, 142* 209
of four vectors, 43 Reduced cofactor, 331 *
of matrices, 170* Reflections, 164f, 419
of quaternions, 404" Regular point of a surface, 203*
of tensor componcnl., 1150* Relative acceleration, 123*
of vectors, 410* -motion, 117
by numbers, 6* -scalar, 346*
outer, 354* -tensor, 345*f
scalar, of motors, 6,s" -velocity, 110*f, 123*
of vectors, 29*f Relativity, special, 401
scalar triple, 41 * Revolving fluid, 262
vector, 29, 34*f -unit vector, 90, 91
vector triple, 40* Ricci identity, 384
Pythagorean Theoroii-., 2 7 -tensor, 384
438 INDEX
Riemannian geometry, 366f Surface, of constant total curvature,
-space, 366*f, 377 307f, 310
Rodrigues' construction, 429 of revolution, 305, 307, 322
Rolling curve, 119 of translation, 322
Roots of a quaternion, 412 parallel, 325
Rotation (rot f, V x f), 183*f, 194, parametric equations of, 203*
195, 263 * quadric, 175, 321, 322
surface (Rot f, V. x f), 207* ruled, 303, 314*
Rotations in space, 164f, 396, 417f, tangent, 315*, 323
420, 421 Surface charges, 242f
Ruled surface, 303, 314 -divergence (Div f), 207*
Ruling, 303, 314 -gradient (Grad f), 206*f
-integral, 218f, 233f
Scalar, 1 -invariants, 209
Scalar product: -normal, 205 *
of two motors, 68*f -rotation (Rot f),
of two vectors, 29*f Symmetric dyadics, 141*, 156f
of three vectors, 41f -tensors, 352 *
Shear, components of, 256 Symmetry, of affine connection, 357,
Shift formulas, 52 361
Singular dyadic, 140* of curvature tensor, 381
-point of a surface, 203* of inertia dyadic, 160
-quadratic form, 337 of stress dyadic, 256
Solenoidal vector, 201 *f, 227
Solid angle, 236* T, N, B, 92*
Space charges, 245f t (geodesic torsion), 284
Spatial invariants, 209 z (torsion), 92*, 96
Speed, 108* Tangent vector, unit (T), 90*, 92
Spherical trigonometry, 44f, 416 Taylor, J. 11., 42
Static Equilibrium, Principle of, 76 Tensor, 172, 344, 348*
Statics, 75f absolute, 343f
Step path, 224 algebra, 350f
Stokes tensor, 372* antisymmetric, 352, 353
-Theorem, 220*, 309 contraction of, 351
Strain, homogeneous, 163 contravariant, 344
Stream-lines, 262*, 268, 269 covariant, 344
Stress dyadic, 253f covariant derivative of, 362*, 365*f
for a fluid, 255 curl of, 374 *
Sum, of dual numbers, 64* curvature, 379, 380*f
of dyadics, 137* differential invariants of a, 209f
of matrices, 170* V f,Vxf,188*
of motors, 67* V. f, V. x f, 207*
of quaternions, 403* divergence of, 372*
of tensors, 350* dual of, 370*
of vectors, 3*f gradient of, 362*
Summation convention, 188, 328*f Of, grad f, 188*
Surface, developable, 303, 304, V8f, Grad f, 206*
314*f metric, 339, 349, 361, 366
minimal, 316*f mixed, 344
INDEX 139

Tensor, relative, 345f Vector, bound, 2*


Iticci, 384* contravariant, 48*, 335
Stokes, 372*f covariant, 48*, 335
transformation equations of a, in Darboux, 93*, 106, 122, 133, 232
afline group, 344, 345 dual, 63 *
in general, 348* free, 2*
Tensor equations, 348 irrotational, 198*f
-operations, 350 line, 2*, 55, 63, 66
Tetradic, 172 moment, 63*
Theorems egregium (Gauss), 307 position, 5*, 8
Torricclli's Law, 270 proper, 1
Torsion (r), 92* rate of change of, 121
geodesic (t), 284*, 286 resultant, 63*
radius of (v), 93* solenoidal, 201 *f, 227
Total curvature (K), 286*, 287, 292, unit, 2*
306f, 392 zero, 1 *, 4
-differential, 197* Vector addition, 3*
Tractrix, 105*, 308 -algebra, 3f, 29f, 354
Transformation of integrals, surface -equations, 50
to line, 216, 218f, 221f -quantity, 1 *
volume to surface, 233f Velocity, 108*
Transpose, of a matrix, 167 absolute, 111
Triadic, 172* angular, 115*, 116*f
Trigonometry, plane, 21 complex, 272*
spherical, 21f, 416 general components of, 359
Trivector, 353 * of sound, 281
rectangular components of, 110
U, V, IV, 194* relative, 110f
Umbilical point, 288* transfer, 111 *
Unit dyadic (I), 144* Velocity four-vector, 401
-quaternion, 407*, 412, 414 -motor, 117*, 120
Unit vector, 2* Vortex, combined, 281
along binormal (B), 92* Vortex lines, 2fi6*, 269
along principal normal (N), 92* Vorticity, 263*
along surface normal (n), 205*
along tangent (T), 90*, 92 Wedderburn, J. 11. M., 169
derivative of, 91* Weingarten, Derivative formulas of,
radial (R), 90* 323, 327, 391
revolving, 90 Weight of a tensor, 345*
transverse (P), 90* Wind triangle, 111
Uniformly accelerated motion, 110
Zero dual number, 64*
Valence of a tensor, 172*, 344, 345 -dyadic, 137*
Veblen, 0., 366, 368, 380 -motor, 66*
Vector, 1 * -quaternion, 403*
are, 415 -tensor, 348
base, 24f -vector, 1*, 2*, 4

You might also like