18.07.2014 Views

Introduction to Vectors and Tensors Vol 2 (Bowen 246). - Index of

Introduction to Vectors and Tensors Vol 2 (Bowen 246). - Index of

Introduction to Vectors and Tensors Vol 2 (Bowen 246). - Index of

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

INTRODUCTION TO VECTORS AND TENSORS<br />

Vec<strong>to</strong>r <strong>and</strong> Tensor Analysis<br />

<strong>Vol</strong>ume 2<br />

Ray M. <strong>Bowen</strong><br />

Mechanical Engineering<br />

Texas A&M University<br />

College Station, Texas<br />

<strong>and</strong><br />

C.-C. Wang<br />

Mathematical Sciences<br />

Rice University<br />

Hous<strong>to</strong>n, Texas<br />

Copyright Ray M. <strong>Bowen</strong> <strong>and</strong> C.-C. Wang<br />

(ISBN 0-306-37509-5 (v. 2))


____________________________________________________________________________<br />

PREFACE<br />

To <strong>Vol</strong>ume 2<br />

This is the second volume <strong>of</strong> a two-volume work on vec<strong>to</strong>rs <strong>and</strong> tensors. <strong>Vol</strong>ume 1 is concerned<br />

with the algebra <strong>of</strong> vec<strong>to</strong>rs <strong>and</strong> tensors, while this volume is concerned with the geometrical<br />

aspects <strong>of</strong> vec<strong>to</strong>rs <strong>and</strong> tensors. This volume begins with a discussion <strong>of</strong> Euclidean manifolds. The<br />

principal mathematical entity considered in this volume is a field, which is defined on a domain in a<br />

Euclidean manifold. The values <strong>of</strong> the field may be vec<strong>to</strong>rs or tensors. We investigate results due<br />

<strong>to</strong> the distribution <strong>of</strong> the vec<strong>to</strong>r or tensor values <strong>of</strong> the field on its domain. While we do not discuss<br />

general differentiable manifolds, we do include a chapter on vec<strong>to</strong>r <strong>and</strong> tensor fields defined on<br />

hypersurfaces in a Euclidean manifold.<br />

This volume contains frequent references <strong>to</strong> <strong>Vol</strong>ume 1. However, references are limited <strong>to</strong><br />

basic algebraic concepts, <strong>and</strong> a student with a modest background in linear algebra should be able<br />

<strong>to</strong> utilize this volume as an independent textbook. As indicated in the preface <strong>to</strong> <strong>Vol</strong>ume 1, this<br />

volume is suitable for a one-semester course on vec<strong>to</strong>r <strong>and</strong> tensor analysis. On occasions when we<br />

have taught a one –semester course, we covered material from Chapters 9, 10, <strong>and</strong> 11 <strong>of</strong> this<br />

volume. This course also covered the material in Chapters 0,3,4,5, <strong>and</strong> 8 from <strong>Vol</strong>ume 1.<br />

We wish <strong>to</strong> thank the U.S. National Science Foundation for its support during the<br />

preparation <strong>of</strong> this work. We also wish <strong>to</strong> take this opportunity <strong>to</strong> thank Dr. Kurt Reinicke for<br />

critically checking the entire manuscript <strong>and</strong> <strong>of</strong>fering improvements on many points.<br />

Hous<strong>to</strong>n, Texas<br />

R.M.B.<br />

C.-C.W.<br />

iii


__________________________________________________________________________<br />

CONTENTS<br />

<strong>Vol</strong>. 2<br />

Vec<strong>to</strong>r <strong>and</strong> Tensor Analysis<br />

Contents <strong>of</strong> <strong>Vol</strong>ume 1…………………………………………………<br />

vii<br />

PART III. VECTOR AND TENSOR ANALYSIS<br />

Selected Readings for Part III………………………………………… 296<br />

CHAPTER 9. Euclidean Manifolds………………………………………….. 297<br />

Section 43. Euclidean Point Spaces……………………………….. 297<br />

Section 44. Coordinate Systems…………………………………… 306<br />

Section 45. Transformation Rules for Vec<strong>to</strong>r <strong>and</strong> Tensor Fields…. 324<br />

Section 46. Anholonomic <strong>and</strong> Physical Components <strong>of</strong> <strong>Tensors</strong>…. 332<br />

Section 47. Chris<strong>to</strong>ffel Symbols <strong>and</strong> Covariant Differentiation…... 339<br />

Section 48. Covariant Derivatives along Curves………………….. 353<br />

CHAPTER 10. Vec<strong>to</strong>r Fields <strong>and</strong> Differential Forms………………………... 359<br />

Section 49. Lie Derivatives……………………………………….. 359<br />

Section 5O. Frobenius Theorem…………………………………… 368<br />

Section 51. Differential Forms <strong>and</strong> Exterior Derivative………….. 373<br />

Section 52. The Dual Form <strong>of</strong> Frobenius Theorem: the Poincaré<br />

Lemma……………………………………………….. 381<br />

Section 53. Vec<strong>to</strong>r Fields in a Three-Dimensiona1 Euclidean<br />

Manifold, I. Invariants <strong>and</strong> Intrinsic Equations…….. 389<br />

Section 54. Vec<strong>to</strong>r Fields in a Three-Dimensiona1 Euclidean<br />

Manifold, II. Representations for Special<br />

Class <strong>of</strong> Vec<strong>to</strong>r Fields………………………………. 399<br />

CHAPTER 11.<br />

Hypersurfaces in a Euclidean Manifold<br />

Section 55. Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface Metric… 407<br />

Section 56. Surface Covariant Derivatives………………………. 416<br />

Section 57. Surface Geodesics <strong>and</strong> the Exponential Map……….. 425<br />

Section 58. Surface Curvature, I. The Formulas <strong>of</strong> Weingarten<br />

<strong>and</strong> Gauss…………………………………………… 433<br />

Section 59. Surface Curvature, II. The Riemann-Chris<strong>to</strong>ffel<br />

Tensor <strong>and</strong> the Ricci Identities……………………... 443<br />

Section 60. Surface Curvature, III. The Equations <strong>of</strong> Gauss <strong>and</strong> Codazzi 449<br />

v


vi CONTENTS OF VOLUME 2<br />

Section 61. Surface Area, Minimal Surface………........................ 454<br />

Section 62. Surfaces in a Three-Dimensional Euclidean Manifold. 457<br />

CHAPTER 12.<br />

Elements <strong>of</strong> Classical Continuous Groups<br />

Section 63. The General Linear Group <strong>and</strong> Its Subgroups……….. 463<br />

Section 64. The Parallelism <strong>of</strong> Cartan……………………………. 469<br />

Section 65. One-Parameter Groups <strong>and</strong> the Exponential Map…… 476<br />

Section 66. Subgroups <strong>and</strong> Subalgebras…………………………. 482<br />

Section 67. Maximal Abelian Subgroups <strong>and</strong> Subalgebras……… 486<br />

CHAPTER 13.<br />

Integration <strong>of</strong> Fields on Euclidean Manifolds, Hypersurfaces, <strong>and</strong><br />

Continuous Groups<br />

Section 68. Arc Length, Surface Area, <strong>and</strong> <strong>Vol</strong>ume……………... 491<br />

Section 69. Integration <strong>of</strong> Vec<strong>to</strong>r Fields <strong>and</strong> Tensor Fields……… 499<br />

Section 70. Integration <strong>of</strong> Differential Forms……………………. 503<br />

Section 71. Generalized S<strong>to</strong>kes’ Theorem……………………….. 507<br />

Section 72. Invariant Integrals on Continuous Groups…………... 515<br />

INDEX……………………………………………………………………….<br />

x


______________________________________________________________________________<br />

CONTENTS<br />

<strong>Vol</strong>. 1<br />

Linear <strong>and</strong> Multilinear Algebra<br />

PART 1 BASIC MATHEMATICS<br />

Selected Readings for Part I………………………………………………………… 2<br />

CHAPTER 0 Elementary Matrix Theory…………………………………………. 3<br />

CHAPTER 1 Sets, Relations, <strong>and</strong> Functions……………………………………… 13<br />

Section 1. Sets <strong>and</strong> Set Algebra………………………………………... 13<br />

Section 2. Ordered Pairs" Cartesian Products" <strong>and</strong> Relations…………. 16<br />

Section 3. Functions……………………………………………………. 18<br />

CHAPTER 2 Groups, Rings <strong>and</strong> Fields…………………………………………… 23<br />

Section 4. The Axioms for a Group……………………………………. 23<br />

Section 5. Properties <strong>of</strong> a Group……………………………………….. 26<br />

Section 6. Group Homomorphisms…………………………………….. 29<br />

Section 7. Rings <strong>and</strong> Fields…………………………………………….. 33<br />

PART I1 VECTOR AND TENSOR ALGEBRA<br />

Selected Readings for Part II………………………………………………………… 40<br />

CHAPTER 3 Vec<strong>to</strong>r Spaces……………………………………………………….. 41<br />

Section 8. The Axioms for a Vec<strong>to</strong>r Space…………………………….. 41<br />

Section 9. Linear Independence, Dimension <strong>and</strong> Basis…………….….. 46<br />

Section 10. Intersection, Sum <strong>and</strong> Direct Sum <strong>of</strong> Subspaces……………. 55<br />

Section 11. Fac<strong>to</strong>r Spaces………………………………………………... 59<br />

Section 12. Inner Product Spaces………………………..………………. 62<br />

Section 13. Orthogonal Bases <strong>and</strong> Orthogonal Compliments…………… 69<br />

Section 14. Reciprocal Basis <strong>and</strong> Change <strong>of</strong> Basis……………………… 75<br />

CHAPTER 4. Linear Transformations……………………………………………… 85<br />

Section 15. Definition <strong>of</strong> a Linear Transformation………………………. 85<br />

Section 16. Sums <strong>and</strong> Products <strong>of</strong> Linear Transformations……………… 93


viii CONTENTS OF VOLUME 2<br />

Section 17. Special Types <strong>of</strong> Linear Transformations…………………… 97<br />

Section 18. The Adjoint <strong>of</strong> a Linear Transformation…………………….. 105<br />

Section 19. Component Formulas………………………………………... 118<br />

CHAPTER 5. Determinants <strong>and</strong> Matrices…………………………………………… 125<br />

Section 20. The Generalized Kronecker Deltas<br />

<strong>and</strong> the Summation Convention……………………………… 125<br />

Section 21. Determinants…………………………………………………. 130<br />

Section 22. The Matrix <strong>of</strong> a Linear Transformation……………………… 136<br />

Section 23 Solution <strong>of</strong> Systems <strong>of</strong> Linear Equations…………………….. 142<br />

CHAPTER 6 Spectral Decompositions……………………………………………... 145<br />

Section 24. Direct Sum <strong>of</strong> Endomorphisms……………………………… 145<br />

Section 25. Eigenvec<strong>to</strong>rs <strong>and</strong> Eigenvalues……………………………….. 148<br />

Section 26. The Characteristic Polynomial………………………………. 151<br />

Section 27. Spectral Decomposition for Hermitian Endomorphisms…….. 158<br />

Section 28. Illustrative Examples…………………………………………. 171<br />

Section 29. The Minimal Polynomial……………………………..……… 176<br />

Section 30. Spectral Decomposition for Arbitrary Endomorphisms….….. 182<br />

CHAPTER 7. Tensor Algebra………………………………………………………. 203<br />

Section 31. Linear Functions, the Dual Space…………………………… 203<br />

Section 32. The Second Dual Space, Canonical Isomorphisms…………. 213<br />

Section 33. Multilinear Functions, <strong>Tensors</strong>…………………………..….. 218<br />

Section 34. Contractions…......................................................................... 229<br />

Section 35. <strong>Tensors</strong> on Inner Product Spaces……………………………. 235<br />

CHAPTER 8. Exterior Algebra……………………………………………………... 247<br />

Section 36. Skew-Symmetric <strong>Tensors</strong> <strong>and</strong> Symmetric <strong>Tensors</strong>………….. 247<br />

Section 37. The Skew-Symmetric Opera<strong>to</strong>r……………………………… 250<br />

Section 38. The Wedge Product………………………………………….. 256<br />

Section 39. Product Bases <strong>and</strong> Strict Components……………………….. 263<br />

Section 40. Determinants <strong>and</strong> Orientations………………………………. 271<br />

Section 41. Duality……………………………………………………….. 280<br />

Section 42. Transformation <strong>to</strong> Contravariant Representation……………. 287<br />

INDEX…………………………………………………………………………………….x


_____________________________________________________________________________<br />

PART III<br />

VECTOR AND TENSOR ANALYSIS


Selected Reading for Part III<br />

BISHOP, R. L., <strong>and</strong> R. J. CRITTENDEN, Geometry <strong>of</strong> Manifolds, Academic Press, New York, 1964<br />

BISHOP, R. L., <strong>and</strong> R. J. CRITTENDEN, Tensor Analysis on Manifolds, Macmillian, New York, 1968.<br />

CHEVALLEY, C., Theory <strong>of</strong> Lie Groups, Prince<strong>to</strong>n University Press, Prince<strong>to</strong>n, New Jersey, 1946<br />

COHN, P. M., Lie Groups, Cambridge University Press, Cambridge, 1965.<br />

EISENHART, L. P., Riemannian Geometry, Prince<strong>to</strong>n University Press, Prince<strong>to</strong>n, New Jersey, 1925.<br />

ERICKSEN, J. L., Tensor Fields, an appendix in the Classical Field Theories, <strong>Vol</strong>. III/1.<br />

Encyclopedia <strong>of</strong> Physics, Springer-Verlag, Berlin-Gottingen-Heidelberg, 1960.<br />

FLANDERS, H., Differential Forms with Applications in the Physical Sciences, Academic Press,<br />

New York, 1963.<br />

KOBAYASHI, S., <strong>and</strong> K. NOMIZU, Foundations <strong>of</strong> Differential Geometry, <strong>Vol</strong>s. I <strong>and</strong> II, Interscience,<br />

New York, 1963, 1969.<br />

LOOMIS, L. H., <strong>and</strong> S. STERNBERG, Advanced Calculus, Addison-Wesley, Reading, Massachusetts,<br />

1968.<br />

MCCONNEL, A. J., Applications <strong>of</strong> Tensor Analysis, Dover Publications, New York, 1957.<br />

NELSON, E., Tensor Analysis, Prince<strong>to</strong>n University Press, Prince<strong>to</strong>n, New Jersey, 1967.<br />

NICKERSON, H. K., D. C. SPENCER, <strong>and</strong> N. E. STEENROD, Advanced Calculus, D. Van Nostr<strong>and</strong>,<br />

Prince<strong>to</strong>n, New Jersey, 1958.<br />

SCHOUTEN, J. A., Ricci Calculus, 2nd ed., Springer-Verlag, Berlin, 1954.<br />

STERNBERG, S., Lectures on Differential Geometry, Prentice-Hall, Englewood Cliffs, New Jersey,<br />

1964.<br />

WEATHERBURN, C. E., An <strong>Introduction</strong> <strong>to</strong> Riemannian Geometry <strong>and</strong> the Tensor Calculus,<br />

Cambridge University Press, Cambridge, 1957.


_______________________________________________________________________________<br />

Chapter 9<br />

EUCLIDEAN MANIFOLDS<br />

This chapter is the first where the algebraic concepts developed thus far are combined with<br />

ideas from analysis. The main concept <strong>to</strong> be introduced is that <strong>of</strong> a manifold. We will discuss here<br />

only a special case cal1ed a Euclidean manifold. The reader is assumed <strong>to</strong> be familiar with certain<br />

elementary concepts in analysis, but, for the sake <strong>of</strong> completeness, many <strong>of</strong> these shall be inserted<br />

when needed.<br />

Section 43<br />

Euclidean Point Spaces<br />

Consider an inner produce space V <strong>and</strong> a set E . The set E is a Euclidean point space if<br />

there exists a function f : E × E →V such that:<br />

<strong>and</strong><br />

(a) f( xy , ) = f( xz , ) + f( zy , ), xyz , , ∈E<br />

(b) For every x ∈E <strong>and</strong> v ∈V there exists a unique element y ∈E such that<br />

f ( x, y)<br />

= v .<br />

The elements <strong>of</strong> E are called points, <strong>and</strong> the inner product space V is called the translation space.<br />

We say that f ( x, y ) is the vec<strong>to</strong>r determined by the end point x <strong>and</strong> the initial point y . Condition<br />

b) above is equivalent <strong>to</strong> requiring the function fx : E →V defined by fx( y) = f( x, y ) <strong>to</strong> be one <strong>to</strong><br />

one for each x . The dimension <strong>of</strong> E , written dimE , is defined <strong>to</strong> be the dimension <strong>of</strong> V . If V<br />

does not have an inner product, the set E defined above is called an affine space.<br />

A Euclidean point space is not a vec<strong>to</strong>r space but a vec<strong>to</strong>r space with inner product is made<br />

a Euclidean point space by defining f ( v1, v2)<br />

≡ v1−<br />

v<br />

2<br />

for all v ∈V . For an arbitrary point space<br />

the function f is called the point difference, <strong>and</strong> it is cus<strong>to</strong>mary <strong>to</strong> use the suggestive notation<br />

f ( x, y)<br />

= x − y (43.1)<br />

In this notation (a) <strong>and</strong> (b) above take the forms<br />

297


298 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

x − y = x− z+ z−y (43.2)<br />

<strong>and</strong><br />

x − y = v (43.3)<br />

Theorem 43.1. In a Euclidean point space E<br />

() i x− x = 0<br />

( ii) x− y = −( y−x)<br />

( iii) if x − y = x ' −y ', then x− x' = y−y'<br />

(43.4)<br />

Pro<strong>of</strong>. For (i) take x = y = z in (43.2); then<br />

x− x = x− x+ x−x<br />

which implies x− x = 0. To obtain (ii) take y = x in (43.2) <strong>and</strong> use (i). For (iii) observe that<br />

x − y ' = x − y+ y− y' = x− x' + x' −y<br />

'<br />

from (43.2). However, we are given x − y = x ' −y ' which implies (iii).<br />

The equation<br />

x − y = v<br />

has the property that given any v <strong>and</strong> y , x is uniquely determined. For this reason it is cus<strong>to</strong>mary<br />

<strong>to</strong> write<br />

x = y + v (43.5)


Sec. 43 • Euclidean Point Spaces 299<br />

for the point x uniquely determined by<br />

y ∈E <strong>and</strong> v ∈V .<br />

The distance from x <strong>to</strong> y , written d( x, y ) , is defined by<br />

{( ) ( )} 1/2<br />

d ( x, y)<br />

= x− y = x−y ⋅ x−y (43.6)<br />

It easily fol1ows from the definition (43.6) <strong>and</strong> the properties <strong>of</strong> the inner product that<br />

d( x, y) = d( y, x )<br />

(43.7)<br />

<strong>and</strong><br />

d( x, y) ≤ d( xz , ) + d( zy , )<br />

(43.8)<br />

for all x,<br />

y , <strong>and</strong> z in E . Equation (43.8) is simply rewritten in terms <strong>of</strong> the points x,<br />

y , <strong>and</strong> z<br />

rather than the vec<strong>to</strong>rs x −y,<br />

x−z, <strong>and</strong> z − y . It is also apparent from (43.6) that<br />

d( x, y) ≥ 0 <strong>and</strong> d( x, y) = 0⇔ x = y (43.9)<br />

The properties (43.7)-(43.9) establish that E is a metric space.<br />

There are several concepts from the theory <strong>of</strong> metric spaces which we need <strong>to</strong> summarize.<br />

For simplicity the definitions are sated here in terms <strong>of</strong> Euclidean point spaces only even though<br />

they can be defined for metric spaces in general.<br />

In a Euclidean point space E an open ball <strong>of</strong> radius ε > 0 centered at x ∈ 0<br />

E is the set<br />

{ d ε}<br />

B( x , ε ) = x ( x, x ) <<br />

(43.10)<br />

0 0<br />

<strong>and</strong> a closed ball is the set


300 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

{ d ε}<br />

B( x , ε ) = x ( x, x ) ≤<br />

(43.11)<br />

0 0<br />

A neighborhood <strong>of</strong> x ∈E is a set which contains an open ball centered at x . A subset U <strong>of</strong> E is<br />

open if it is a neighborhood <strong>of</strong> each <strong>of</strong> its points. The empty set ∅ is trivially open because it<br />

contains no points. It also follows from the definitions that E is open.<br />

Theorem 43.2. An open ball is an open set.<br />

Pro<strong>of</strong>. Consider the open ball B( x<br />

0, ε ). Let x be an arbitrary point in B( x<br />

0, ε ). Then<br />

ε − d( xx ,<br />

0) > 0 <strong>and</strong> the open ball B( x, ε − d( x, x<br />

0))<br />

is in B( x<br />

0, ε ), because if<br />

y ∈B( x, ε −d( x, x<br />

0))<br />

, then d( y, x) < ε −d( x0, x ) <strong>and</strong>, by (43.8), d( x0, y) ≤ d( x0, x) + d( x, y ),<br />

which yields d( x0, y ) < ε <strong>and</strong>, thus, y∈ B( x<br />

0, ε ).<br />

A subset U <strong>of</strong> E is closed if its complement, EU, is open. It can be shown that closed<br />

balls are indeed closed sets. The empty set, ∅ , is closed because E = E ∅ is open. By the same<br />

logic E is closed since EE =∅ is open. In fact ∅ <strong>and</strong> E are the only subsets <strong>of</strong> E which are<br />

both open <strong>and</strong> closed. A subset U ⊂ E is bounded if it is contained in some open ball. A subset<br />

U ⊂ E is compact if it is closed <strong>and</strong> bounded.<br />

Theorem 43.3. The union <strong>of</strong> any collection <strong>of</strong> open sets is open.<br />

Pro<strong>of</strong>. Let { U<br />

α<br />

α ∈ I}<br />

be a collection <strong>of</strong> open sets, where I is an index set. Assume that<br />

x ∈∪<br />

α∈IUα<br />

. Then x must belong <strong>to</strong> at least one <strong>of</strong> the sets in the collection, say U<br />

α 0<br />

. Since U<br />

α 0<br />

is<br />

open, there exists an open ball B( x, ε ) ⊂Uα ⊂∪ 0 α∈IUα<br />

. Thus, B( x, ε ) ⊂ ∪<br />

α∈IUα<br />

. Since x is<br />

arbitrary, ∪ is open.<br />

U α∈I<br />

α<br />

Theorem43.4. The intersection <strong>of</strong> a finite collection <strong>of</strong> open sets is open.<br />

Pro<strong>of</strong>. Let { }<br />

1 ,..., α<br />

n<br />

U U be a finite family <strong>of</strong> open sets. If ∩ U i= 1 i<br />

is empty, the assertion is trivial.<br />

Thus, assume ∩ n<br />

U is not empty <strong>and</strong> let x be an arbitrary element <strong>of</strong> ∩ n<br />

U . Then i= 1 i<br />

i= 1 i<br />

x ∈U i<br />

for<br />

i = 1,..., n <strong>and</strong> there is an open ball B( x, ε i<br />

) ⊂U i<br />

for i = 1,..., n. Let ε be the smallest <strong>of</strong> the<br />

positive numbers ,..., n<br />

n<br />

ε1 ε<br />

n<br />

. Then x∈B( x, ε ) ⊂∩ i=<br />

1U<br />

i<br />

. Thus ∩ U i= 1 i<br />

is open.


Sec. 43 • Euclidean Point Spaces 301<br />

It should be noted that arbitrary intersections <strong>of</strong> open sets will not always lead <strong>to</strong> open sets.<br />

− 1 n,1<br />

n ,<br />

The st<strong>and</strong>ard counter example is given by the family <strong>of</strong> open sets <strong>of</strong> R <strong>of</strong> the form ( )<br />

∞<br />

n = 1,2,3.... The intersection ∩ ( ) is the set { 0 } which is not open.<br />

n 1<br />

1 n,1<br />

n<br />

=<br />

−<br />

By a sequence in E we mean a function on the positive integers { 1,2,3,..., n ,...}<br />

with values<br />

in E . The notation { x1, x2, x3..., x<br />

n,...<br />

} , or simply { x n }, is usually used <strong>to</strong> denote the values <strong>of</strong> the<br />

sequence. A sequence { x n } is said <strong>to</strong> converge <strong>to</strong> a limit x ∈E if for every open ball B( x , ε )<br />

centered at x , there exists a positive integer n ( ) 0<br />

ε such that x ∈ B( x, ε ) whenever n≥<br />

n ( ε ).<br />

0<br />

Equivalently, a sequence { x n } converges <strong>to</strong> x if for every real number ε > 0 there exists a positive<br />

integer n ( ) 0<br />

ε such that d( x , x ) < ε for all n><br />

n ( ε ).<br />

If { }<br />

0<br />

x converges <strong>to</strong> x , it is conventional <strong>to</strong><br />

write<br />

n<br />

n<br />

n<br />

x = lim x or x→x<br />

as n →∞<br />

n→∞<br />

n<br />

n<br />

Theorem 43.5. If { x n } converges <strong>to</strong> a limit, then the limit is unique.<br />

Pro<strong>of</strong>. Assume that<br />

x = lim x , y = lim x<br />

n→∞<br />

n<br />

n→∞<br />

n<br />

Then, from (43.8)<br />

d( x, y) ≤ d( xx , ) + d( x, y )<br />

n<br />

n<br />

for every n . Let ε be an arbitrary positive real number. Then from the definition <strong>of</strong> convergence<br />

<strong>of</strong> { x n }, there exists an integer n ( ε ) such that n≥<br />

n ( ε ) implies ( , )<br />

0 0<br />

d xx<br />

n<br />

< ε <strong>and</strong> d( xn, y ) < ε .<br />

Therefore,<br />

d( xy , ) ≤ 2ε<br />

for arbitrary ε . This result implies d ( xy , ) = 0 <strong>and</strong>, thus, x = y .


302 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

A point x ∈E is a limit point <strong>of</strong> a subset U ⊂ E if every neighborhood <strong>of</strong> x contains a<br />

point <strong>of</strong> U distinct from x . Note that x need not be in U . For example, the sphere<br />

{ x d( x, x<br />

0)<br />

= ε}<br />

are limit points <strong>of</strong> the open ball B( x<br />

0, ε ). The closure <strong>of</strong> U ⊂ E , written U , is<br />

the union <strong>of</strong> U <strong>and</strong> its limit points. For example, the closure <strong>of</strong> the open ball B( x<br />

0, ε ) is the<br />

closed ball B( x<br />

0, ε ). It is a fact that the closure <strong>of</strong> U is the smallest closed set containing U .<br />

Thus U is closed if <strong>and</strong> only if U = U .<br />

The reader is cautioned not <strong>to</strong> confuse the concepts limit <strong>of</strong> a sequence <strong>and</strong> limit point <strong>of</strong> a<br />

subset. A sequence is not a subset <strong>of</strong> E ; it is a function with values in E . A sequence may have a<br />

limit when it has no limit point. Likewise the set <strong>of</strong> pints which represent the values <strong>of</strong> a sequence<br />

may have a limit point when the sequence does not converge <strong>to</strong> a limit. However, these two<br />

concepts are related by the following result from the theory <strong>of</strong> metric spaces: a point x is a limit<br />

point <strong>of</strong> a set U if <strong>and</strong> only if there exists a convergent sequence <strong>of</strong> distinct points <strong>of</strong> U with x as<br />

a limit.<br />

A mapping f : U →E ', where U is an open set in E <strong>and</strong> E ' is a Euclidean point space or<br />

an inner product space, is continuous at x ∈ 0<br />

U if for every real number ε > 0 there exists a real<br />

number δ ( x<br />

0, ε ) > 0 such that d( xx ,<br />

0) < δ ( ε, x<br />

0)<br />

implies d'( f( x0), f( x )) < ε . Here d ' is the<br />

distance function for E ' . When f is continuous at x<br />

0<br />

, it is conventional <strong>to</strong> write<br />

lim f( x) or f( x) → f( x ) as x→x<br />

x→x0<br />

0 0<br />

The mapping f is continuous on U if it is continuous at every point <strong>of</strong> U . A continuous mapping<br />

is called a homomorphism if it is one-<strong>to</strong>-one <strong>and</strong> if its inverse is also continuous. What we have<br />

just defined is sometimes called a homeomorphism in<strong>to</strong>. If a homomorphism is also on<strong>to</strong>, then it is<br />

called specifically a homeomorphism on<strong>to</strong>. It is easily verified that a composition <strong>of</strong> two<br />

continuous maps is a continuous map <strong>and</strong> the composition <strong>of</strong> two homomorphisms is a<br />

homomorphism.<br />

A mapping f : U →E ' , where U <strong>and</strong> E ' are defined as before, is differentiable at<br />

there exists a linear transformation A ∈LV ( ; V ') such that<br />

x<br />

x ∈U if<br />

f( x+ v) = f( x) + A v+<br />

o( x, v )<br />

(43.12)<br />

x


Sec. 43 • Euclidean Point Spaces 303<br />

where<br />

lim o( x , v ) = 0<br />

v→0<br />

v<br />

(43.13)<br />

In the above definition V ' denotes the translation space <strong>of</strong> E ' .<br />

Theorem 43.6. The linear transformation<br />

A<br />

x<br />

in (43.12) is unique.<br />

Pro<strong>of</strong>. If (43.12) holds for<br />

A<br />

x<br />

<strong>and</strong><br />

A<br />

x<br />

, then by subtraction we find<br />

By (43.13), ( x<br />

−<br />

x)<br />

( Ax<br />

− Ax) v = o( x, v ) −o( x, v )<br />

A A e must be zero for each unit vec<strong>to</strong>r e , so that A = A .<br />

x<br />

x<br />

If f is differentiable at every point <strong>of</strong> U , then we can define a mapping<br />

grad f : U → L ( V ; V ') , called the gradient <strong>of</strong> f , by<br />

grad f ( x) = A , x∈U (43.14)<br />

x<br />

1<br />

If grad f is continuous on U , then f is said <strong>to</strong> be <strong>of</strong> class C . If grad f exists <strong>and</strong> is<br />

1<br />

2<br />

r<br />

itself <strong>of</strong> class C , then f is <strong>of</strong> class C . More generally, f is <strong>of</strong> class C , r > 0 , if it is <strong>of</strong> class<br />

r 1<br />

C − r−1<br />

<strong>and</strong> its ( r −1)<br />

st gradient, written grad f<br />

continuous on U . If f is a<br />

r<br />

is called a C diffeomorphism.<br />

, is <strong>of</strong> class<br />

r<br />

C one-<strong>to</strong>-one map with a<br />

1<br />

C . Of course, f is <strong>of</strong> class<br />

r<br />

C inverse<br />

0<br />

C if it is<br />

1<br />

f − defined on f ( U ), then f<br />

If f is differentiable at x , then it follows from (43.12) that<br />

f( x+ τ v) − f( x)<br />

d<br />

Av<br />

x<br />

= lim = f ( x+<br />

τ v )<br />

(43.15)<br />

τ →0<br />

τ dτ<br />

τ = 0


304 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

for all<br />

v ∈V . To obtain (43.15) replace v by τ v , τ > 0 in (43.12) <strong>and</strong> write the result as<br />

f( x+ τ v) − f( x) o( x, τ v )<br />

Av<br />

x<br />

= −<br />

(43.16)<br />

τ<br />

τ<br />

By (43.13) the limit <strong>of</strong> the last term is zero as τ → 0 , <strong>and</strong> (43.15) is obtained. Equation (43.15)<br />

holds for all v ∈V because we can always choose τ in (43.16) small enough <strong>to</strong> ensure that x+<br />

τ v<br />

is in U , the domain <strong>of</strong> f . If f is differentiable at every x ∈U , then (43.15) can be written<br />

d<br />

grad ( x) v = f( x+<br />

τ v )<br />

(43.17)<br />

d<br />

( f )<br />

τ τ = 0<br />

A function f : U →R, where U is an open subset <strong>of</strong> E , is called a scalar field. Similarly,<br />

f : U →V is a vec<strong>to</strong>r field, <strong>and</strong> f : U → T ( V ) is a tensor field <strong>of</strong> order q . It should be noted<br />

that the term field is defined here is not the same as that in Section 7.<br />

q<br />

Before closing this section there is an important theorem which needs <strong>to</strong> be recorded for<br />

later use. We shall not prove this theorem here, but we assume that the reader is familiar with the<br />

result known as the inverse mapping theorem in multivariable calculus.<br />

r<br />

Theorem 43.7. Let f : U →E ' be a C mapping <strong>and</strong> assume that grad f ( x<br />

0)<br />

is a linear<br />

isomorphism. Then there exists a neighborhood U<br />

1<br />

<strong>of</strong> x<br />

0<br />

such that the restriction <strong>of</strong> f <strong>to</strong> U<br />

1<br />

is a<br />

r<br />

C diffeomorphism. In addition<br />

−<br />

( ) 1<br />

x x (43.18)<br />

−1<br />

grad f ( f( 0)) = grad f( 0)<br />

This theorem provides a condition under which one can asert the existence <strong>of</strong> a local inverse <strong>of</strong> a<br />

smooth mapping.<br />

Exercises<br />

43.1 Let a sequence { x n } converge <strong>to</strong> x . Show that every subsequence <strong>of</strong> { n }<br />

converges <strong>to</strong> x .<br />

x also


Sec. 43 • Euclidean Point Spaces 305<br />

43.2 Show that arbitrary intersections <strong>and</strong> finite unions <strong>of</strong> closed sets yields closed sets.<br />

43.3 Let f : U → E ' , where U is open in E , <strong>and</strong> E ' is either a Euclidean point space or an<br />

inner produce space. Show that f is continuous on U if <strong>and</strong> only if f −1<br />

( D ) is open in<br />

E for all D open in f ( U ) .<br />

43.4 Let f : U → E ' be a homeomorphism. Show that f maps any open set in U on<strong>to</strong> an<br />

open set in E ' .<br />

43.5 If f is a differentiable scalar valued function on LV ( ; V ) , show that the gradient <strong>of</strong> f<br />

at A ∈LV ( ; V ) , written<br />

∂f<br />

( A)<br />

∂A<br />

is a linear transformation in LV ( ; V ) defined by<br />

⎛ ∂f<br />

T ⎞ df<br />

tr ⎜ ( ) ⎟ = +<br />

⎝∂A<br />

AB ⎠ d<br />

A B<br />

( τ )<br />

τ τ = 0<br />

for all B ∈LV ( ; V )<br />

43.6 Show that<br />

∂μ 1<br />

∂μ ( ) = <strong>and</strong> N<br />

( ) = adj<br />

∂A<br />

A I ∂A<br />

A A<br />

( )<br />

T


306 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

Section 44 Coordinate Systems<br />

r<br />

Given a Euclidean point space E <strong>of</strong> dimension N , we define a C -chart at x ∈E <strong>to</strong> be a<br />

N<br />

r<br />

U , xˆ<br />

, where U is an open set in E containing x <strong>and</strong> xˆ : U →R is a C diffeomorphism.<br />

i<br />

U , xˆ<br />

, there are N scalar fields xˆ : U →R such that<br />

pair ( )<br />

Given any chart ( )<br />

1<br />

N<br />

( )<br />

xˆ( x) = xˆ ( x),..., xˆ<br />

( x )<br />

(44.1)<br />

for all x ∈U . We call these fields the coordinate functions <strong>of</strong> the chart, <strong>and</strong> the mapping ˆx is also<br />

called a coordinate map or a coordinate system on U . The set U is called the coordinate<br />

neighborhood.<br />

N<br />

N<br />

Two charts xˆ : U1<br />

→R <strong>and</strong> yˆ : U2<br />

→R , where U1∩U 2<br />

≠∅, yield the coordinate<br />

−1<br />

−1<br />

transformation yˆ xˆ : xˆ( U ˆ<br />

1∩U2) → y( U1∩U2)<br />

<strong>and</strong> its inverse xˆ yˆ : yˆ( U ˆ<br />

1∩U2) → x( U1∩U2)<br />

.<br />

Since<br />

1<br />

N<br />

( )<br />

yˆ( x) = yˆ ( x),..., yˆ<br />

( x )<br />

(44.2)<br />

The coordinate transformation can be written as the equations<br />

( ˆ 1 ˆ N<br />

) ˆ ˆ −<br />

( ),..., ( ) 1 ( ˆ 1 ( ),..., ˆ N<br />

y x y x = y x x x x ( x)<br />

)<br />

(44.3)<br />

<strong>and</strong> the inverse can be written<br />

( ˆ 1 ˆ N<br />

) ˆ ˆ −<br />

( ),..., ( ) 1 ( ˆ 1 ( ),..., ˆ N<br />

x x x x = x y y x y ( x)<br />

)<br />

(44.4)<br />

The component forms <strong>of</strong> (44.3) <strong>and</strong> (44.4) can be written in the simplified notation<br />

j j 1 N j k<br />

y = y ( x ,..., x ) ≡ y ( x )<br />

(44.5)


Sec. 44 • Coordinate Systems 307<br />

<strong>and</strong><br />

j j 1 N j k<br />

x = x ( y ,..., y ) ≡ x ( y )<br />

(44.6)<br />

1 N<br />

1 N<br />

j j<br />

j j<br />

The two N-tuples ( y ,..., y ) <strong>and</strong> ( x ,..., x ) , where y = yˆ ( x ) <strong>and</strong> x = xˆ ( x ), are the<br />

coordinates <strong>of</strong> the point x ∈U1∩U 2. Figure 6 is useful in underst<strong>and</strong>ing the coordinate<br />

transformations.<br />

N<br />

R<br />

U ∩U<br />

1 2<br />

U<br />

1<br />

U<br />

2<br />

xˆ( U ∩U<br />

)<br />

1 2<br />

ˆx<br />

ŷ<br />

yˆ( U ∩U<br />

)<br />

1 2<br />

yˆ<br />

ˆ<br />

1<br />

x −<br />

xˆ<br />

ˆ<br />

1<br />

y −<br />

Figure 6<br />

It is important <strong>to</strong> note that the quantities<br />

1 N<br />

x ,..., x ,<br />

1 N<br />

y ,..., y are scalar fields, i.e., realvalued<br />

functions defined on certain subsets <strong>of</strong> E . Since U<br />

1<br />

<strong>and</strong> U<br />

2<br />

are open sets, U1∩U 2<br />

is open,<br />

r<br />

<strong>and</strong> because ˆx <strong>and</strong> ŷ are<br />

ˆx ∩ ŷ U ∩U are open subsets <strong>of</strong><br />

C diffeomorphisms, ( U U ) <strong>and</strong> ( )<br />

N<br />

1<br />

R . In addition, the mapping yˆ xˆ<br />

−<br />

1<br />

<strong>and</strong> xˆ yˆ<br />

− are<br />

diffeomorphism, equation (44.1) written in the form<br />

1 2<br />

1 2<br />

r<br />

C diffeomorphisms. Since ˆx is a<br />

1 N<br />

ˆ( ) ( ,..., )<br />

x x = x x<br />

(44.7)<br />

can be inverted <strong>to</strong> yie1d


308 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

1<br />

( ,..., N<br />

j<br />

x x ) ( x )<br />

(44.8)<br />

x = x = x<br />

where x is a diffeomorphism<br />

x : xˆ<br />

( U)<br />

→U<br />

.<br />

r<br />

r<br />

A C -atlas on E is a family (not necessarily countable) <strong>of</strong> -<br />

where I is an index set, such that<br />

{ , x I<br />

α α<br />

α ∈ }<br />

C charts ( ˆ )<br />

U ,<br />

E<br />

=<br />

∪ Uα<br />

(44.9)<br />

α∈I<br />

Equation (44.9) states that E is covered by the family <strong>of</strong> open sets { U<br />

α<br />

α ∈ I}<br />

. A<br />

r<br />

C -Euclidean<br />

r<br />

∞<br />

manifold is a Euclidean point space equipped with a C -atlas . A C -atlas<br />

<strong>and</strong> a C ∞ -Euclidean<br />

manifold are defined similarly. For simplicity, we shall assume that E is C ∞ .<br />

A C ∞ curve in E is a C ∞ mapping λ : ( ab , ) →E , where ( ab , ) is an open interval <strong>of</strong> R .<br />

A C ∞ curve λ passes through x ∈ 0<br />

E if there exists a c∈<br />

( a , b ) such that λ( c ) = x<br />

0<br />

. Given a chart<br />

( U , xˆ<br />

) <strong>and</strong> a point x ∈U<br />

, the 0 jth coordinate curve passing through x<br />

0<br />

is the curve λ<br />

j<br />

defined by<br />

1 j− 1 j j+<br />

1 N<br />

λ<br />

j( t) = x ( x0,..., x0 , x0 + t, x0 ,..., x0<br />

)<br />

(44.10)<br />

1 j− 1 j j+<br />

1 N<br />

k<br />

for all t such that ( x ˆ<br />

0,..., x0 , x0 + t, x0 ,..., x0 ) ∈x( U ) , where ( x ˆ<br />

0) = x( x<br />

0)<br />

. The subset <strong>of</strong> U<br />

obtained by requiring<br />

x<br />

j<br />

j<br />

= xˆ ( x ) = const<br />

(44.11)<br />

is called the j th coordinate surface <strong>of</strong> the chart<br />

Euclidean manifolds possess certain special coordinate systems <strong>of</strong> major interest. Let<br />

i i be an arbitrary basis, not necessarily orthonormal, for V . We define N constant vec<strong>to</strong>r<br />

{ 1 ,..., N<br />

}<br />

j<br />

fields i : E →V , j = 1,..., N , by the formulas


Sec. 44 • Coordinate Systems 309<br />

j j<br />

i ( x) = i , x∈U (44.12)<br />

The use <strong>of</strong> the same symbol for the vec<strong>to</strong>r field <strong>and</strong> its value will cause no confusion <strong>and</strong> simplifies<br />

the notation considerably. If 0 E<br />

denotes a fixed element <strong>of</strong> E , then a Cartesian coordinate system<br />

1 2 N<br />

on E is defined by the N scalar fields zˆ , zˆ ,..., z ˆ such that<br />

z<br />

= zˆ ( x) = ( x−0E ) ⋅i , x∈E (44.13)<br />

j j j<br />

If the basis { i 1 ,..., i N<br />

} is orthonormal, the Cartesian system is called a rectangular Cartesian<br />

system. The point 0 E<br />

is called the origin <strong>of</strong> the Cartesian coordinate system. The vec<strong>to</strong>r field<br />

defined by<br />

rx ( )= x−<br />

0 E<br />

(44.14)<br />

for all<br />

If { }<br />

1 ,..., N<br />

x ∈E is the position vec<strong>to</strong>r field relative <strong>to</strong> 0 E<br />

. The value rx ( ) is the position vec<strong>to</strong>r <strong>of</strong> x .<br />

i i is the basis reciprocal <strong>to</strong> { 1 ,..., N<br />

}<br />

i i , then (44.13) implies<br />

x− 0 = x − 0 = ˆ x i = i<br />

(44.15)<br />

E<br />

1 N j j<br />

( z ,..., z )<br />

E<br />

z ( )<br />

j<br />

z<br />

j<br />

Defining constant vec<strong>to</strong>r fields i 1<br />

,..., i N<br />

as before, we can write (44.15) as<br />

ˆ j<br />

r = z i (44.16)<br />

j<br />

The product <strong>of</strong> the scalar field z ˆ j with the vec<strong>to</strong>r field i<br />

j<br />

in (44.16) is defined pointwise; i.e., if f<br />

is a scalar field <strong>and</strong> v is a vec<strong>to</strong>r field, then fv is a vec<strong>to</strong>r field defined by<br />

fvx ( ) = f( xvx ) ( )<br />

(44.17)<br />

for all x in the intersection <strong>of</strong> the domains <strong>of</strong> f <strong>and</strong> v . An equivalent version <strong>of</strong> (44.13) is<br />

ˆ j<br />

z =<br />

r⋅<br />

i j<br />

(44.18)


310 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

j<br />

where the opera<strong>to</strong>r r⋅<br />

i between vec<strong>to</strong>r fields is defined pointwise in a similar fashion as in<br />

1 2 N<br />

i , i ,..., i be another basis for V related <strong>to</strong> the original basis<br />

(44.17). As an illustration, let { }<br />

by<br />

j k<br />

= Q k<br />

j<br />

i i (44.19)<br />

<strong>and</strong> let 0 E<br />

be another fixed point <strong>of</strong> E . Then by (44.13)<br />

z<br />

= ( x−0 ) ⋅ i = ( x−0 ) ⋅ i + ( 0 −0 ) ⋅ i<br />

j j j j<br />

E E E E<br />

= Q ( x−0 ) ⋅ i + ( 0 −0 ) ⋅ i = Q z + c<br />

j k j j k j<br />

k E E E<br />

k<br />

(44.20)<br />

where (44.19) has been used. Also in (44.20) the constant scalars<br />

k<br />

c , k 1,...,<br />

= N , are defined by<br />

k<br />

k<br />

c = ( 0E −0E ) ⋅ i<br />

(44.21)<br />

1 2<br />

If the bases { , ,..., N<br />

1 2 N<br />

i i i } <strong>and</strong> { , ,..., }<br />

i i i are both orthonormal, then the matrix<br />

orthogonal. Note that the coordinate neighborhood is the entire space E .<br />

⎡<br />

⎣<br />

j<br />

Q k<br />

⎤<br />

⎦ is<br />

curve<br />

The j th coordinate curve which passes through 0 E<br />

<strong>of</strong> the Cartesian coordinate system is the<br />

λ() t = ti j<br />

+ 0 E<br />

(44.22)<br />

1 2<br />

N<br />

Equation (44.22) follows from (44.10), (44.15), <strong>and</strong> the fact that for x = 0 E<br />

, z = z =⋅⋅⋅= z = 0 .<br />

As (44.22) indicates, the coordinate curves are straight lines passing through 0 E<br />

. Similarly, the j th<br />

coordinate surface is the plane defined by


Sec. 44 • Coordinate Systems 311<br />

j<br />

( x−0 ) ⋅ i = const<br />

E<br />

3<br />

z<br />

i<br />

3<br />

0 E<br />

i<br />

2<br />

2<br />

z<br />

i<br />

1<br />

1<br />

z<br />

Figure 7<br />

Geometrically, the Cartesian coordinate system can be represented by Figure 7 for N=3. The<br />

coordinate transformation represented by (44.20) yields the result in Figure 8 (again for N=3).<br />

Since every inner product space has an orthonormal basis (see Theorem 13.3), there is no loss <strong>of</strong><br />

generality in assuming that associated with every point <strong>of</strong> E as origin we can introduce a<br />

rectangular Cartesian coordinate system.


312 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

3<br />

z<br />

3<br />

z<br />

i<br />

3<br />

r<br />

r<br />

3<br />

i<br />

1<br />

z<br />

i<br />

1<br />

0 E<br />

i<br />

2<br />

2<br />

z<br />

1<br />

z<br />

1<br />

i<br />

0 E<br />

2<br />

i<br />

2<br />

z<br />

Figure 8<br />

1 N<br />

Given any rectangular coordinate system ( zˆ<br />

,..., z ˆ ) , we can characterize a general or a<br />

curvilinear coordinate system as follows: Let ( , xˆ<br />

)<br />

U be a chart. Then it can be specified by the<br />

coordinate transformation from ẑ <strong>to</strong> ˆx as described earlier, since in this case the overlap <strong>of</strong> the<br />

coordinate neighborhood is U = E ∩U . Thus we have<br />

z z = zˆ<br />

xˆ<br />

x x<br />

1 N<br />

−1 1 N<br />

( ,..., ) ( ,..., )<br />

x x = xˆ<br />

zˆ<br />

z z<br />

1 N<br />

−1 1 N<br />

( ,..., ) ( ,..., )<br />

(44.23)<br />

−1<br />

where zˆ xˆ<br />

: xˆ( U) → zˆ( U)<br />

<strong>and</strong><br />

<strong>of</strong> (44.23) are<br />

ˆ ˆ : ˆ( ) → ˆ( U)<br />

are diffeomorphisms. Equivalent versions<br />

−1<br />

x z zU<br />

x<br />

z = z x x = z x<br />

j j 1 N j k<br />

( ,..., ) ( )<br />

x = x z z = x z<br />

j j 1 N j k<br />

( ,..., ) ( )<br />

(44.24)


Sec. 44 • Coordinate Systems 313<br />

As an example <strong>of</strong> the above ideas, consider the cylindrical coordinate system 1 . In this case<br />

N=3 <strong>and</strong> equations (44.23) take the special form<br />

1 2 3 1 2 1 2 3<br />

( z , z , z ) ( x cos x , x sin x , x )<br />

= (44.25)<br />

In order for (44.25) <strong>to</strong> qualify as a coordinate transformation, it is necessary for the transformation<br />

functions <strong>to</strong> be C ∞ 1<br />

. It is apparent from (44.25) that ẑ xˆ<br />

− is C ∞ 3<br />

on every open subset <strong>of</strong> R .<br />

1<br />

Also, by examination <strong>of</strong> (44.25), ẑ xˆ<br />

− is one-<strong>to</strong>-one if we restrict it <strong>to</strong> an appropriate domain, say<br />

3<br />

1<br />

(0, ∞ ) × (0,2 π ) × ( −∞, ∞ ) . The image <strong>of</strong> this subset <strong>of</strong> R under ẑ xˆ<br />

− is easily seen <strong>to</strong> be the set<br />

{( z , z , z ) z ≥ 0, z = 0}<br />

R 3 1 2 3 1 2 <strong>and</strong> the inverse transformation is<br />

2<br />

1 2 3 ⎛ 1 2 2 2<br />

1/2<br />

−1 z 3⎞<br />

( x , x , x ) = ⎜⎡<br />

⎣( z ) + ( z ) ⎤<br />

⎦ ,tan , z<br />

1 ⎟<br />

⎝<br />

z ⎠<br />

(44.26)<br />

which is also C ∞ . Consequently we can choose the coordinate neighborhood <strong>to</strong> be any open subset<br />

U in E such that<br />

{ }<br />

3 1 2 3 1 2<br />

zˆ( ) ⊂ ( z , z , z ) z ≥ 0, z = 0<br />

U R (44.27)<br />

or, equivalently,<br />

xˆ( U ) ⊂(0, ∞ ) × (0,2 π ) × ( −∞, ∞)<br />

1<br />

Figure 9 describes the cylindrical system. The coordinate curves are a straight line (for x ), a<br />

1 2<br />

2<br />

circle lying in a plane parallel <strong>to</strong> the ( z , z ) plane (for x ), <strong>and</strong> a straight line coincident with<br />

3<br />

1<br />

3<br />

(for x ). The coordinate surface x = const is a circular cylinder whose genera<strong>to</strong>rs are the z<br />

lines. The remaining coordinate surfaces are planes.<br />

3<br />

z<br />

1 The computer program Maple has a plot comm<strong>and</strong>, coordplot3d, that is useful when trying <strong>to</strong> visualize coordinate<br />

curves <strong>and</strong> coordinate surfaces. The program MATLAB will also produce useful <strong>and</strong> instructive plots.


314 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

z<br />

= x<br />

3 3<br />

2<br />

z<br />

1<br />

z<br />

2<br />

x<br />

Figure 9<br />

1<br />

x<br />

Returning <strong>to</strong> the general transformations (44.5) <strong>and</strong> (44.6), we can substitute the second in<strong>to</strong><br />

the first <strong>and</strong> differentiate the result <strong>to</strong> find<br />

∂y<br />

∂x<br />

∂x<br />

∂y<br />

i<br />

k<br />

1 N 1 N i<br />

( x ,..., x ) ( y ,..., y ) = δ<br />

k<br />

j<br />

j<br />

(44.28)<br />

By a similar argument with x <strong>and</strong> y interchanged,<br />

∂y<br />

∂x<br />

∂x<br />

∂y<br />

k<br />

i<br />

1 N 1 N i<br />

( x ,..., x ) ( y ,..., y ) = δ<br />

j<br />

k<br />

j<br />

(44.29)<br />

Each <strong>of</strong> these equations ensures that<br />

⎡∂y<br />

⎤<br />

i<br />

1 N<br />

det ⎢ ( x ,..., x ) 0<br />

k<br />

j<br />

x<br />

⎥ = ≠<br />

⎣∂ ⎦ ⎡∂x<br />

1 N ⎤<br />

det ( y ,..., y )<br />

l<br />

⎢<br />

⎣∂y<br />

1<br />

⎥<br />

⎦<br />

(44.30)<br />

For example, ( j k<br />

⎡<br />

⎣ )( )<br />

j k 1 N<br />

determinant det ⎡( ∂x ∂y )( y ,..., y ) ⎤<br />

⎤<br />

⎦<br />

1 2 3 1<br />

det ∂z ∂ x x , x , x = x ≠0<br />

⎣ ⎦<br />

for the cylindrical coordinate system. The<br />

is the Jacobian <strong>of</strong> the coordinate transformation (44.6).


Sec. 44 • Coordinate Systems 315<br />

Just as a vec<strong>to</strong>r space can be assigned an orientation, a Euclidean manifold can be oriented<br />

by assigning the orientation <strong>to</strong> its translation space V . In this case E is called an oriented<br />

Euclidean manifold. In such a manifold we use Cartesian coordinate systems associated with<br />

positive basis only <strong>and</strong> these coordinate systems are called positive. A curvilinear coordinate<br />

system is positive if its coordinate transformation relative <strong>to</strong> a positive Cartesian coordinate system<br />

has a positive Jacobian.<br />

Given a chart ( , xˆ<br />

)<br />

<strong>and</strong> obtain a C ∞<br />

U for E , we can compute the gradient <strong>of</strong> each coordinate function x ˆi<br />

vec<strong>to</strong>r field on U . We shall denote each <strong>of</strong> these fields by<br />

i<br />

g , namely<br />

i<br />

i<br />

g = grad xˆ<br />

(44.31)<br />

i<br />

for i = 1,..., N. From this definition, it is clear that g ( x ) is a vec<strong>to</strong>r in V normal <strong>to</strong> the i th<br />

coordinate surface. From (44.8) we can define N vec<strong>to</strong>r fields g ,..., 1<br />

g<br />

N<br />

on U by<br />

[ ]<br />

g = grad x (0,..., 1,...,0)<br />

(44.32)<br />

i<br />

i<br />

Or, equivalently,<br />

g<br />

i<br />

<br />

= ≡<br />

t<br />

∂ x<br />

1 i N 1 N<br />

x( x ,..., x + t,..., x ) −x( x ,..., x ) ∂x<br />

1 N<br />

lim ( x ,..., x )<br />

t→0<br />

i<br />

(44.33)<br />

for all<br />

Since<br />

x ∈U . Equations (44.10) <strong>and</strong> (44.33) show that g ( x ) is tangent <strong>to</strong> the i th coordinate curve.<br />

i<br />

1<br />

xˆ i ( ( x ,..., x N )) = x<br />

i<br />

x (44.34)<br />

The chain rule along with the definitions (44.31) <strong>and</strong> (44.32) yield<br />

i<br />

i<br />

g ( x) ⋅ g ( x ) = δ<br />

(44.35)<br />

j<br />

j<br />

as they should.


316 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

The values <strong>of</strong> the vec<strong>to</strong>r fields { g ,..., 1<br />

g<br />

N } form a linearly independent set <strong>of</strong> vec<strong>to</strong>rs<br />

{ g x g x } at each x ∈U . To see this assertion, assume x ∈U <strong>and</strong><br />

( ),..., ( )<br />

1 N<br />

N<br />

λ g ( x) + λ g ( x) +⋅⋅⋅+ λ g ( x)<br />

= 0<br />

1 2<br />

1 2<br />

N<br />

for<br />

1 N<br />

j<br />

λ ,..., λ ∈R . Taking the inner product <strong>of</strong> this equation with g ( x ) <strong>and</strong> using equation (44.35)<br />

j<br />

, we see that λ = 0 , j = 1,..., N which proves the assertion.<br />

Because V has dimension N, { g ( ),..., ( )<br />

1<br />

x g<br />

N<br />

x } forms a basis for V at each ∈<br />

Equation (44.35) shows that { g 1 ( x),..., g N ( x )}<br />

is the basis reciprocal <strong>to</strong> { ( ),..., ( )<br />

1 N }<br />

<strong>of</strong> the special geometric interpretation <strong>of</strong> the vec<strong>to</strong>rs { }<br />

N<br />

x U .<br />

g x g x . Because<br />

g ( ),..., ( )<br />

1<br />

x g<br />

N<br />

x <strong>and</strong> { g 1 ( x ),..., g ( x ) }<br />

mentioned above, these bases are called the natural bases <strong>of</strong> ˆx at x . Any other basis field which<br />

cannot be determined by either (44.31) or (44.32) relative <strong>to</strong> any coordinate system is called an<br />

anholonomic or nonintegrable basis. The constant vec<strong>to</strong>r fields { i ,..., 1<br />

i<br />

N } <strong>and</strong> { i 1 ,..., i N<br />

} yield the<br />

natural bases for the Cartesian coordinate systems.<br />

U are two charts such that U1∩U 2<br />

≠∅, we can determine the<br />

transformation rules for the changes <strong>of</strong> natural bases at x ∈U1∩<br />

U<br />

2<br />

in the following way: We shall<br />

j<br />

let the vec<strong>to</strong>r fields h , j = 1,..., N , be defined by<br />

If ( U , x ) <strong>and</strong> ( , y )<br />

1 ˆ<br />

2<br />

ˆ<br />

j<br />

j<br />

h = grad yˆ<br />

(44.36)<br />

Then, from (44.5) <strong>and</strong> (44.31),<br />

∂y<br />

h x = x = x<br />

∂ x<br />

j<br />

∂y<br />

1 N i<br />

= ( x ,..., x ) g ( x)<br />

∂<br />

i<br />

x<br />

j<br />

j j 1 N i<br />

( ) grad yˆ<br />

( ) ( x ,..., x )grad xˆ<br />

( )<br />

i<br />

(44.37)<br />

for all x ∈U1∩U 2. A similar calculation shows that


Sec. 44 • Coordinate Systems 317<br />

i<br />

∂x<br />

1 N<br />

h<br />

j( x) = ( y ,..., y ) gi( x )<br />

(44.38)<br />

∂<br />

j<br />

y<br />

for all x ∈U1∩U 2. Equations (44.37) <strong>and</strong> (44.38) are the desired transformations.<br />

Given a chart ( , x )<br />

U , we can define 2<br />

ij<br />

2N scalar fields g : U →R <strong>and</strong> g : U<br />

1 ˆ<br />

ij<br />

→R by<br />

g ( x) = g ( x) ⋅ g ( x) = g ( x )<br />

(44.39)<br />

ij i j ji<br />

<strong>and</strong><br />

ij i j ji<br />

g ( x) = g ( x) ⋅ g ( x) = g ( x )<br />

(44.40)<br />

for all<br />

x ∈U . It immediately follows from (44.35) that<br />

−1<br />

ij<br />

⎣<br />

⎡g ( x) ⎦<br />

⎤ = ⎡ ⎣ g ij<br />

( x ) ⎤ ⎦<br />

(44.41)<br />

since we have<br />

i ij<br />

g = g g<br />

j<br />

(44.42)<br />

<strong>and</strong><br />

i<br />

g<br />

j<br />

= g<br />

jig (44.43)<br />

where the produce <strong>of</strong> the scalar fields with vec<strong>to</strong>r fields is defined by (44.17). If θ<br />

ij<br />

is the angle<br />

between the i th <strong>and</strong> the j th coordinate curves at x ∈U , then from (44.39)<br />

gij( x)<br />

cos θ<br />

ij<br />

=<br />

⎡ ⎣gii( x) g<br />

jj( x)<br />

⎤ ⎦<br />

1/2<br />

(no sum)<br />

(44.44)


318 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

Based upon (44.44), the curvilinear coordinate system is orthogonal if g<br />

ij<br />

= 0 when i ≠ j. The<br />

symbol g denotes a scalar field on U defined by<br />

g( x) = det ⎡<br />

⎣g ij<br />

( x ) ⎤<br />

⎦<br />

(44.45)<br />

for all<br />

x ∈U .<br />

At the point<br />

x ∈U , the differential element <strong>of</strong> arc ds is defined by<br />

2<br />

ds = d ⋅ d<br />

x x (44.46)<br />

<strong>and</strong>, by (44.8), (44.33), <strong>and</strong> (44.39),<br />

∂x<br />

∂x<br />

ˆ<br />

∂x<br />

∂x<br />

i j<br />

= g ( x)<br />

dx dx<br />

2<br />

i j<br />

ds = ( x( x)) ⋅ ( x( x))<br />

dx dx<br />

i<br />

j<br />

ij<br />

ˆ<br />

(44.47)<br />

If ( U , x ) <strong>and</strong> ( , y )<br />

1 ˆ<br />

U are charts where U1∩U 2<br />

≠∅, then at x ∈U1∩<br />

U<br />

2<br />

2<br />

ˆ<br />

h ( x) = h ( x) ⋅h ( x)<br />

ij i j<br />

∂x<br />

∂x<br />

= ∂ y<br />

∂ y<br />

k<br />

l<br />

1 N<br />

1 N<br />

( y ,..., y ) ( y ,..., y ) gkl<br />

( x)<br />

i<br />

j<br />

(44.48)<br />

Equation (44.48) is helpful for actual calculations <strong>of</strong> the quantities g ( x ) . For example, for<br />

the transformation (44.23), (44.48) can be arranged <strong>to</strong> yield<br />

ij<br />

k<br />

k<br />

∂z<br />

1 N ∂z<br />

1 N<br />

gij( x ) = ( x ,..., x ) ( x ,..., x )<br />

(44.49)<br />

∂<br />

i j<br />

x ∂ x<br />

since k l kl<br />

⋅ =δ i i . For the cylindrical coordinate system defined by (44.25) a simple calculation<br />

based upon (44.49) yields


Sec. 44 • Coordinate Systems 319<br />

⎡1 0 0⎤<br />

1 2<br />

⎡g<br />

( )<br />

⎢<br />

ij<br />

0 ( x ) 0<br />

⎥<br />

⎣ x ⎤<br />

⎦ =<br />

(44.50)<br />

⎢ ⎥<br />

⎢⎣<br />

0 0 1⎥⎦<br />

Among other things, (44.50) shows that this coordinate system is orthogonal. By (44.50) <strong>and</strong><br />

(44.41)<br />

⎡1 0 0⎤<br />

ij<br />

1 2<br />

⎡g<br />

( )<br />

⎢<br />

0 1 ( x ) 0<br />

⎥<br />

⎣ x ⎤<br />

⎦ =<br />

(44.51)<br />

⎢ ⎥<br />

⎢⎣<br />

0 0 1⎥⎦<br />

And, from (44.50) <strong>and</strong> (44.45),<br />

g( ) ( )<br />

1 2<br />

x = x<br />

(44.52)<br />

It follows from (44.50) <strong>and</strong> (44.47) that<br />

2 1 2 1 2 2 2 3 2<br />

ds = ( dx ) + ( x ) ( dx ) + ( dx )<br />

(44.53)<br />

Exercises<br />

44.1 Show that<br />

i<br />

∂x<br />

k<br />

= <strong>and</strong> i = grad ˆ<br />

∂z<br />

k<br />

k<br />

z<br />

k<br />

for any Cartesian coordinate system ẑ associated with { i j<br />

}.<br />

44.2 Show that<br />

I= grad r( x )


320 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

44.3 Spherical coordinates<br />

1 2 3<br />

( x , x , x ) are defined by the coordinate transformation<br />

z = x sin x cos x<br />

1 1 2 3<br />

z = x sin x sin x<br />

2 1 2 3<br />

z = x cos x<br />

3 1 2<br />

1 2 3<br />

relative <strong>to</strong> a rectangular Cartesian coordinate system ẑ . How must the quantity ( x , x , x )<br />

1<br />

be restricted so as <strong>to</strong> make ẑ xˆ<br />

− one-<strong>to</strong>-one? Discuss the coordinate curves <strong>and</strong> the<br />

coordinate surfaces. Show that<br />

⎡1 0 0 ⎤<br />

1 2<br />

⎡g<br />

( )<br />

⎢<br />

ij<br />

0 ( x ) 0<br />

⎥<br />

⎣ x ⎤<br />

⎦ =<br />

⎢ ⎥<br />

1 2 2<br />

⎢⎣<br />

0 0 ( x sin x ) ⎥⎦<br />

44.4 Paraboloidal coordinates<br />

1 2 3<br />

( x , x , x ) are defined by<br />

z = x x cos x<br />

1 1 2 3<br />

z = x x sin x<br />

2 1 2 3<br />

1<br />

z = ( x ) − ( x )<br />

2<br />

( )<br />

3 1 2 2 2<br />

1 2 3<br />

Relative <strong>to</strong> a rectangular Cartesian coordinate system ẑ . How must the quantity ( x , x , x )<br />

1<br />

be restricted so as <strong>to</strong> make ẑ xˆ<br />

− one-<strong>to</strong>-one. Discuss the coordinate curves <strong>and</strong> the<br />

coordinate surfaces. Show that<br />

1 2 2 2<br />

⎡( x ) + ( x ) 0 0 ⎤<br />

⎢<br />

1 2 2 2 ⎥<br />

⎡<br />

⎣gij( x ) ⎤<br />

⎦ =<br />

⎢<br />

0 ( x ) + ( x ) 0<br />

⎥<br />

1 2 2<br />

⎢ 0 0 ( xx)<br />

⎥<br />

⎣<br />

⎦<br />

44.5 A bispherical coordinate system<br />

coordinate system by<br />

1 2 3<br />

( x , x , x ) is defined relative <strong>to</strong> a rectangular Cartesian


Sec. 44 • Coordinate Systems 321<br />

z<br />

z<br />

1<br />

2<br />

2 3<br />

asin<br />

x cos x<br />

=<br />

cosh x − cos x<br />

2 3<br />

asin<br />

x sin x<br />

=<br />

cosh x − cos x<br />

1 2<br />

1 2<br />

<strong>and</strong><br />

z<br />

3<br />

1<br />

asinh<br />

x<br />

=<br />

cosh x − cos x<br />

1 2<br />

1 2 3<br />

1<br />

where a > 0 . How must ( x , x , x ) be restricted so as <strong>to</strong> make ẑ xˆ<br />

− one-<strong>to</strong>-one?<br />

Discuss the coordinate curves <strong>and</strong> the coordinate surfaces. Also show that<br />

⎡<br />

⎢<br />

⎢<br />

⎢<br />

⎢<br />

a<br />

1 2<br />

( cosh x − cos x )<br />

2<br />

2<br />

0 0<br />

2<br />

⎡gij( x ) ⎤ = ⎢ 0<br />

2<br />

0 ⎥<br />

⎣<br />

⎦<br />

⎢<br />

⎢<br />

⎢<br />

⎢<br />

⎣<br />

a<br />

1 2<br />

( cosh x − cos x )<br />

0 0<br />

2 2 2<br />

a (sin x )<br />

1 2<br />

( cosh x − cos x )<br />

2<br />

⎤<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎦<br />

44.6 Prolate spheroidal coordinates<br />

1 2 3<br />

( x , x , x ) are defined by<br />

z = asinh x sin x cos x<br />

1 1 2 3<br />

z = asinh x sin x sin x<br />

2 1 2 3<br />

z = acosh<br />

x cos x<br />

3 1 2<br />

relative <strong>to</strong> a rectangular Cartesian coordinate system ẑ , where a > 0 . How must<br />

1 2 3<br />

( x , x , x )<br />

1<br />

be restricted so as <strong>to</strong> make ẑ xˆ<br />

− one-<strong>to</strong>-one? Also discuss the coordinate curves <strong>and</strong> the<br />

coordinate surfaces <strong>and</strong> show that


322 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

( − )<br />

2 2 1 2 2<br />

⎡a cosh x cos x<br />

0 0 ⎤<br />

⎢<br />

⎥<br />

2 2 1 2 2<br />

⎡<br />

⎣gij( x ) ⎤<br />

⎦ = ⎢ 0 a ( cosh x −cos x ) 0 ⎥<br />

⎢<br />

⎥<br />

2 2 1 2 2<br />

⎢ 0 0 a sinh x sin x ⎥<br />

⎣<br />

⎦<br />

44.7 Elliptical cylindrical coordinates<br />

coordinate system by<br />

1 2 3<br />

( x , x , x ) are defined relative <strong>to</strong> a rectangular Cartesian<br />

z = acosh<br />

x cos x<br />

1 1 2<br />

z = asinh<br />

x sin x<br />

z<br />

2 1 2<br />

= x<br />

3 3<br />

1 2 3<br />

1<br />

where a > 0 . How must ( x , x , x ) be restricted so as <strong>to</strong> make ẑ xˆ<br />

− one-<strong>to</strong>-one?<br />

Discuss the coordinate curves <strong>and</strong> coordinate surfaces. Also, show that<br />

( + )<br />

2 2 1 2 2<br />

⎡a sinh x sin x<br />

0 0⎤<br />

⎢<br />

⎥<br />

2 2 1 2 2<br />

⎡<br />

⎣gij( x ) ⎤<br />

⎦ = ⎢ 0 a ( sinh x + sin x ) 0⎥<br />

⎢<br />

⎥<br />

⎢ 0 0 1⎥<br />

⎣<br />

⎦<br />

44.8 For the cylindrical coordinate system show that<br />

g = (cos x ) i + (sin x ) i<br />

2 2<br />

1 1 2<br />

g =− x (sin x ) i + x (cos x ) i<br />

g<br />

1 2 1 2<br />

2 1 2<br />

= i<br />

3 3<br />

44.9 At a point x in E , the components <strong>of</strong> the position vec<strong>to</strong>r rx ( ) = x−<br />

0 E<br />

with respect <strong>to</strong> the<br />

basis { i ,..., 1<br />

i<br />

N } associated with a rectangular Cartesian coordinate system are<br />

1 N<br />

z ,..., z .<br />

This observation follows, <strong>of</strong> course, from (44.16). Compute the components <strong>of</strong> rx ( ) with<br />

respect <strong>to</strong> the basis { g1( x), g2( x), g3( x )}<br />

for (a) cylindrical coordinates, (b) spherical<br />

coordinates, <strong>and</strong> (c) parabolic coordinates. You should find that


Sec. 44 • Coordinate Systems 323<br />

rx ( ) = x g( x) + x g( x) for (a)<br />

rx<br />

1 3<br />

1 3<br />

1<br />

( ) = x g1( x) for (b)<br />

1 1 1 2<br />

rx ( ) = x g1( x) + x g2( x) for (c)<br />

2 2<br />

44.10 Toroidal coordinates<br />

system by<br />

1 2 3<br />

( x , x , x ) are defined relative <strong>to</strong> a rectangular Cartesian coordinate<br />

z<br />

z<br />

1<br />

2<br />

1 3<br />

asinh<br />

x cos x<br />

=<br />

1 2<br />

cosh x − cos x<br />

1 3<br />

asinh<br />

x sin x<br />

=<br />

1 2<br />

cosh x − cos x<br />

<strong>and</strong><br />

z<br />

3<br />

2<br />

asin<br />

x<br />

=<br />

cosh x − cos x<br />

1 2<br />

1 2 3<br />

where a > 0 . How must ( x , x , x ) be restricted so as <strong>to</strong> make<br />

the coordinate surfaces. Show that<br />

1<br />

ẑ xˆ<br />

− one <strong>to</strong> one? Discuss<br />

⎡<br />

⎢<br />

⎢<br />

⎢<br />

⎢<br />

a<br />

1 2<br />

( cosh x − cos x )<br />

2<br />

2<br />

0 0<br />

2<br />

⎡gij( x ) ⎤ = ⎢ 0<br />

2<br />

0 ⎥<br />

⎣<br />

⎦<br />

⎢<br />

⎢<br />

⎢<br />

⎢<br />

⎣<br />

a<br />

1 2<br />

( cosh x − cos x )<br />

0 0<br />

a<br />

sinh<br />

x<br />

2 2 2<br />

1 2<br />

( cos x − cos x )<br />

2<br />

⎤<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />

⎥<br />


324 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

Section 45. Transformation Rules for Vec<strong>to</strong>rs <strong>and</strong> Tensor Fields<br />

In this section, we shall formalize certain ideas regarding fields on E <strong>and</strong> then investigate<br />

the transformation rules for vec<strong>to</strong>rs <strong>and</strong> tensor fields. Let U be an open subset <strong>of</strong> E ; we shall<br />

denote by F ∞ ( U ) the set <strong>of</strong> C ∞ functions f : U →R. First we shall study the algebraic structure<br />

<strong>of</strong> F ∞ ( U ) . If f <strong>and</strong> 1<br />

f<br />

2<br />

are in F ∞ ( U ) , then their sum f1+ f2<br />

is an element <strong>of</strong> F ∞ ( U ) defined by<br />

( f + f )( x) = f ( x) + f ( x )<br />

(45.1)<br />

1 2 1 2<br />

<strong>and</strong> their produce f1f 2<br />

is also an element <strong>of</strong> F ∞ ( U ) defined by<br />

( ff)( x) = f( x) f( x )<br />

(45.2)<br />

1 2 1 2<br />

for all<br />

x ∈U . For any real number λ ∈R the constant function is defined by<br />

λ( x ) = λ<br />

(45.3)<br />

for all x ∈U . For simplicity, the function <strong>and</strong> the value in (45.3) are indicated by the same<br />

symbol. Thus, the zero function in F ∞ ( U ) is denoted simply by 0 <strong>and</strong> for every f ∈ F ∞ ( U )<br />

f<br />

+ 0= f<br />

(45.4)<br />

It is also apparent that<br />

1f<br />

= f<br />

(45.5)<br />

In addition, we define<br />

− f = ( − 1) f<br />

(45.6)


Sec. 45 • Transformation Rules 325<br />

It is easily shown that the operations <strong>of</strong> addition <strong>and</strong> multiplication obey commutative, associative,<br />

<strong>and</strong> distributive laws. These facts show that F ∞ ( U ) is a commutative ring (see Section 7).<br />

An important collection <strong>of</strong> scalar fields can be constructed as follows: Given two charts<br />

2<br />

U ˆ<br />

2, y , where U1∩U 2<br />

≠∅, we define the<br />

∂y ∂ x x 1 ,..., x<br />

1 N<br />

2<br />

x ,..., x ∈xˆ<br />

U ∩U . Using a suggestive notation, we can define N C ∞ functions<br />

( U , x ) <strong>and</strong> ( )<br />

1 ˆ<br />

at every ( ) ( )<br />

i j<br />

∂y<br />

∂x<br />

: U ∩U →R by<br />

1 2<br />

1 2<br />

N partial derivatives ( i j )( N<br />

)<br />

∂y<br />

∂x<br />

i<br />

j<br />

i<br />

∂y<br />

( x) = xˆ<br />

( x)<br />

(45.7)<br />

j<br />

∂x<br />

for all x ∈U1∩U 2.<br />

As mentioned earlier, a C ∞ vec<strong>to</strong>r field on an open set U <strong>of</strong> E is a C ∞ map v : U →V ,<br />

where V is the translation space <strong>of</strong> E . The fields defined by (44.31) <strong>and</strong> (44.32) are special cases<br />

<strong>of</strong> vec<strong>to</strong>r fields. We can express v in component forms on U1∩U 2,<br />

i<br />

j<br />

v = υ gi<br />

= υ<br />

jg (45.8)<br />

υ<br />

i<br />

j<br />

= g<br />

jiυ<br />

(45.9)<br />

i<br />

As usual, we can computer υ : U ∩U1<br />

→R by<br />

i<br />

i<br />

υ ( x) = v( x) ⋅g ( x),<br />

x∈U ∩U (45.10)<br />

1 2<br />

<strong>and</strong><br />

i<br />

component form <strong>of</strong><br />

υ by (45.9). In particular, if ( , y)<br />

U is another chart such that U1∩U 2<br />

≠∅, then the<br />

2<br />

ˆ<br />

h<br />

j<br />

relative <strong>to</strong> ˆx is<br />

i<br />

∂x<br />

h<br />

j<br />

= g<br />

j i<br />

(45.11)<br />

∂ y


326 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

With respect <strong>to</strong> the chart ( U , y)<br />

2<br />

ˆ<br />

, we have also<br />

k<br />

v=υ<br />

h<br />

k<br />

(45.12)<br />

k<br />

where υ : U ∩U2<br />

→R. From (45.12), (45.11), <strong>and</strong> (45.8), the transformation rule for the<br />

components <strong>of</strong> v relative <strong>to</strong> the two charts ( U , x ) <strong>and</strong> ( , y )<br />

1 ˆ<br />

U is<br />

2<br />

ˆ<br />

υ<br />

∂x<br />

y<br />

i<br />

i<br />

j<br />

= υ<br />

(45.13)<br />

∂<br />

j<br />

for all x ∈U1∩U2∩U .<br />

As in (44.18), we can define an inner product operation between vec<strong>to</strong>r fields. If<br />

v1: U1<br />

→V <strong>and</strong> v2: U2<br />

→V are vec<strong>to</strong>r fields, then v1⋅<br />

v<br />

2<br />

is a scalar field defined on U1∩U 2<br />

by<br />

v ⋅ v ( x) = v ( x) ⋅v ( x),<br />

x∈U ∩U (45.14)<br />

1 2 1 2 1 2<br />

Then (45.10) can be written<br />

i<br />

υ =<br />

i<br />

v ⋅ g (45.15)<br />

Now let us consider tensor fields in general. Let T ∞ q<br />

( U ) denote the set <strong>of</strong> all tensor fields <strong>of</strong><br />

order q defined on an open set U in E . As with the set F ∞ ( U ) , the set T ∞ q<br />

( U ) can be assigned<br />

an algebraic structure. The sum <strong>of</strong> A : U → Tq( V ) <strong>and</strong> B : U → Tq( V ) is a C ∞ tensor field<br />

A+ B : U →T ( V ) defined by<br />

q<br />

( A+ B)( x) = A( x) + B ( x)<br />

(45.16)<br />

for all<br />

x∈U . If f ∈ F ∞ ( U ) <strong>and</strong> A ∈T ∞ ( U ), then we can define fA ∈T ∞ ( U ) by<br />

q<br />

q


Sec. 45 • Transformation Rules 327<br />

fA( x) = f( x) A ( x)<br />

(45.17)<br />

Clearly this multiplication operation satisfies the usual associative <strong>and</strong> distributive laws with<br />

respect <strong>to</strong> the sum for all x∈U . As with F ∞ ( U ) , constant tensor fields in T ∞ ( U ) are given the<br />

same symbol as their value. For example, the zero tensor field is 0 : U → T ( V ) <strong>and</strong> is defined by<br />

q<br />

q<br />

0 ( x)=<br />

0<br />

(45.18)<br />

for all<br />

x∈U . If 1 is the constant function in F ∞ ( U ) , then<br />

− A = ( −1)<br />

A (45.19)<br />

The algebraic structure on the set T ∞ q<br />

( V ) just defined is called a module over the ring F ∞ ( U ) .<br />

scalar fields<br />

The components <strong>of</strong> a tensor field : U → T ( V )<br />

A : U ∩U →R defined by<br />

i1 ... i 1<br />

q<br />

A with respect <strong>to</strong> a chart ( , x )<br />

q<br />

U are the q<br />

N<br />

1 ˆ<br />

A<br />

...<br />

( x) = A ( x)( g ( x),..., g ( x))<br />

(45.20)<br />

i1 iq<br />

i1<br />

iq<br />

for all x∈U ∩U 1<br />

. Clearly we can regard tensor fields as multilinear mappings on vec<strong>to</strong>r fields<br />

with values as scalar fields. For example, A ( g ,..., g ) is a scalar field defined by<br />

i1<br />

i q<br />

A ( g ,..., g )( x ) = A ( x )( g ( x ),..., g ( x ))<br />

i1 iq<br />

i1<br />

iq<br />

for all x∈U ∩U 1<br />

. In fact we can, <strong>and</strong> shall, carry over <strong>to</strong> tensor fields the many algebraic<br />

operations previously applied <strong>to</strong> tensors. In particular a tensor field A : U → T ( V ) has the<br />

representation<br />

q<br />

i<br />

i<br />

1<br />

q<br />

A i 1 ... i<br />

g g<br />

q<br />

A = ⊗⋅⋅⋅⊗<br />

(45.21)


328 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

U is the coordinate neighborhood for a chart ( )<br />

for all x∈U ∩U 1<br />

, where U , x<br />

1<br />

. The scalar fields<br />

A are the covariant components <strong>of</strong> A <strong>and</strong> under a change <strong>of</strong> coordinates obey the<br />

i1 ... iq<br />

transformation rule<br />

1 ˆ<br />

A<br />

i1<br />

∂x<br />

∂x<br />

= ⋅⋅⋅<br />

∂y<br />

∂y<br />

iq<br />

A<br />

k1... kq<br />

k1<br />

kq<br />

i1...<br />

iq<br />

(45.22)<br />

Equation (45.22) is a relationship among the component fields <strong>and</strong> holds at all points<br />

x∈U1∩U2∩U where the charts involved are ( U , xˆ<br />

1 ) <strong>and</strong> ( U ˆ ) 2, y . We encountered an example <strong>of</strong><br />

(45.22) earlier with (44.48). Equation (44.48) shows that the g<br />

ij<br />

are the covariant components <strong>of</strong> a<br />

tensor field I whose value is the identity or metric tensor, namely<br />

I = g g ⊗ g = g ⊗ g = g ⊗ g = g g ⊗g (45.23)<br />

i j j j ij<br />

ij j j i j<br />

for points ∈<br />

U , x<br />

1<br />

. Equations (45.23) show that the<br />

components <strong>of</strong> a constant tensor field are not necessarily constant scalar fields. It is only in<br />

Cartesian coordinates that constant tensor fields have constant components.<br />

x U , where the chart in question is ( )<br />

1 ˆ<br />

Another important tensor field is the one constructed from the positive unit volume tensor<br />

i , which has positive orientation, E is given by (41.6),<br />

E . With respect <strong>to</strong> an orthonormal basis { j }<br />

i.e.,<br />

E = i ∧⋅⋅⋅∧ i = i ⊗⋅⋅⋅⊗i<br />

(45.24)<br />

1 N<br />

εi 1 ⋅⋅⋅ iN i1<br />

iN<br />

Given this tensor, we define as usual a constant tensor field E : E → Tˆ<br />

( V ) by<br />

N<br />

E( x)=<br />

E (45.25)<br />

for all<br />

x ∈E . With respect <strong>to</strong> a chart ( U , x ) , it follows from the general formula (42.27) that<br />

1 ˆ<br />

E = E ⊗⋅⋅⋅⊗ = ⊗⋅⋅⋅⊗<br />

(45.26)<br />

i1<br />

iN<br />

i1<br />

iN<br />

i1 i<br />

E ⋅⋅⋅<br />

⋅⋅⋅<br />

g g g<br />

N<br />

i<br />

g<br />

1<br />

iN


Sec. 45 • Transformation Rules 329<br />

where<br />

i1 iN<br />

E ⋅⋅⋅<br />

<strong>and</strong> E ⋅⋅⋅ are scalar fields on U<br />

1<br />

defined by<br />

i1 iN<br />

E<br />

= e gε<br />

(45.27)<br />

i1⋅⋅⋅iN<br />

i1⋅⋅⋅iN<br />

<strong>and</strong><br />

E<br />

e<br />

=<br />

g ε<br />

(45.28)<br />

i1⋅⋅⋅iN<br />

i1⋅⋅⋅iN<br />

where, as in Section 42, e is 1<br />

+ if { ( )}<br />

g x is positively oriented <strong>and</strong> 1<br />

i<br />

− if { ( )}<br />

g x is negatively<br />

oriented, <strong>and</strong> where g is the determinant <strong>of</strong> ⎡<br />

⎣gij<br />

⎤<br />

⎦ as defined by (44.45). By application <strong>of</strong> (42.28),<br />

it follows that<br />

i<br />

i1⋅⋅⋅iN ij 1 1 iN jN<br />

E g g E j 1⋅⋅⋅<br />

j N<br />

= ⋅⋅⋅ (45.29)<br />

An interesting application <strong>of</strong> the formulas derived thus far is the derivation <strong>of</strong> an expression<br />

for the differential element <strong>of</strong> volume in curvilinear coordinates. Given the position vec<strong>to</strong>r r<br />

U , x , the differential <strong>of</strong> r can be written<br />

defined by (44.14) <strong>and</strong> a chart ( )<br />

1 ˆ<br />

i<br />

dr = dx = g ( x ) dx<br />

(45.30)<br />

i<br />

where (44.33) has been used. Given N differentials <strong>of</strong> r , dr1, dr2, ⋅⋅⋅, dr N<br />

, the differential volume<br />

element dυ generated by them is defined by<br />

( r ⋅⋅⋅ r )<br />

dυ = E d , , 1<br />

d<br />

N<br />

(45.31)<br />

or, equivalently,<br />

dυ = dr ∧⋅⋅⋅∧dr (45.32)<br />

1 N


330 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

If we select<br />

N<br />

dr = g ( x) dx , dr = g ( x) dx , ⋅⋅⋅ , dr = g ( x ) dx , we can write (45.31)) as<br />

1 2<br />

1 1 2 2<br />

N<br />

N<br />

1 2<br />

( g x g x )<br />

( ), , ( ) N<br />

dυ = E<br />

1<br />

⋅⋅⋅<br />

N<br />

dx dx ⋅⋅⋅dx<br />

(45.33)<br />

By use <strong>of</strong> (45.26) <strong>and</strong> (45.27), we then get<br />

( g ( x), , g ( x)<br />

)<br />

E ⋅⋅⋅ = E = e g<br />

(45.34)<br />

1 N<br />

12⋅⋅⋅N<br />

Therefore,<br />

1 2 N<br />

dυ = gdx dx ⋅⋅⋅ dx<br />

(45.35)<br />

For example, in the parabolic coordinates mentioned in Exercise 44.4,<br />

(( ) ( ) )<br />

2 2<br />

d x x x x dx dx dx<br />

1 2 1 2 1 2 3<br />

υ = + (45.36)<br />

Exercises<br />

45.1 Let v be a C ∞ vec<strong>to</strong>r field <strong>and</strong> f be a C ∞ function both defined on U an open set in E .<br />

We define v f : U →R by<br />

v f( x) = v( x) ⋅grad f( x),<br />

x∈U (45.37)<br />

Show that<br />

( λf + μg) = λ( f ) + μ( g)<br />

v v v<br />

<strong>and</strong>


Sec. 45 • Transformation Rules 331<br />

( fg) = ( f ) g+<br />

f ( g)<br />

v v v<br />

For all constant functions λ , μ <strong>and</strong> all C ∞ functions f <strong>and</strong> g . In differential geometry,<br />

an opera<strong>to</strong>r on F ∞ ( U ) with the above properties is called a derivation. Show that,<br />

conversely, every derivation on F ∞ ( U ) corresponds <strong>to</strong> a unique vec<strong>to</strong>r field on U by<br />

(45.37).<br />

45.2 By use <strong>of</strong> the definition (45.37), the Lie bracket <strong>of</strong> two vec<strong>to</strong>r fields v : U →V <strong>and</strong><br />

: →<br />

uv, , is a vec<strong>to</strong>r field defined by<br />

u U V , written [ ]<br />

[ uv , ] f = u ( v f ) −v ( u f )<br />

(45.38)<br />

For all scalar fields f ∈ F ∞ ( U ) . Show that [ , ]<br />

defines a derivation on [ uv. , ] Also, show that<br />

uv is well defined by verifying that (45.38)<br />

[ , ] = ( grad ) −( grad )<br />

uv vu u v<br />

<strong>and</strong> then establish the following results:<br />

(a) [ uv , ] =−[ vu , ]<br />

⎡⎣v u w ⎤⎦+ ⎡⎣u w v ⎤⎦+ ⎡⎣w v u ⎤⎦<br />

= 0<br />

(b) ,[ , ] ,[ , ] ,[ , ]<br />

for all vec<strong>to</strong>r fields uv , <strong>and</strong> w .<br />

ˆ 1<br />

g<br />

i<br />

. Show that ⎡<br />

⎣gi,<br />

g ⎤<br />

j⎦<br />

= 0 .<br />

The results (a) <strong>and</strong> (b) are known as Jacobi’s identities.<br />

45.3 In a three-dimensional Euclidean space the differential element <strong>of</strong> area normal <strong>to</strong> the plan<br />

formed from dr<br />

1<br />

<strong>and</strong> dr<br />

2<br />

is defined by<br />

(c) Let ( U , x ) be a chart with natural basis field { }<br />

dσ = dr × dr<br />

1 2<br />

Show that<br />

j k i<br />

dσ<br />

= E dx1dx2 g ( x )<br />

ijk


332 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

Section 46. Anholonomic <strong>and</strong> Physical Components <strong>of</strong> <strong>Tensors</strong><br />

In many applications, the components <strong>of</strong> interest are not always the components with<br />

g<br />

j . For definiteness let us call the components <strong>of</strong> a<br />

respect <strong>to</strong> the natural basis fields { } i<br />

g <strong>and</strong> { }<br />

tensor field A ∈T ∞<br />

q<br />

( U ) is defined by (45.20) the holonomic components <strong>of</strong> A . In this section, we<br />

shall consider briefly the concept <strong>of</strong> the anholonomic components <strong>of</strong> A ; i.e., the components <strong>of</strong><br />

A taken with respect <strong>to</strong> an anholonomic basis <strong>of</strong> vec<strong>to</strong>r fields. The concept <strong>of</strong> the physical<br />

components <strong>of</strong> a tensor field is a special case <strong>and</strong> will also be discussed.<br />

Let<br />

1<br />

e a<br />

denote a set <strong>of</strong> N vec<strong>to</strong>rs fields on U<br />

1<br />

, which are<br />

linearly independent, i.e., at each x ∈U 1<br />

, { e a } is a basis <strong>of</strong> V . If A is a tensor field in T ∞ q<br />

( U )<br />

where U ∩ ≠∅<br />

1<br />

U , then by the same type <strong>of</strong> argument as used in Section 45, we can write<br />

U be an open set in E <strong>and</strong> let { }<br />

a<br />

a<br />

1<br />

q<br />

A aa 1 2 ... a<br />

e e<br />

q<br />

A = ⊗⋅⋅⋅⊗<br />

(46.1)<br />

or, for example,<br />

b1<br />

... b q<br />

eb<br />

e<br />

1<br />

bq<br />

A = A ⊗⋅⋅⋅⊗<br />

(46.2)<br />

where { a<br />

}<br />

e is the reciprocal basis field <strong>to</strong> { }<br />

e defined by<br />

a<br />

a<br />

a<br />

e ( x) ⋅ e ( x ) = δ<br />

(46.3)<br />

b<br />

b<br />

for all x ∈U 1<br />

. Equations (46.1) <strong>and</strong> (46.2) hold on U ∩U 1<br />

, <strong>and</strong> the component fields as defined by<br />

A<br />

...<br />

= A ( e ,..., e )<br />

(46.4)<br />

aa 1 2 aq<br />

a1<br />

aq<br />

<strong>and</strong><br />

b1 ... bq<br />

b b<br />

1<br />

q<br />

A = A ( e ,..., e )<br />

(46.5)


Sec. 46 • Anholonomic <strong>and</strong> Physical Components 333<br />

are scalar fields on U ∩U 1<br />

. These fields are the anholonomic components <strong>of</strong> A when the bases<br />

e are not the natural bases <strong>of</strong> any coordinate system.<br />

{ a<br />

}<br />

e <strong>and</strong> { }<br />

a<br />

Given a set <strong>of</strong> N vec<strong>to</strong>r fields { e a } as above, one can show that a necessary <strong>and</strong> sufficient<br />

condition for { e a } <strong>to</strong> be the natural basis field <strong>of</strong> some chart is<br />

[ , ]<br />

e e 0<br />

a b<br />

=<br />

for all ab , = 1,..., N, where the bracket product is defined in Exercise 45.2. We shall prove this<br />

important result in Section 49. Formulas which generalize (45.22) <strong>to</strong> anholonomic components can<br />

easily be derived. If { e ˆ a } is an anholonomic basis field defined on an open set U<br />

2<br />

such that<br />

U1∩U 2<br />

≠∅, then we can express each vec<strong>to</strong>r field e ˆ b<br />

in anholonomic component form relative <strong>to</strong><br />

the basis { e a }, namely<br />

a<br />

eˆ<br />

= T e (46.6)<br />

b b a<br />

where the<br />

T are scalar fields on U1∩U 2<br />

defined by<br />

a<br />

b<br />

a<br />

T = eˆ<br />

⋅e<br />

b<br />

b<br />

a<br />

The inverse <strong>of</strong> (46.6) can be written<br />

ˆ b<br />

e = T e ˆ<br />

(46.7)<br />

a a ba<br />

where<br />

ˆ b ( )<br />

a ( )<br />

b<br />

T x T x = δ<br />

(46.8)<br />

a c c<br />

<strong>and</strong>


334 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

a<br />

( ) ˆ c a<br />

T x T ( x ) = δ<br />

(46.9)<br />

c b b<br />

for all x ∈U1∩U 2. It follows from (46.4) <strong>and</strong> (46.7) that<br />

A = T ⋅⋅⋅ T Aˆ<br />

(46.10)<br />

ˆb<br />

b<br />

1 ˆ q<br />

a1... aq a1 aq b1...<br />

bq<br />

where<br />

A ˆ = A ( eˆ ,..., eˆ<br />

)<br />

(46.11)<br />

b1...<br />

bq<br />

b1<br />

bq<br />

Equation (46.10) is the transformation rule for the anholonomic components <strong>of</strong> A . Of course,<br />

(46.10) is a field equation which holds at every point <strong>of</strong> U1∩U2∩<br />

U . Similar transformation rules<br />

for the other components <strong>of</strong> A can easily be derived by the same type <strong>of</strong> argument used above.<br />

We define the physical components <strong>of</strong> A , denoted by<br />

A<br />

a1 ,..., aq<br />

, <strong>to</strong> be the anholonomic<br />

components <strong>of</strong> A relative <strong>to</strong> the field <strong>of</strong> orthonomal basis { g i } whose basis vec<strong>to</strong>rs g are unit<br />

i<br />

vec<strong>to</strong>rs in the direction <strong>of</strong> the natural basis vec<strong>to</strong>rs g<br />

i<br />

<strong>of</strong> an orthogonal coordinate system. Let<br />

( , x )<br />

U by such a coordinate system with g = 0, i ≠ j. Then we define<br />

1 ˆ<br />

ij<br />

g ( x) g ( x) g ( x ) (no sum)<br />

(46.12)<br />

i<br />

=<br />

i<br />

i<br />

at every x ∈U 1<br />

. By (44.39), an equivalent version <strong>of</strong> (46.12) is<br />

g x = g x g x (46.13)<br />

12<br />

( )<br />

i( ) (<br />

ii( )) (no sum)<br />

i<br />

Since { } i<br />

g is orthogonal, it follows from (46.13) <strong>and</strong> (44.39) that { i }<br />

g is orthonormal:<br />

g ⋅ g = δ<br />

a b ab<br />

(46.14)


Sec. 46 • Anholonomic <strong>and</strong> Physical Components 335<br />

as it should, <strong>and</strong> it follows from (44.41) that<br />

⎡1 g11( x) 0 ⋅ ⋅ ⋅ 0 ⎤<br />

⎢<br />

0 1 g22( x) 0<br />

⎥<br />

⎢<br />

⎥<br />

⎢ ⋅ ⋅ ⋅ ⎥<br />

ij<br />

⎡<br />

⎣g<br />

( x)<br />

⎤= ⎦ ⎢ ⎥<br />

⎢<br />

⋅ ⋅ ⋅<br />

⎥<br />

⎢ ⋅ ⋅ ⋅ ⎥<br />

⎢<br />

⎥<br />

⎣ 0 0 ⋅ ⋅ ⋅ 1 gNN<br />

( x)<br />

⎦<br />

(46.15)<br />

This result shows that<br />

g can also be written<br />

i<br />

g ( x)<br />

g x x g x<br />

( g ( x))<br />

i<br />

12 i<br />

( ) = = ( g ( )) ( ) (no sum)<br />

i ii 12 ii<br />

(46.16)<br />

Equation (46.13) can be viewed as a special case <strong>of</strong> (46.7), where<br />

⎡1 g11<br />

0 ⋅ ⋅ ⋅ 0 ⎤<br />

⎢<br />

⎥<br />

⎢ 0 1 g22<br />

0 ⎥<br />

⎢<br />

ˆ b ⋅ ⋅ ⋅<br />

⎥<br />

⎡Ta<br />

⎤= ⎢ ⎥<br />

⎣ ⎦ ⎢ ⋅ ⋅ ⋅ ⎥<br />

⎢<br />

⎥<br />

⎢<br />

⋅ ⋅ ⋅<br />

⎥<br />

⎢ 0 0 1 g ⎥<br />

⎣<br />

⋅ ⋅ ⋅<br />

NN ⎦<br />

By the transformation rule (46.10), the physical components <strong>of</strong> A are related <strong>to</strong> the covariant<br />

components <strong>of</strong> A by<br />

(46.17)<br />

( )<br />

−12<br />

≡ ( g , g ,..., g ) =<br />

1 2... 1 2<br />

aa<br />

⋅⋅⋅<br />

1 1 a 1...<br />

(no sum)<br />

aa a a a<br />

q<br />

aq<br />

qaq a aq<br />

A A g g A<br />

(46.18)<br />

Equation (46.18) is a field equation which holds for all x ∈U . Since the coordinate system is<br />

orthogonal, we can replace (46.18) 1 with several equivalent formulas as follows:


336 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

aa 1 2...<br />

aq<br />

12<br />

a1<br />

... aq<br />

( aa )<br />

1 1 aqaq<br />

−<br />

( ) ( )<br />

q q<br />

A = g ⋅⋅⋅g A<br />

12 12<br />

a2<br />

... aq<br />

aa 1 1 aa 2 2 a a a1<br />

= g g ⋅⋅⋅g A<br />

⋅<br />

⋅<br />

⋅<br />

−<br />

( ) ( )<br />

12 12<br />

aq<br />

aa 1 1 aq−1aq−1 aqaq a1...<br />

aq−1<br />

= g ⋅⋅⋅g g A<br />

(46.19)<br />

In mathematical physics, tensor fields <strong>of</strong>ten arise naturally in component forms relative <strong>to</strong><br />

e are fields <strong>of</strong> bases,<br />

product bases associated with several bases. For example, if { } a<br />

ˆ b<br />

e <strong>and</strong> { }<br />

possibly anholonomic, then it might be convenient <strong>to</strong> express a second-order tensor field A as a<br />

field <strong>of</strong> linear transformations such that<br />

bˆ<br />

a aeb<br />

A e = A ˆ , a = 1,..., N<br />

(46.20)<br />

In this case A has naturally the component form<br />

bˆ<br />

a b<br />

a<br />

A = A eˆ<br />

⊗ e<br />

(46.21)<br />

a<br />

Relative <strong>to</strong> the product basis { eˆ<br />

⊗ b<br />

e } formed by { e ˆ b } <strong>and</strong> { e<br />

a<br />

}<br />

basis <strong>of</strong> { e a } as usual. For definiteness, we call { ˆ ⊗ a<br />

b }<br />

with the bases { e ˆ b } <strong>and</strong> { e<br />

a<br />

}. Then the scalar fields<br />

called the composite components <strong>of</strong> A , <strong>and</strong> they are given by<br />

, the latter being the reciprocal<br />

e e a composite product basis associated<br />

ˆb<br />

A<br />

a<br />

defined by (46.20) or (46.21), may be<br />

bˆ<br />

a<br />

b<br />

A = A ( eˆ<br />

⊗e<br />

)<br />

(46.22)<br />

a<br />

Similarly we may define other types <strong>of</strong> composite components, e.g.,<br />

ba ˆ<br />

b a<br />

Aˆ = A( eˆ , e ), A = A ( eˆ<br />

, e )<br />

(46.23)<br />

ba<br />

b<br />

a<br />

etc., <strong>and</strong> these components are related <strong>to</strong> one another by


Sec. 46 • Anholonomic <strong>and</strong> Physical Components 337<br />

A = A gˆ<br />

= A g = A gˆ<br />

g<br />

(46.24)<br />

cˆ<br />

c cd ˆ<br />

ba ˆ a cb ˆˆ bˆ ca bc ˆˆ<br />

da<br />

etc. Further, the composite components are related <strong>to</strong> the regular tensor components associated<br />

with a single basis field by<br />

c ˆ cˆ<br />

ˆ ˆ ,<br />

ba ca b bc ˆˆ a<br />

A = A T = A T etc.<br />

(46.25)<br />

a<br />

where T ˆ <strong>and</strong><br />

{ } ˆ b<br />

b<br />

ˆ a ˆ<br />

b<br />

T are given by (46.6) <strong>and</strong> (46.7) as before, In the special case where { }<br />

e <strong>and</strong><br />

e are orthogonal but not orthonormal, we define the normalized basis vec<strong>to</strong>rs e <strong>and</strong> e ˆ<br />

a<br />

a<br />

as<br />

before. Then the composite physical components A ba<br />

<strong>of</strong> A are given by<br />

ˆ,<br />

a<br />

A = A ( e ˆ , e )<br />

(46.26)<br />

ba ˆ,<br />

bˆ<br />

a<br />

<strong>and</strong> these are related <strong>to</strong> the composite components by<br />

ba ˆ,<br />

bˆ<br />

aa<br />

( ˆ<br />

ˆˆ) a ( )<br />

12 12<br />

A = g A g<br />

(no sum)<br />

(46.27)<br />

bb<br />

Clearly the concepts <strong>of</strong> composite components <strong>and</strong> composite physical components can be<br />

defined for higher order tensors also.<br />

Exercises<br />

46.1 In a three-dimensional Euclidean space the covariant components <strong>of</strong> a tensor field A<br />

relative <strong>to</strong> the cylindrical coordinate system are A<br />

ij<br />

. Determine the physical components <strong>of</strong><br />

A .<br />

46.2 Relative <strong>to</strong> the cylindrical coordinate system the helical basis { e a } has the component form


338 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

e<br />

= g<br />

1 1<br />

( cosα) ( sinα)<br />

( sinα) ( cosα)<br />

e = g + g<br />

2 2 3<br />

e =− g + g<br />

3 2 3<br />

(46.28)<br />

where α is a constant called the pitch, <strong>and</strong> where { g<br />

α } is the orthonormal basis associated<br />

with the natural basis <strong>of</strong> the cylindrical system. Show that { e a } is anholonomic. Determine<br />

the anholonomic components <strong>of</strong> the tensor field A in the preceding exercise relative <strong>to</strong> the<br />

helical basis.<br />

46.3 Determine the composite physical components <strong>of</strong> A relative <strong>to</strong> the composite product basis<br />

e ⊗ g<br />

{ a b }


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 339<br />

Section 47. Chris<strong>to</strong>ffel Symbols <strong>and</strong> Covariant Differentiation<br />

In this section we shall investigate the problem <strong>of</strong> representing the gradient <strong>of</strong> various<br />

tensor fields in components relative <strong>to</strong> the natural basis <strong>of</strong> arbitrary coordinate systems. We<br />

consider first the simple case <strong>of</strong> representing the tangent <strong>of</strong> a smooth curve in E . Let<br />

λ :( ab , ) →E be a smooth curve passing through a point x , say x = λ ( c)<br />

. Then the tangent vec<strong>to</strong>r<br />

<strong>of</strong> λ at x is defined by<br />

dλ()<br />

t<br />

λ =<br />

(47.1)<br />

x<br />

dt<br />

=<br />

t c<br />

Given the chart ( , xˆ<br />

)<br />

{ g i } <strong>of</strong> ˆx . First, the coordinates <strong>of</strong> the curve λ are given by<br />

U covering x , we can project the vec<strong>to</strong>r equation (47.1) in<strong>to</strong> the natural basis<br />

1 N<br />

1<br />

N<br />

( ) = ( λ λ ) ( ) = ( λ λ )<br />

xˆ λ( t) ( t),..., ( t) , λ t x ( t),..., ( t)<br />

(47.2)<br />

for all t such that λ () t ∈U . Differentiating (47.2) 2 with respect <strong>to</strong> t , we get<br />

j<br />

∂x<br />

dλ<br />

λ <br />

= (47.3)<br />

x ∂<br />

j<br />

x dt<br />

t=<br />

c<br />

By (47.3) this equation can be rewritten as<br />

j<br />

dλ<br />

λ = g<br />

j ( x)<br />

(47.4)<br />

x<br />

dt<br />

Thus, the components <strong>of</strong> λ relative <strong>to</strong> { g j } are simply the derivatives <strong>of</strong> the coordinate<br />

representations <strong>of</strong> λ in ˆx . In fact (44.33) can be regarded as a special case <strong>of</strong> (47.3) when λ<br />

coincides with the i th coordinate curve <strong>of</strong> ˆx .<br />

t=<br />

c<br />

An important consequence <strong>of</strong> (47.3) is that


340 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

1 1<br />

N N<br />

( +Δ ) = ( λ + υ Δ + Δ λ + υ Δ + Δ )<br />

xˆ λ ( t t) ( t) t o( t),..., ( t) t o( t)<br />

(47.5)<br />

i<br />

where υ denotes a component <strong>of</strong> λ , i.e.,<br />

υ<br />

i i<br />

= dλ<br />

dt<br />

(47.6)<br />

In particular, if λ is a straight line segment, say<br />

λ()<br />

t = x+<br />

tv (47.7)<br />

so that<br />

λ () t = v<br />

(47.8)<br />

for all t , then (47.5) becomes<br />

1 1 N N<br />

( ) ( υ υ )<br />

xˆ λ ( t) = x + t,..., x + t + o( t)<br />

(47.9)<br />

for sufficiently small t .<br />

Next we consider the gradient <strong>of</strong> a smooth function f defined on an open set U ⊂ E .<br />

From (43.17) the gradient <strong>of</strong> f is defined by<br />

d<br />

grad f( x) ⋅ v= f( x+<br />

τ v )<br />

(47.10)<br />

τ = 0<br />

dτ for all<br />

v ∈V . As before, we choose a chart near x ; then f can be represented by the function<br />

1 N<br />

1 N<br />

f ( x ,..., x ) ≡ f x ( x ,..., x )<br />

(47.11)


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 341<br />

For definiteness, we call the function f<br />

see that<br />

x the coordinate representation <strong>of</strong> f . From (47.9), we<br />

1 1<br />

N N<br />

f( x+ τ v ) = f( x + υτ+ o( τ),..., x + υ τ+<br />

o( τ))<br />

(47.12)<br />

As a result, the right-h<strong>and</strong> side <strong>of</strong> (47.10) is given by<br />

d<br />

d<br />

∂f( x)<br />

j<br />

f ( x+ τ v ) = υ<br />

(47.13)<br />

∂<br />

j<br />

x<br />

τ = 0<br />

τ<br />

Now since v is arbitrary, (47.13) can be combined with (47.10) <strong>to</strong> obtain<br />

grad f<br />

∂f<br />

x<br />

j<br />

= g (47.14)<br />

∂<br />

j<br />

j<br />

where g is a natural basis vec<strong>to</strong>r associated with the coordinate chart as defined by (44.31). In<br />

fact, that equation can now be regarded as a special case <strong>of</strong> (47.14) where f reduces <strong>to</strong> the<br />

coordinate function x ˆi<br />

.<br />

Having considered the tangent <strong>of</strong> a curve <strong>and</strong> the gradient <strong>of</strong> a function, we now turn <strong>to</strong> the<br />

problem <strong>of</strong> representing the gradient <strong>of</strong> a tensor field in general. Let A ∈T ∞<br />

q ( U ) be such a field<br />

<strong>and</strong> suppose that x is an arbitrary point in its domain U . We choose an arbitrary chart ˆx covering<br />

x . Then the formula generalizing (47.4) <strong>and</strong> (47.14) is<br />

∂A( x)<br />

j<br />

grad A ( x) = ⊗g ( x)<br />

(47.15)<br />

j<br />

∂x<br />

j<br />

where the quantity ∂A ( x)<br />

∂x on the right-h<strong>and</strong> side is the partial derivative <strong>of</strong> the coordinate<br />

representation <strong>of</strong> A , i.e.,<br />

∂A( x)<br />

∂<br />

1 N<br />

= A x<br />

( x ,..., x )<br />

(47.16)<br />

j j<br />

∂x<br />

∂x<br />

From (47.15) we see that grad A is a tensor field <strong>of</strong> order 1<br />

q + on U , grad ∈ ( U )<br />

T ∞ q+<br />

1<br />

A .


342 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

To prove (47.15) we shall first regard grad A ( x)<br />

as in LV ( ; Tq( V )) . Then by (43.15),<br />

when A is smooth, we get<br />

d<br />

A A (47.17)<br />

=<br />

dτ ( grad ( x) ) v= ( x+<br />

τ v)<br />

τ 0<br />

for all<br />

v ∈V . By using exactly the same argument from (47.10) <strong>to</strong> (47.13), we now have<br />

∂A( x)<br />

j<br />

A v = υ<br />

(47.18)<br />

∂<br />

j<br />

x<br />

( grad ( x)<br />

)<br />

Since this equation must hold for all<br />

v ∈V , we may take v = g ( x ) <strong>and</strong> find<br />

k<br />

∂A( x)<br />

grad A ( ) gk = (47.19)<br />

∂<br />

k<br />

x<br />

( x )<br />

which is equivalent <strong>to</strong> (47.15) by virtue <strong>of</strong> the canonical isomorphism between LV ( ; Tq( V )) <strong>and</strong><br />

T ( ) q+1<br />

V .<br />

Since grad A ( x) ∈Tq+<br />

1( V ) , it can be represented by its component form relative <strong>to</strong> the<br />

natural basis, say<br />

( ) 1 ... q<br />

i1<br />

i i j<br />

j<br />

iq<br />

grad A( x) = grad A ( x) g ( x) ⋅⋅⋅g ( x) ⊗g ( x)<br />

(47.20)<br />

Comparing this equation with (47.15), we se that<br />

... ∂A( x)<br />

grad A ( x) gi ( x) ⋅⋅⋅ gi ( x)<br />

= (47.21)<br />

j<br />

q<br />

∂<br />

j<br />

x<br />

i<br />

( ) 1 iq<br />

1


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 343<br />

for all j = 1,..., N . In applications it is convenient <strong>to</strong> express the components <strong>of</strong> grad A ( x)<br />

in terms<br />

<strong>of</strong> the components <strong>of</strong> A ( x)<br />

relative <strong>to</strong> the same coordinate chart. If we write the component form<br />

<strong>of</strong> A ( x)<br />

as usual by<br />

i1<br />

... iq<br />

A ( x) = A ( x) g ( x) ⊗⋅⋅⋅⊗g ( x)<br />

(47.22)<br />

i1<br />

iq<br />

then the right-h<strong>and</strong> side <strong>of</strong> (47.21) is given by<br />

i1<br />

... iq<br />

x ∂A<br />

x<br />

= ( ) ( )<br />

j<br />

j g<br />

i1<br />

x ⊗⋅⋅⋅⊗ g<br />

iq<br />

x<br />

∂x<br />

i1 ... i<br />

⎡∂g<br />

( )<br />

q<br />

i<br />

x<br />

∂gi<br />

x<br />

1<br />

q<br />

+ A ( x) ⎢ ⊗⋅⋅⋅⊗ gi<br />

( x) + g ( )<br />

j<br />

q i<br />

x ⊗⋅⋅⋅⊗<br />

1<br />

j<br />

∂A( ) ( )<br />

∂x<br />

⎣<br />

∂x<br />

( ) ⎤<br />

⎥<br />

∂x<br />

⎦<br />

(47.23)<br />

From this representation we see that it is important <strong>to</strong> express the gradient <strong>of</strong> the basis vec<strong>to</strong>r g<br />

i<br />

in<br />

component form first, since from (47.21) for the case A = gi<br />

, we have<br />

∂gi( x)<br />

=<br />

∂<br />

( )<br />

k<br />

grad g ( x) g ( x )<br />

(47.24)<br />

i<br />

x j j<br />

k<br />

or, equivalently,<br />

∂gi<br />

( x)<br />

⎧k<br />

⎫<br />

=<br />

k<br />

( )<br />

j ⎨ ⎬g x (47.25)<br />

∂x<br />

⎩ij⎭<br />

⎧k<br />

⎫<br />

We call ⎨ ⎬<br />

⎩ij⎭ the Chris<strong>to</strong>ffel symbol associated with the chart ˆx . Notice that, in general, ⎧k<br />

⎫<br />

⎨ ⎬<br />

⎩ij⎭ is a<br />

function <strong>of</strong> x , but we have suppressed the argument x in the notation. More accurately, (47.25)<br />

should be replaced by the field equation<br />

∂g<br />

∂x<br />

i<br />

j<br />

⎧k<br />

⎫<br />

= ⎨ ⎬g k<br />

(47.26)<br />

⎩ij⎭


344 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

which is valid at each point x in the domain <strong>of</strong> the chart ˆx , for all i, j = 1,..., N . We shall consider<br />

some preliminary results about the Chris<strong>to</strong>ffel symbols first.<br />

From (47.26) <strong>and</strong> (44.35) we have<br />

⎧k<br />

⎫ ∂<br />

⎨ ⎬= ⋅<br />

⎩ij⎭<br />

∂x<br />

gi<br />

k<br />

j g (47.27)<br />

By virtue <strong>of</strong> (44.33), this equation can be rewritten as<br />

2<br />

⎧k<br />

⎫ ∂ x<br />

⎨ ⎬= ⋅g<br />

i j<br />

⎩ij⎭<br />

∂x∂x<br />

k<br />

(47.28)<br />

It follows from (47.28) that the Chris<strong>to</strong>ffel symbols are symmetric in the pair ( ij ) , namely<br />

⎧k⎫ ⎧k<br />

⎫<br />

⎨ ⎬=<br />

⎨ ⎬<br />

⎩ij⎭ ⎩ji⎭<br />

(47.29)<br />

for all i, j, k = 1,..., N. Now by definition<br />

g = g ⋅g (47.30)<br />

ij i j<br />

Taking the partial derivative <strong>of</strong> (47.30) with respect <strong>to</strong><br />

we get<br />

k<br />

x <strong>and</strong> using the component form (47.26),<br />

∂gij<br />

∂g ∂g<br />

i<br />

j<br />

= ⋅<br />

k k g<br />

j<br />

+ g<br />

i⋅<br />

k<br />

(47.31)<br />

∂x ∂x ∂x<br />

When the symmetry property (47.29) is used, equation (47.31) can be solved for the Chris<strong>to</strong>ffel<br />

symbols:


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 345<br />

⎧k⎫ 1 kl ⎛∂g<br />

∂g il jl<br />

∂gij<br />

⎞<br />

⎨ ⎬= g<br />

j i l<br />

ij 2<br />

⎜ + −<br />

∂x ∂x ∂x<br />

⎟<br />

⎩ ⎭ ⎝ ⎠<br />

(47.32)<br />

where<br />

kl<br />

g denotes the contravariant components <strong>of</strong> the metric tensor, namely<br />

kl k l<br />

g = g ⋅g (47.33)<br />

The formula (47.32) is most convenient for the calculation <strong>of</strong> the Chris<strong>to</strong>ffel symbols in any given<br />

chart.<br />

As an example, we now compute the Chris<strong>to</strong>ffel symbols for the cylindrical coordinate<br />

system in a three-dimensional space. In Section 44 we have shown that the components <strong>of</strong> the<br />

metric tensor are given by (44.50) <strong>and</strong> (44.51) relative <strong>to</strong> this coordinate system. Substituting those<br />

components in<strong>to</strong> (47.32), we obtain<br />

⎧ 1 ⎫ 1 ⎧ 2 ⎫ ⎧ 2⎫<br />

1<br />

⎨ ⎬= − x , ⎨ ⎬= ⎨ ⎬=<br />

1<br />

⎩22⎭ ⎩12⎭ ⎩12⎭<br />

x<br />

(47.34)<br />

<strong>and</strong> all other Chris<strong>to</strong>ffel symbols are equal <strong>to</strong> zero.<br />

, respectively, the<br />

transformation rule for the Chris<strong>to</strong>ffel symbols can be derived in the following way:<br />

Given two charts ˆx <strong>and</strong> ŷ with natural bases { g i } <strong>and</strong> { h i }<br />

k<br />

s<br />

⎧k ⎫ k ∂gi<br />

∂x l ∂ ⎛∂y<br />

⎞<br />

⎨ ⎬= g ⋅ = h ⋅<br />

j l j ⎜ h<br />

i s ⎟<br />

⎩ij⎭ ∂x ∂y ∂x ⎝∂x<br />

⎠<br />

∂x ⎛ ∂ y ∂y ∂h<br />

∂y<br />

= ⋅ ⎜ +<br />

∂y ⎝∂x ∂x ∂x ∂y ∂x<br />

k 2 s s l<br />

l<br />

s<br />

h h<br />

l j i s i l j<br />

k 2 l k s l<br />

∂x ∂ y ∂x ∂y ∂y<br />

l ∂hs<br />

h<br />

l j i l i j l<br />

= + ⋅<br />

∂y ∂x ∂x ∂y ∂x ∂x ∂y<br />

k 2 l k s l<br />

∂x ∂ y ∂x ∂y ∂y<br />

⎧ l ⎫<br />

= + ⎨ ⎬<br />

st<br />

l j i l i j<br />

∂y ∂x ∂x ∂y ∂x ∂x<br />

⎩ ⎭<br />

⎞<br />

⎟<br />

⎠<br />

(47.35)


346 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

⎧k<br />

⎫<br />

where ⎨ ⎬<br />

⎩ij⎭ <strong>and</strong> ⎧ l ⎫<br />

⎨ ⎬ denote the Chris<strong>to</strong>ffel symbols associated with ˆx <strong>and</strong> ŷ , respectively. Since<br />

⎩st⎭ (47.35) is different from the tensor transformation rule (45.22), it follows that the Chris<strong>to</strong>ffel<br />

symbols are not the components <strong>of</strong> a particular tensor field. In fact, if ŷ is a Cartesian chart, then<br />

⎧ l ⎫<br />

⎨ ⎬<br />

⎩st⎭ vanishes since the natural basis vec<strong>to</strong>rs h<br />

s<br />

, are constant. In that case (47.35) reduces <strong>to</strong><br />

k 2 l<br />

k x y<br />

⎧ ⎫ ∂ ∂<br />

⎨ ⎬=<br />

⎩ij⎭<br />

∂y ∂x ∂x<br />

l j i<br />

(47.36)<br />

⎧k<br />

⎫<br />

<strong>and</strong> ⎨ ⎬ need not vanish unless ˆx is also a Cartesian chart. The formula (47.36) can also be used<br />

⎩ij⎭ <strong>to</strong> calculate the Chris<strong>to</strong>ffel symbols when the coordination transformation from ˆx <strong>to</strong> a Cartesian<br />

system ŷ is given.<br />

Having presented some basis properties <strong>of</strong> the Chris<strong>to</strong>ffel symbols, we now return <strong>to</strong> the<br />

general formula (47.23) for the components <strong>of</strong> the gradient <strong>of</strong> a tensor field. Substituting (47.26)<br />

in<strong>to</strong> (47.23) yields<br />

i1<br />

... iq<br />

∂A( x) ⎡∂A<br />

( x)<br />

ki2... i ⎧ i<br />

q 1 ⎫<br />

i1...<br />

iq<br />

1k<br />

⎧ iq<br />

⎫⎤<br />

−<br />

= A<br />

A<br />

i<br />

( ) ( )<br />

j ⎢ +<br />

j ⎨ ⎬+⋅⋅⋅+ ⎨ ⎬⎥g x ⊗⋅⋅⋅⊗g 1<br />

i<br />

x<br />

q<br />

∂x ⎢⎣<br />

∂x ⎩kj<br />

⎭ ⎩kj<br />

⎭⎥⎦<br />

(47.37)<br />

Comparing this result with (47.21), we finally obtain<br />

i1<br />

... iq<br />

i1... iq i1<br />

... iq<br />

∂A<br />

ki2... i ⎧ i<br />

q 1 ⎫<br />

i1...<br />

iq<br />

1k<br />

⎧ iq<br />

⎫<br />

−<br />

A ,<br />

j<br />

≡ ( gradA ) = + A A<br />

j j ⎨ ⎬+⋅⋅⋅+<br />

⎨ ⎬ (47.38)<br />

∂x<br />

⎩kj<br />

⎭ ⎩kj<br />

⎭<br />

1<br />

This particular formula gives the components A ... , <strong>of</strong> the gradient <strong>of</strong> A in terms <strong>of</strong> the<br />

i1 ... iq<br />

contravariant components A <strong>of</strong> A . If the mixed components <strong>of</strong> A are used, the formula<br />

becomes<br />

i i q<br />

j


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 347<br />

i1<br />

... ir<br />

1... i1 ... i ∂A i ir r<br />

j1<br />

... js<br />

li2... i ⎧ i1<br />

⎫<br />

r i1...<br />

ir<br />

1l<br />

⎧ ir<br />

⎫<br />

−<br />

A<br />

j ( )<br />

1... j<br />

, grad<br />

s k<br />

≡ A = + A<br />

j1<br />

... j ,<br />

j1... j<br />

A<br />

s j1...<br />

j<br />

s k<br />

k<br />

⎨ ⎬+⋅⋅⋅+<br />

⎨ ⎬<br />

s<br />

∂x<br />

⎩lk<br />

⎭ ⎩lk<br />

⎭<br />

−A<br />

⎧ l<br />

⎫<br />

−⋅⋅⋅− A<br />

⎧ l<br />

i1... ir<br />

i1...<br />

ir<br />

lj2... j ⎨ ⎬<br />

s<br />

j1...<br />

js<br />

1l<br />

⎨ ⎬<br />

−<br />

⎩jk<br />

1 ⎭ ⎩jk<br />

s ⎭<br />

⎫<br />

(47.39)<br />

We leave the pro<strong>of</strong> <strong>of</strong> this general formula as an exercise. From (47.39), if A ∈T ∞ ( U ), where<br />

q= r+ s, then grad ∈T ∞ i1<br />

A<br />

q+<br />

1( U ) . Further, if the coordinate system is Cartesian, then A<br />

i1<br />

... ir<br />

k<br />

reduces <strong>to</strong> the ordinary partial derivative <strong>of</strong> A wih respect <strong>to</strong> x .<br />

j1<br />

... js<br />

q<br />

... ir<br />

j1<br />

... j<br />

,<br />

s k<br />

Some special cases <strong>of</strong> (47.39) should be noted here. First, since the metric tensor is a<br />

constant second-order tensor field, we have<br />

ij i<br />

g , = g , = δ , = 0<br />

(47.40)<br />

ij k k j k<br />

for all ilk , , = 1,..., N. In fact, (47.40) is equivalent <strong>to</strong> (47.31), which we have used <strong>to</strong> obtain the<br />

formula (47.32) for the Chris<strong>to</strong>ffel symbols. An important consequence <strong>of</strong> (47.40) is that the<br />

operations <strong>of</strong> raising <strong>and</strong> lowering <strong>of</strong> indices commute with the operation <strong>of</strong> gradient or covariant<br />

differentiation.<br />

Another constant tensor field on E is the tensor field E defined by (45.26). While the sign<br />

<strong>of</strong> E depends on the orientation, we always have<br />

E<br />

i1<br />

... iN<br />

i1<br />

... i<br />

, , 0<br />

N k<br />

E<br />

k<br />

= = (47.41)<br />

If we substitute (45.27) <strong>and</strong> (45.28) in<strong>to</strong> (47.41), we can rewrite the result in the form<br />

1<br />

g<br />

∂ g ⎧ k ⎫<br />

= ⎨ ⎬<br />

x lk<br />

i<br />

∂ ⎩ ⎭<br />

(47.42)<br />

where g is the determinant <strong>of</strong><br />

⎡g<br />

⎤ as defined by (44.45).<br />

⎣ ij ⎦


348 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

Finally, the operation <strong>of</strong> skew-symmetrization<br />

constant tensor. So we have<br />

K<br />

r<br />

introduced in Section 37 is also a<br />

δ<br />

...<br />

, = 0<br />

(47.43)<br />

i1<br />

... ir<br />

j1<br />

jr<br />

k<br />

in any coordinate system. Thus, the operations <strong>of</strong> skew-symmetrization <strong>and</strong> covariant<br />

differentiation commute, provided that the indices <strong>of</strong> the covariant differentiations are not affected<br />

by the skew-symmetrization.<br />

Some classical differential opera<strong>to</strong>rs can be derived from the gradient. First, if A is a<br />

tensor field <strong>of</strong> order q ≥ 1, then the divergence <strong>of</strong> A is defined by<br />

qq+ , 1<br />

( grad )<br />

div A = C A<br />

where C denotes the contraction operation. In component form we have<br />

i1...<br />

iq<br />

1k<br />

A − kgi g<br />

1 iq−1<br />

div A = , ⊗⋅⋅⋅⊗<br />

(47.44)<br />

so that div A is a tensor field <strong>of</strong> order q − 1. In particular, for a vec<strong>to</strong>r field v , (47.44) reduces <strong>to</strong><br />

i ij<br />

div v = υ , = g υ ,<br />

(47.45)<br />

i i j<br />

By use <strong>of</strong> (47.42) <strong>and</strong> (47.40), we an rewrite this formula as<br />

i<br />

( gυ<br />

)<br />

1 ∂<br />

div v =<br />

(47.46)<br />

i<br />

g ∂x<br />

This result is useful since it does not depend on the Chris<strong>to</strong>ffel symbols explicitly.<br />

by<br />

The Laplacian <strong>of</strong> a tensor field <strong>of</strong> order q + 1 is a tensor field <strong>of</strong> the same order q defined


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 349<br />

Δ A =div ( grad A )<br />

(47.47)<br />

In component form we have<br />

i1<br />

... iq<br />

kl<br />

Δ A = g A , g ⊗⋅⋅⋅⊗g<br />

(47.48)<br />

kl i1<br />

iq<br />

For a scalar field f the Laplacian is given by<br />

kl 1 ∂ ⎛ kl ∂f<br />

⎞<br />

Δ f = g f,<br />

kl<br />

= gg<br />

l ⎜ k ⎟<br />

g ∂x<br />

⎝ ∂x<br />

⎠<br />

(47.49)<br />

where (47.42), (47.14) <strong>and</strong> (47.39) have been used. Like (47.46), the formula (47.49) does not<br />

i1 depend explicitly on the Chris<strong>to</strong>ffel symbols. In (47.48), ... i<br />

A<br />

q<br />

, denotes the components <strong>of</strong> the<br />

second gradient ( )<br />

1<br />

grad grad A <strong>of</strong> A . The reader will verify easily that A<br />

... ,<br />

kl<br />

i i q<br />

kl<br />

, like the ordinary<br />

second partial derivative, is symmetric in the pair ( kl , ). Indeed, if the coordinate system is<br />

i1 Cartesian, then ... i q<br />

2 i1...<br />

iq<br />

k l<br />

A , reduces <strong>to</strong> ∂ A ∂x ∂ x .<br />

kl<br />

Finally, the classical curl opera<strong>to</strong>r can be defined in the following way. If v is a vec<strong>to</strong>r<br />

field, then curl v is a skew-symmetric second-order tensor field defined by<br />

2<br />

( grad v)<br />

curl v ≡ K (47.50)<br />

where K<br />

2<br />

is the skew-symmetrization opera<strong>to</strong>r. In component form (47.50) becomes<br />

1<br />

i j<br />

curl v = ( υi, j−υ<br />

j,<br />

i)<br />

g ⊗g (47.51)<br />

2<br />

By virtue <strong>of</strong> (47.29) <strong>and</strong> (47.39) this formula can be rewritten as<br />

1 ⎛∂υ ∂υ<br />

i j ⎞ i j<br />

curl v =<br />

j i<br />

2<br />

⎜ − ⊗<br />

∂x<br />

∂x<br />

⎟g g (47.52)<br />

⎝ ⎠


350 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

which no longer depends on the Chris<strong>to</strong>ffel symbols. We shall generalize the curl opera<strong>to</strong>r <strong>to</strong><br />

arbitrary skew-symmetric tensor fields in the next chapter.<br />

Exercises<br />

47.1 Show that in spherical coordinates on a three-dimensional Euclidean manifold the nonzero<br />

Chris<strong>to</strong>ffel symbols are<br />

⎧ 2 ⎫ ⎧ 2 ⎫ ⎧ 3 ⎫ ⎧ 3⎫<br />

1<br />

⎨ ⎬= ⎨ ⎬= ⎨ ⎬= ⎨ ⎬=<br />

1<br />

⎩21⎭ ⎩12⎭ ⎩31⎭ ⎩13⎭<br />

x<br />

⎧ 1 ⎫ 1<br />

⎨ ⎬ =− x<br />

⎩22⎭<br />

⎧ 1 ⎫ 1 2 2<br />

⎨ ⎬ =− x (sin x )<br />

⎩33⎭<br />

⎧ 3 ⎫ ⎧ 3⎫<br />

2<br />

⎨ ⎬= ⎨ ⎬=<br />

cot x<br />

⎩32⎭ ⎩23⎭<br />

⎧ 2⎫ 2 2<br />

⎨ ⎬ =− sin x cos x<br />

⎩33⎭<br />

47.2 On an oriented three-dimensional Euclidean manifold the curl <strong>of</strong> a vec<strong>to</strong>r field can be<br />

regarded as a vec<strong>to</strong>r field by<br />

∂υ<br />

ijk<br />

ijk j<br />

curl v ≡− E υ<br />

j,<br />

g<br />

k i<br />

=−E<br />

g<br />

k i<br />

(47.53)<br />

∂x<br />

ijk<br />

where E denotes the components <strong>of</strong> the positive volume tensor E . Show that<br />

curl curl v = grad div v −div grad v . Also show that<br />

( ) ( ) ( )<br />

curl( grad f ) = 0<br />

(47.54)<br />

For any scalar field f <strong>and</strong> that


Sec. 47 • Chris<strong>to</strong>ffel Symbols, Covariant Differentiation 351<br />

( )<br />

div curl v = 0<br />

(47.55)<br />

for any vec<strong>to</strong>r field v .<br />

47.3 Verify that<br />

k<br />

⎧k<br />

⎫ ∂g<br />

⎨ ⎬= − ⋅g j i<br />

(47.56)<br />

⎩ij⎭<br />

∂x<br />

Therefore,<br />

∂g<br />

∂x<br />

k<br />

j<br />

⎧k<br />

⎫<br />

=−⎨ ⎬<br />

⎩ij⎭<br />

i<br />

g (47.57)<br />

47.4 Prove the formula (47.37).<br />

47.5 Prove the formula (47.42) <strong>and</strong> show that<br />

1 ∂ g<br />

div g j<br />

=<br />

x j<br />

(47.58)<br />

g ∂<br />

47.6 Show that for an orthogonal coordinate system<br />

⎧k<br />

⎫<br />

⎨ ⎬ = 0,<br />

⎩ij⎭<br />

i ≠ j, i ≠ k,<br />

j ≠ k<br />

⎧j<br />

⎫ 1 ∂gii<br />

⎨ ⎬=−<br />

j<br />

⎩ii⎭<br />

2g jj<br />

∂x<br />

if i ≠ j<br />

⎧ i⎫ ⎧ i ⎫ 1 ∂gii<br />

⎨ ⎬= ⎨ ⎬= −<br />

j<br />

⎩ij⎭ ⎩ji⎭<br />

2gii<br />

∂x<br />

if i ≠<br />

⎧ i⎫<br />

1 ∂gii<br />

⎨ ⎬=<br />

i<br />

⎩ii⎭<br />

2gii<br />

∂x<br />

j<br />

Where the indices i <strong>and</strong> j are not summed.


352 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

47.7 Show that<br />

∂<br />

∂x<br />

j<br />

⎧k⎫ ∂ ⎧k⎫ ⎧t ⎫⎧k⎫ ⎧t ⎫⎧k⎫<br />

⎨ ⎬− 0<br />

l ⎨ ⎬+ ⎨ ⎬⎨ ⎬− ⎨ ⎬⎨ ⎬=<br />

⎩il⎭ ∂x<br />

⎩ij⎭ ⎩il⎭⎩tj⎭ ⎩ij⎭⎩tl⎭<br />

The quantity on the left-h<strong>and</strong> side <strong>of</strong> this equation is the component <strong>of</strong> a fourth-order tensor<br />

R , called the curvature tensor, which is zero for any Euclidean manifold 2 .<br />

2 The Maple computer program contains a package tensor that will produce Chris<strong>to</strong>ffel symbols <strong>and</strong> other important<br />

tensor quantities associated with various coordinate systems.


Sec. 48 • Covariant Derivatives along Curves 353<br />

Section 48. Covariant Derivatives along Curves<br />

In the preceding section we have considered covariant differentiation <strong>of</strong> tensor fields which<br />

are defined on open submanifolds in E . In applications, however, we <strong>of</strong>ten encounter vec<strong>to</strong>r or<br />

tensor fields defined only on some smooth curve in E . For example, if λ : ( ab , ) →E is a smooth<br />

curve, then the tangent vec<strong>to</strong>r λ is a vec<strong>to</strong>r field on E . In this section we shall consider the<br />

problem <strong>of</strong> representing the gradients <strong>of</strong> arbitrary tensor fields defined on smooth curves in a<br />

Euclidean space.<br />

Given any smooth curve λ :( ab , ) →E <strong>and</strong> a field A :( ab , ) → Tq( V ) we can regard the<br />

value A () t as a tensor <strong>of</strong> order q at λ () t . Then the gradient or covariant derivative <strong>of</strong> A along λ<br />

is defined by<br />

d A () t ( t+ Δt) − () t<br />

≡ lim<br />

A A<br />

dt Δ→ t 0 Δt<br />

(48.1)<br />

for all t∈ ( a, b)<br />

. If the limit on the right-h<strong>and</strong> side <strong>of</strong> (48.1) exists, then dA () t dt is itself also a<br />

2 2<br />

tensor field <strong>of</strong> order q on λ . Hence, we can define the second gradient d A () t dt by replacing<br />

A by dA () t dt in (48.1). Higher gradients <strong>of</strong> A are defined similarly. If all gradients <strong>of</strong> A<br />

exist, then A is C ∞ -smooth on λ . We are interested in representing the gradients <strong>of</strong> A in<br />

component form.<br />

Let ˆx be a coordinate system covering some point <strong>of</strong> λ . Then as before we can<br />

λ i ( t), i = 1,..., N . Similarly, we can express A in<br />

characterize λ by its coordinate representation ( )<br />

component form<br />

i1<br />

... iq<br />

A ( t) = A ( t) g ( λ( t)) ⊗⋅⋅⋅⊗g ( λ( t))<br />

(48.2)<br />

i1<br />

iq<br />

where the product basis is that <strong>of</strong> ˆx at λ () t , the point where A () t is defined. Differentiating<br />

(48.2) with respect <strong>to</strong> t , we obtain


354 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

i1<br />

... iq<br />

dA() t dA () t<br />

= gi<br />

( λ( t)) ⊗⋅⋅⋅⊗g ( ( ))<br />

1<br />

i<br />

λ t<br />

q<br />

dt dt<br />

i1 ... i ⎡dg<br />

( ( ))<br />

q i<br />

λ t<br />

dg<br />

1<br />

+ A ( t) ⎢ ⊗⋅⋅⋅⊗ gi<br />

( λ( t)) +⋅⋅⋅+ g ( ( ))<br />

q<br />

i<br />

λ t ⊗⋅⋅⋅⊗<br />

1<br />

⎣ dt<br />

iq<br />

( λ( t))<br />

⎤<br />

⎥<br />

dt ⎦<br />

(48.3)<br />

By application <strong>of</strong> the chain rule, we obtain<br />

j<br />

dgi( λ( t)) ∂gi( λ( t)) dλ<br />

( t)<br />

=<br />

j<br />

dt ∂x dt<br />

(48.4)<br />

where (47.5) <strong>and</strong> (47.6) have been used. In the preceding section we have represented the partial<br />

derivative <strong>of</strong> g<br />

i<br />

by the component form (47.26). Hence we can rewrite (48.4) as<br />

j<br />

dgi( λ( t)) ⎧k<br />

⎫dλ<br />

( t)<br />

= ⎨ ⎬ gk<br />

( λ ( t))<br />

(48.5)<br />

dt ⎩ij⎭<br />

dt<br />

Substitution <strong>of</strong> (48.5) in<strong>to</strong> (48.3) yields the desired component representation<br />

i1<br />

... iq<br />

j<br />

dA() t ⎧⎪dA () t ⎡<br />

ki2... i ⎧ i1<br />

⎫<br />

1...<br />

()<br />

q<br />

i iq<br />

1k<br />

⎧ iq<br />

⎫⎤dλ<br />

t ⎫<br />

−<br />

⎪<br />

= ⎨ + ⎢A () t ⎨ ⎬+⋅⋅⋅+<br />

A () t ⎨ ⎬⎥ ⎬<br />

dt ⎪⎩<br />

dt ⎢⎣<br />

⎩kj<br />

⎭ ⎩kj<br />

⎭⎥⎦<br />

dt ⎪⎭<br />

× g ( λ( t)) ⊗⋅⋅⋅⊗g ( λ( t))<br />

i1<br />

iq<br />

(48.6)<br />

where the Chris<strong>to</strong>ffel symbols are evaluated at the position λ ( t)<br />

. The representation (48.6) gives<br />

i i<br />

dA()<br />

t dt<br />

q<br />

<strong>of</strong> dA () t dt in terms <strong>of</strong> the contravariant components<br />

the contravariant components ( ) 1...<br />

i1...<br />

i q<br />

A () t <strong>of</strong> A () t relative <strong>to</strong> the same coordinate chart ˆx . If the mixed components are used, the<br />

representation becomes<br />

i1<br />

... ir<br />

dA()<br />

t ⎪⎧ dA<br />

j1 ... j<br />

() t ⎡<br />

s ki2... i ⎧ i1<br />

⎫<br />

r i1...<br />

ir<br />

1k<br />

⎧ ir<br />

⎫<br />

−<br />

= ⎨ + ⎢ A<br />

j1... j<br />

() t ⎨ ⎬+⋅⋅⋅+<br />

A<br />

1...<br />

()<br />

s<br />

j j<br />

t ⎨ ⎬<br />

s<br />

dt ⎪⎩<br />

dt ⎣ ⎩kl<br />

⎭ ⎩kl<br />

⎭<br />

l<br />

k k<br />

i1... i ⎧ ⎫<br />

1...<br />

()<br />

r<br />

i i ⎧ ⎫⎤ dλ<br />

t ⎫<br />

r<br />

−A kj2... j<br />

() t ⎨ ⎬−⋅⋅⋅−<br />

A<br />

1...<br />

()<br />

s<br />

j js<br />

1k<br />

t ⎨ ⎬<br />

−<br />

⎥ ⎬<br />

⎩jl<br />

1 ⎭ ⎩jl<br />

s ⎭⎦<br />

dt ⎭<br />

j1<br />

js<br />

× g ( λ( t))<br />

⊗⋅⋅⋅⊗gi ( λ( t)) ⊗g ( λ( t)) ⊗⋅⋅⋅⊗g ( λ( t))<br />

i1<br />

r<br />

(48.7)


Sec. 48 • Covariant Derivatives along Curves 355<br />

We leave the pro<strong>of</strong> <strong>of</strong> this general formula as an exercise. In view <strong>of</strong> the representations (48.6) <strong>and</strong><br />

(48.7), we see that it is important <strong>to</strong> distinguish the notation<br />

⎛<br />

⎜<br />

⎝<br />

dA<br />

dt<br />

⎞<br />

⎟<br />

⎠<br />

i1<br />

... ir<br />

j1<br />

... js<br />

which denotes a component <strong>of</strong> the covariant derivative dA<br />

dt, from the notation<br />

dA<br />

i1<br />

... ir<br />

j1<br />

... js<br />

dt<br />

which denotes the derivative <strong>of</strong> a component <strong>of</strong> A . For this reason we shall denote the former by<br />

the new notation<br />

DA<br />

i1<br />

... ir<br />

j1<br />

... js<br />

Dt<br />

(48.6)<br />

As an example we compute the components <strong>of</strong> the gradient <strong>of</strong> a vec<strong>to</strong>r field v along λ . By<br />

or, equivalently,<br />

i<br />

j<br />

dv() t ⎧dυ<br />

() t i<br />

k ⎧ ⎫dλ<br />

() t ⎫<br />

= ⎨ + υ () t ⎨ ⎬ ⎬gi( λ ()) t<br />

(48.8)<br />

dt ⎩ dt ⎩kj⎭<br />

dt ⎭<br />

i i j<br />

Dυ () t dυ () t k ⎧ i ⎫dλ<br />

() t<br />

= + υ () t ⎨ ⎬<br />

Dt dt ⎩kj⎭<br />

dt<br />

(48.9)<br />

In particular, when v is the i th basis vec<strong>to</strong>r<br />

i<br />

g <strong>and</strong> then λ is the j th coordinate curve, then (48.8)<br />

reduces <strong>to</strong>


356 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

∂g<br />

∂x<br />

i<br />

j<br />

⎧k<br />

⎫<br />

= ⎨ ⎬g<br />

⎩ij⎭<br />

k<br />

which is the previous equation (47.26).<br />

Next if v is the tangent vec<strong>to</strong>r λ <strong>of</strong> λ , then (48.8) reduces <strong>to</strong><br />

λ() ⎧ λ () λ () ⎧ i ⎫ λ () ⎫<br />

⎨ ⎨ ⎬ ⎬gi<br />

( λ ( t))<br />

(48.10)<br />

⎩<br />

⎩ ⎭ ⎭<br />

2 2 i k j<br />

d t d t d t d t<br />

= +<br />

2 2<br />

dt dt dt kj dt<br />

where (47.4) has been used. In particular, if λ is a straight line with homogeneous parameter, i.e.,<br />

if λ<br />

= v = const , then<br />

2 i k j<br />

d () t d () t d () t i<br />

2<br />

⎨<br />

dt dt dt kj<br />

λ + λ λ ⎧ ⎫<br />

⎬= 0, i=<br />

1,..., N<br />

⎩ ⎭<br />

(48.11)<br />

This equation shows that, for the straight line λ( t)<br />

= x+<br />

tv given by (47.7), we can sharpen the<br />

result (47.9) <strong>to</strong><br />

⎛<br />

1<br />

1 1 k j 1⎧ ⎫ 2 N N 1 k j ⎧ N ⎫ 2⎞<br />

2<br />

xˆ ( λ ( t) ) = ⎜x + υ t− υ υ ⎨ ⎬t ,..., x + υ t− υ υ ⎨ ⎬t ⎟+<br />

o( t ) (48.12)<br />

⎝<br />

2⎩kj⎭ 2 ⎩kj<br />

⎭ ⎠<br />

for sufficiently small t , where the Chris<strong>to</strong>ffel symbols are evaluated at the point x = λ (0) .<br />

Now suppose that A is a tensor field on U , say<br />

in U . Then the restriction <strong>of</strong> A on λ is a tensor field<br />

A ∈T ∞<br />

q<br />

, <strong>and</strong> let λ :( ab , ) →<br />

U be a curve<br />

A ˆ () t ≡ A ( λ()),<br />

t t∈(<br />

a, b)<br />

(48.13)<br />

In this case we can compute the gradient <strong>of</strong> A along λ either by (48.6) or directly by the chain rule<br />

<strong>of</strong> (48.13). In both cases the result is


Sec. 48 • Covariant Derivatives along Curves 357<br />

i1<br />

... iq<br />

j<br />

dA( λ( t)) ⎡∂A ( λ( t)) ki2... i ⎧ i1<br />

⎫<br />

1...<br />

( )<br />

q<br />

i iq<br />

1k<br />

⎧ iq<br />

⎫⎤dλ<br />

t<br />

−<br />

= ⎢<br />

+ A () t A () t<br />

j<br />

⎨ ⎬+⋅⋅⋅+<br />

⎨ ⎬⎥<br />

dt ⎣⎢<br />

∂x ⎩kj<br />

⎭ ⎩kj<br />

⎭⎦⎥<br />

dt<br />

× g ( λ( t)) ⊗⋅⋅⋅⊗g ( λ( t))<br />

i1<br />

iq<br />

(48.14)<br />

or, equivalently,<br />

dA( λ( t))<br />

=<br />

dt<br />

( λ t )<br />

grad A( ( )) λ ( t)<br />

(48.15)<br />

where grad A ( λ( t))<br />

is regarded as an element <strong>of</strong> LV ( ; T( V )) as before. Since grad A is given<br />

by (47.15), the result (48.15) also can be written as<br />

q<br />

i<br />

dA( λ( t)) ∂A( λ( t)) dλ<br />

( t)<br />

=<br />

j<br />

dt ∂x dt<br />

(48.16)<br />

A special case <strong>of</strong> (48.16), when A reduces <strong>to</strong> g<br />

i<br />

, is (48.4).<br />

By (48.16) it follows that the gradient <strong>of</strong> the metric tensor, the volume tensor, <strong>and</strong> the skewsymmetrization<br />

opera<strong>to</strong>r all vanish along any curve.<br />

Exercises<br />

48.1 Prove the formula (48.7).<br />

48.2 If the parameter t <strong>of</strong> a curve λ is regarded as time, then the tangent vec<strong>to</strong>r<br />

d<br />

v = λ<br />

dt<br />

is the velocity <strong>and</strong> the gradient <strong>of</strong> v


358 Chap. 9 • EUCLIDEAN MANIFOLDS<br />

d<br />

a = v<br />

dt<br />

is the acceleration. For a curve λ in a three-dimensional Euclidean manifold, express the<br />

acceleration in component form relative <strong>to</strong> the spherical coordinates.


______________________________________________________________________________<br />

Chapter 10<br />

VECTOR FIELDS AND DIFFERENTIAL FORMS<br />

Section 49. Lie Derivatives<br />

Let E be a Euclidean manifold <strong>and</strong> let<br />

In Exercise 45.2 we have defined the Lie bracket [ uv , ] by<br />

u<strong>and</strong><br />

vbe vec<strong>to</strong>r fields defined on some open set U inE.<br />

[ , ] = ( grad ) −( grad )<br />

uv v u u v (49.1)<br />

In this section first we shall explain the geometric meaning <strong>of</strong> the formula (49.1), <strong>and</strong> then we<br />

generalize the operation <strong>to</strong> the Lie derivative <strong>of</strong> arbitrary tensor fields.<br />

To interpret the operation on the right-h<strong>and</strong> side <strong>of</strong> (49.1), we start from the concept <strong>of</strong><br />

the flow generated by a vec<strong>to</strong>r field. We say that a curve λ :( ab , ) →U is an integral curve <strong>of</strong> a<br />

vec<strong>to</strong>r field v if<br />

for all t ( a b)<br />

dλ<br />

( t)<br />

dt<br />

( () t )<br />

= v λ (49.2)<br />

∈ , . By (47.1), the condition (49.2) means that λ is an integral curve <strong>of</strong> v if <strong>and</strong><br />

only if its tangent vec<strong>to</strong>r coincides with the value <strong>of</strong> v at every point λ ( t).<br />

An integral curve<br />

may be visualized as the orbit <strong>of</strong> a point flowing with velocity v . Then the flow generated by v<br />

λ t along any integral curve <strong>of</strong><br />

is defined <strong>to</strong> be the mapping that sends a point λ ( ) <strong>to</strong> a point ( )<br />

v .<br />

To make this concept more precise, let us introduce a local coordinate system ˆx . Then<br />

the condition (49.2) can be represented by<br />

t 0<br />

( )<br />

i<br />

( ) / υ ( )<br />

i<br />

dλ<br />

t dt t<br />

= λ (49.3)<br />

i<br />

where (47.4) has been used. This formula shows that the coordinates ( )<br />

( t , i 1, , N)<br />

λ = … <strong>of</strong> an<br />

integral curve are governed by a system <strong>of</strong> first-order differential equations. Now it is proved in<br />

i<br />

the theory <strong>of</strong> differential equations that if the fields υ on the right-h<strong>and</strong> side <strong>of</strong> (49.3) are<br />

smooth, then corresponding <strong>to</strong> any initial condition, say<br />

359


360 Chap. 10 • VECTOR FIELDS<br />

λ ( 0 ) = x<br />

0<br />

(49.4)<br />

or equivalently<br />

i<br />

( ) 0<br />

a unique solution <strong>of</strong> (49.3) exists on a certain interval ( − δ , δ )<br />

i<br />

λ 0 = x , i= 1, …,<br />

N<br />

(49.5)<br />

, where δ may depend on the<br />

initial point x<br />

0<br />

but it may be chosen <strong>to</strong> be a fixed, positive number for all initial points in a<br />

sufficiently small neighborhood <strong>of</strong> x . 0<br />

For definiteness, we denote the solution <strong>of</strong> (49.2)<br />

x by λ t, x ; then it is known that the mapping from x <strong>to</strong><br />

corresponding <strong>to</strong> the initial point ( )<br />

λ ( t, x ); then it is known that the mapping from x <strong>to</strong> λ ( t, )<br />

x is smooth for each t belonging <strong>to</strong><br />

the interval <strong>of</strong> existence <strong>of</strong> the solution. We denote this mapping by ρt,<br />

namely<br />

( ) = ( )<br />

ρt x λ t, x (49.6)<br />

<strong>and</strong> we call it the flow generated by v . In particular,<br />

( )<br />

ρ x = x x∈U (49.7)<br />

0<br />

,<br />

reflecting the fact that x is the initial point <strong>of</strong> the integral curve λ ( t, x ).<br />

i<br />

Since the fields υ are independent <strong>of</strong> t, they system (49.3) is said <strong>to</strong> be au<strong>to</strong>nomous. A<br />

characteristic property <strong>of</strong> such a system is that the flow generated by v forms a local oneparameter<br />

group. That is, locally,<br />

ρ ρ ρ<br />

(49.8)<br />

t1+ t<br />

=<br />

2 t1 t2<br />

or equivalently<br />

( )<br />

( t + t , ) = t , ( t , )<br />

λ x λ λ x (49.9)<br />

1 2 2 1<br />

for all t 1<br />

, t 2<br />

, x such that the mappings in (49.9) are defined. Combining (49.7) with (49.9), we see<br />

that the flow ρ<br />

t<br />

is a local diffeomorphism <strong>and</strong>, locally,<br />

ρ = ρ (49.10)<br />

−1<br />

t<br />

−t


Sec. 49 • Lie Derivatives 361<br />

Consequently the gradient, grad ρt,<br />

is a linear isomorphism which carries a vec<strong>to</strong>r at any point<br />

x <strong>to</strong> a vec<strong>to</strong>r at the point ρt<br />

( x ). We call this linear isomorphism the parallelism generated by<br />

the flow <strong>and</strong>, for brevity, denote it by P<br />

t<br />

. The parallelism P<br />

t<br />

is a typical two-point tensor whose<br />

component representation has the form<br />

i<br />

( ) P ( ) ( )<br />

t t j i t<br />

j<br />

( ) ( )<br />

P x = x g ρ x ⊗g x (49.11)<br />

or, equivalently,<br />

{ }<br />

{ i } j( t( ))<br />

i<br />

( ) = P ( ) ( ( ))<br />

( ) ⎤ ( )<br />

⎡⎣Pt x ⎦ g<br />

j<br />

x<br />

t<br />

x g<br />

j i<br />

ρt<br />

x (49.12)<br />

where g ( x)<br />

<strong>and</strong> g ρ x are the natural bases <strong>of</strong> ˆx at the points x <strong>and</strong> ( )<br />

respectively.<br />

ρ x ,<br />

Now the parallelism generated by the flow <strong>of</strong> v gives rise <strong>to</strong> a difference quotient <strong>of</strong> the<br />

vec<strong>to</strong>r field u in the following way: At any point x ∈U , Pt<br />

( x ) carries the vec<strong>to</strong>r u( x)at<br />

x <strong>to</strong><br />

P x u x at ( ),<br />

u ρ x also at<br />

the vec<strong>to</strong>r ( )<br />

( t )( ( ))<br />

ρ ( x ). Thus we define the difference quotient<br />

t<br />

ρ x which can then be compared with the vec<strong>to</strong>r ( )<br />

t<br />

t<br />

( t )<br />

1<br />

( t( )) ( t( ))( ( ))<br />

t ⎡ ⎣<br />

u ρ x − P x u x ⎤ ⎦<br />

(49.13)<br />

The limit <strong>of</strong> this difference quotient as t approaches zero is called the Lie derivative <strong>of</strong> u with<br />

respect <strong>to</strong> v <strong>and</strong> is denoted by<br />

1<br />

L u( x) ≡lim ⎡ ( t( )) −( t( ))( ( ))<br />

⎤<br />

v<br />

t→0<br />

t ⎣<br />

u ρ x P x u x<br />

⎦<br />

(49.14)<br />

We now derive a representation for the Lie derivative in terms <strong>of</strong> a local coordinate<br />

system x ˆ. In view <strong>of</strong> (49.14) we see that we need an approximate representation for P<br />

t<br />

<strong>to</strong> within<br />

first-order terms in t. Let x<br />

0<br />

be a particular reference point. From (49.3) we have<br />

λ<br />

( t,<br />

) = x + υ ( ) t+<br />

o( t)<br />

x x (49.15)<br />

i i i<br />

0 0 0<br />

i i i<br />

Suppose that x is an arbitrary neighboring point <strong>of</strong> x<br />

0<br />

, say x = x0<br />

+Δ x . Then by the same<br />

argument as (49.15) we have also


362 Chap. 10 • VECTOR FIELDS<br />

λ<br />

( t,<br />

x ) = x + υ ( x) t+<br />

o( t)<br />

i i i<br />

( x0<br />

) ( ) ( ) ()<br />

i<br />

i i i<br />

∂υ<br />

j k<br />

= x0+Δ x + υ ( x0)<br />

t+ Δ x t+ o Δ x + o t<br />

j<br />

∂x<br />

(49.16)<br />

Subtracting (49.15) from (49.16), we obtain<br />

It follows from (49.17) that<br />

( x0<br />

) ( ) ( ) ()<br />

i<br />

i i i<br />

∂υ<br />

j k<br />

λ ( t, x) − λ ( t,<br />

x<br />

0 ) =Δ x + Δ x t+ o Δ x + o t<br />

(49.17)<br />

j<br />

∂x<br />

( x )<br />

i<br />

P<br />

i<br />

i i<br />

∂υ<br />

0<br />

t( x0 ) ≡ ( grad ρt( x<br />

o)<br />

) = δ<br />

t j<br />

+ + 0<br />

j<br />

() t<br />

(49.18)<br />

j<br />

j<br />

∂x<br />

( t 0 )<br />

Now assuming that u is also smooth, we can represent ( )<br />

( x )<br />

u ρ x approximately by<br />

i<br />

i i<br />

∂u<br />

0 j<br />

u ( ρt<br />

( x0)<br />

) = u ( x0)<br />

+ υ<br />

j<br />

( x<br />

0) t+<br />

o( t)<br />

(49.19)<br />

∂x<br />

where (49.15) has been used. Substituting (49.18) <strong>and</strong> (49.19) in<strong>to</strong> (49.14) <strong>and</strong> taking the limit,<br />

we finally obtain<br />

i<br />

( x ) ∂υ<br />

( x )<br />

i<br />

⎡∂u<br />

0 j<br />

0 j<br />

⎤<br />

L u( x0)<br />

= ⎢ υ − u<br />

j<br />

j<br />

( x0) ⎥ gi<br />

( x0)<br />

(49.20)<br />

v<br />

⎣ ∂x<br />

∂x<br />

⎦<br />

where x<br />

0<br />

is arbitrary. Thus the field equation for (49.20) is<br />

⎡ ∂u<br />

⎣∂x<br />

∂υ<br />

∂x<br />

⎤<br />

i<br />

i<br />

j<br />

j<br />

L u = ⎢ υ − u<br />

j<br />

j ⎥ gi<br />

(49.21)<br />

v<br />

⎦<br />

By (47.39) or by using a Cartesian system, we can rewrite (49.21) as<br />

( u<br />

i , υ<br />

j υ<br />

i , u<br />

j<br />

)<br />

L u = − g<br />

(49.22)<br />

v<br />

j j i<br />

Consequently the Lie derivative has the following coordinate-free representation:<br />

( grad ) ( grad )<br />

L u= u v−<br />

v u<br />

(49.23)<br />

v


Sec. 49 • Lie Derivatives 363<br />

Comparing (49.23) with (49.1), we obtain<br />

[ v,<br />

u]<br />

L u=<br />

(49.24)<br />

v<br />

L<br />

v<br />

u<br />

Since the Lie derivative is defined by the limit <strong>of</strong> the difference quotient (49.13),<br />

vanishes if <strong>and</strong> only if u commutes with the flow <strong>of</strong> v in the following sense:<br />

u ρ = Pu<br />

(49.25)<br />

t<br />

i<br />

When u satisfies this condition, it may be called invariant with respect <strong>to</strong> v . Clearly, this<br />

uv , , since the Lie bracket is skew-symmetric, namely<br />

condition is symmetric for the pair ( )<br />

[ , ] = − [ , ]<br />

uv vu (49.26)<br />

In particular, (49.25) is equivalent <strong>to</strong><br />

v ϕ = Q v<br />

(49.27)<br />

t<br />

t<br />

where<br />

ϕ<br />

t<br />

<strong>and</strong><br />

Q<br />

t<br />

denote the flow <strong>and</strong> the parallelism generated by u .<br />

Another condition equivalent <strong>to</strong> (49.25) <strong>and</strong> (49.27) is<br />

ϕ ρ<br />

= ρ ϕ (49.28)<br />

s t t s<br />

which means that<br />

( an integralcurve<strong>of</strong> )<br />

( an integralcurve<strong>of</strong> v)<br />

ρ u an integralcurve <strong>of</strong> u<br />

t<br />

=<br />

ϕ an integralcurve <strong>of</strong> v<br />

(49.29)<br />

s<br />

=<br />

To show that (49.29) is necessary <strong>and</strong> sufficient for<br />

we choose any particular integral curve :( −δδ<br />

, )<br />

integral curve μ ( st , ) for u such that<br />

[ uv , ] = 0 (49.30)<br />

λ →U for v , At each point λ () t we define an<br />

( 0,t) = ( t)<br />

μ λ (49.31)


364 Chap. 10 • VECTOR FIELDS<br />

The integral curves μ ( ⋅,t ) with t ∈( − δ , δ ) then sweep out a surface parameterized by ( st )<br />

We shall now show that (49.30) requires that the curves ( s,<br />

⋅)<br />

As before, let ˆx be a local coordinate system. Then by definition<br />

, .<br />

μ be integral curves <strong>of</strong> v for all s.<br />

<strong>and</strong><br />

( )<br />

i<br />

( st , ) / s u ( st , )<br />

i<br />

∂μ<br />

∂ = μ (49.32)<br />

( )<br />

i<br />

( 0, t) / s υ ( 0, t)<br />

i<br />

∂μ<br />

∂ = μ (49.33)<br />

<strong>and</strong> we need <strong>to</strong> prove that<br />

( )<br />

i<br />

( st , ) / s υ ( st , )<br />

i<br />

∂μ<br />

∂ = μ (49.34)<br />

for s ≠ 0. We put<br />

i<br />

ζ<br />

( st)<br />

( st , )<br />

i<br />

∂μ<br />

i<br />

, ≡ −υ<br />

( μ ( st , ))<br />

(49.35)<br />

∂t<br />

By (49.33)<br />

i<br />

We now show that ( st , )<br />

<strong>to</strong> s <strong>and</strong> use (49.32), we obtain<br />

( t)<br />

i<br />

ζ 0, = 0<br />

(49.36)<br />

ζ vanishes identically. Indeed, if we differentiate (49.35) with respect<br />

i i i j<br />

∂ζ ∂ ⎛∂μ ⎞ ∂υ ∂μ<br />

= −<br />

j<br />

∂s ∂t ⎜<br />

s<br />

⎟<br />

⎝ ∂ ⎠ ∂x ∂s<br />

i j i<br />

∂u<br />

∂μ<br />

∂υ<br />

= − u<br />

j<br />

j<br />

∂x ∂t ∂x<br />

∂u<br />

⎛∂u<br />

∂υ<br />

= + −<br />

∂x ⎜<br />

⎝∂x ∂x<br />

i i i<br />

j i j<br />

ζ υ u<br />

j j j<br />

j<br />

⎞<br />

⎟<br />

⎠<br />

(49.37)<br />

i<br />

As a result, when (49.30) holds, ζ are governed by the system <strong>of</strong> differential equations<br />

i<br />

∂ζ<br />

∂s<br />

i<br />

∂u<br />

j<br />

= ζ<br />

j<br />

∂x<br />

(49.38)


Sec. 49 • Lie Derivatives 365<br />

i<br />

<strong>and</strong> subject <strong>to</strong> the initial condition (49.36) for each fixed t. Consequently, ζ = 0 is the only<br />

i<br />

solution. Conversely, when ζ vanishes identically on the surface, (49.37) implies immediately<br />

that the Lie bracket <strong>of</strong> u<strong>and</strong><br />

v vanishes. Thus the condition (49.28) is shown <strong>to</strong> be equivalent <strong>to</strong><br />

the condition (49.30).<br />

The result just established can be used <strong>to</strong> prove the theorem mentioned in Section 46 that<br />

a field <strong>of</strong> basis is holonomic if <strong>and</strong> only if<br />

⎡<br />

⎣hi, h<br />

j⎤ ⎦ = 0<br />

(49.39)<br />

for all i, j 1, , N.<br />

h i<br />

is holonomic, the components <strong>of</strong><br />

i<br />

each h<br />

i<br />

relative <strong>to</strong> the coordinate system corresponding <strong>to</strong> { h i } are the constants δ<br />

j.<br />

Hence<br />

from (49.21) we must have (49.39). Conversely, suppose (49.39) holds. Then by (49.28) there<br />

exists a surface swept out by integral curves <strong>of</strong> the vec<strong>to</strong>r fields h1 <strong>and</strong> h<br />

2.<br />

We denote the<br />

surface parameters by x1 <strong>and</strong> x<br />

2.<br />

Now if we define an integral curve for h<br />

3<br />

at each surface point<br />

( )<br />

1 2<br />

− … Necessity is obvious, since when { }<br />

x , x , then by the conditions<br />

[ h h ] [ h h ]<br />

, = , = 0<br />

1 3 2 3<br />

we see that the integral curves <strong>of</strong> h1, h2,<br />

h<br />

3<br />

form a three-dimensional net which can be regarded<br />

as a “surface coordinate system” ( x1, x2,<br />

x<br />

3)<br />

on a three-dimensional hypersurface in the N-<br />

dimensional Euclidean manifold E . By repeating the same process based on the condition<br />

(49.39), we finally arrive at an N-dimensional net formed by integral curves <strong>of</strong> h , , .<br />

1<br />

… h N<br />

The<br />

corresponding N-dimensional coordinate system x , 1<br />

… x now forms a chart in E <strong>and</strong> its<br />

natural basis is the given basis { h i }. Thus the theorem is proved. In the next section we shall<br />

make use <strong>of</strong> this theorem <strong>to</strong> prove the classical Frobenius theorem.<br />

So far we have considered the Lie derivative L u <strong>of</strong> a vec<strong>to</strong>r field u relative <strong>to</strong> v only.<br />

Since the parallelism P<br />

t<br />

generated by v is a linear isomorphism, it can be extended <strong>to</strong> tensor<br />

fields. As usual for simple tensors we define<br />

v<br />

, N<br />

( ) ( ) ( )<br />

P a⊗ b⊗ = Pa ⊗ Pb ⊗ (49.40)<br />

t t t<br />

Then we extend P<br />

t<br />

<strong>to</strong> arbitrary tensors by linearity. Using this extended parallelism, we define<br />

the Lie derivative <strong>of</strong> a tensor field A with respect <strong>to</strong> v by<br />

1<br />

L A( x) ≡lim ⎡ ( t( )) −( t( ))( ( ))<br />

⎤<br />

v<br />

t→0<br />

t ⎣<br />

A ρ x P x A x<br />

⎦<br />

(49.41)


366 Chap. 10 • VECTOR FIELDS<br />

which is clearly a generalization <strong>of</strong> (49.14). In terms <strong>of</strong> a coordinate system it can be shown that<br />

( L A)<br />

v<br />

i1…<br />

ir<br />

j1…<br />

js<br />

∂A<br />

=<br />

∂x<br />

i1…<br />

ir<br />

j1…<br />

js<br />

k<br />

υ<br />

k<br />

−A<br />

∂υ<br />

−<br />

∂x<br />

−<br />

+A<br />

∂υ<br />

∂x<br />

A<br />

i1<br />

ki2…<br />

ir<br />

i1…<br />

ir−1k<br />

j<br />

A<br />

1…<br />

j<br />

<br />

s k<br />

j1…<br />

js<br />

∂υ<br />

k<br />

∂x<br />

∂υ<br />

∂x<br />

k<br />

k<br />

i1…<br />

ir<br />

i1…<br />

ir<br />

k…<br />

j<br />

+ <br />

s j1<br />

j1…<br />

js−1k js<br />

ik<br />

(49.42)<br />

which generalizes the formula (49.21). We leave the pro<strong>of</strong> <strong>of</strong> (49.42) as an exercise. By the<br />

same argument as (49.22), the partial derivatives in (49.42) can be replaced by covariant<br />

derivatives.<br />

It should be noted that the operations <strong>of</strong> raising <strong>and</strong> lowering <strong>of</strong> indices by the Euclidean<br />

metric do not commute with the Lie derivative, since the parallelism P<br />

t<br />

generated by the flow<br />

generally does not preserve the metric. Consequently, <strong>to</strong> compute the Lie derivative <strong>of</strong> a tensor<br />

field A,<br />

we must assign a particular contravariant order <strong>and</strong> covariant order <strong>to</strong> A . The formula<br />

(49.42) is valid when A is regarded as a tensor field <strong>of</strong> contravariant order r <strong>and</strong> covariant order<br />

s.<br />

By the same <strong>to</strong>ken, the Lie derivative <strong>of</strong> a constant tensor such as the volume tensor or<br />

the skew-symmetric opera<strong>to</strong>r generally need not vanish.<br />

Exercises<br />

49.1 Prove the general representation formula (49.42) for the Lie derivative.<br />

49.2 Show that the right-h<strong>and</strong> side <strong>of</strong> (49.42) obeys the transformation rule <strong>of</strong> the components<br />

<strong>of</strong> a tensor field.<br />

49.3 In the two-dimensional Euclidean plane E , consider the vec<strong>to</strong>r field v whose<br />

1 2<br />

x , x are<br />

components relative <strong>to</strong> a rectangular Cartesian coordinate system ( )<br />

υ = αx<br />

, υ = αx<br />

1 1 2 2<br />

where α is a positive constant. Determine the flow generated by v . In particular, find<br />

the integral curve passing through the initial point<br />

( x 1 2<br />

0, x<br />

0 ) = ( 1,1)<br />

49.4 In the same two-dimensional plane E consider the vec<strong>to</strong>r field such that


Sec. 49 • Lie Derivatives 367<br />

υ = x , υ = x<br />

1 2 2 1<br />

Show that the flow generated by this vec<strong>to</strong>r field is the group <strong>of</strong> rotations <strong>of</strong> E . In<br />

particular, show that the Euclidean metric is invariant with respect <strong>to</strong> this vec<strong>to</strong>r field.<br />

49.5 Show that the flow <strong>of</strong> the au<strong>to</strong>nomous system (49.3) possesses the local one-parameter<br />

group property (49.8).


368 Chap. 10 • VECTOR FIELDS<br />

Section 50. The Frobenius Theorem<br />

The concept <strong>of</strong> the integral curve <strong>of</strong> a vec<strong>to</strong>r field has been considered in some detail in<br />

the preceding section. In this section we introduce a somewhat more general concept. If v is a<br />

non-vanishing vec<strong>to</strong>r field in E , then at each point x in the domain <strong>of</strong> v we can define a onedimensional<br />

Euclidean space D ( x)<br />

spanned by the vec<strong>to</strong>r v( x ). In this sense an integral curve<br />

<strong>of</strong> v corresponds <strong>to</strong> a one-dimensional hypersurface in E which is tangent <strong>to</strong> D ( x)<br />

at each<br />

point <strong>of</strong> the hypersurface. Clearly this situation can be generalized if we allow the field <strong>of</strong><br />

Euclidean spaces D <strong>to</strong> be multidimensional. For definiteness, we call such a field D a<br />

distribution in E , say <strong>of</strong> dimension D. Then a D-dimensional hypersurface L in E is called an<br />

integral surface <strong>of</strong> D if L is tangent <strong>to</strong> D at every point <strong>of</strong> the hypersurface.<br />

Unlike a one-dimensional distribution, which corresponds <strong>to</strong> some non-vanishing vec<strong>to</strong>r<br />

field <strong>and</strong> hence always possesses many integral curves, a D-dimensional distribution D in<br />

general need not have any integral hypersurface at all. The problem <strong>of</strong> characterizing those<br />

distributions that do possess integral hypersurfaces is answered precisely by the Frobenius<br />

theorem. Before entering in<strong>to</strong> the details <strong>of</strong> this important theorem, we introduce some<br />

preliminary concepts first.<br />

Let D be a D-dimensional distribution <strong>and</strong> let v be a vec<strong>to</strong>r field. Then v is said <strong>to</strong> be<br />

v x ∈D x at each x in the domain <strong>of</strong> v . Since D is D-dimensional, there<br />

v<br />

α<br />

, α = 1, …,D<br />

contained in D such that D ( x)<br />

is spanned by<br />

contained in D if ( ) ( )<br />

exist D vec<strong>to</strong>r fields { }<br />

{ α ( ), α = 1, ,D}<br />

{ , 1, ,D}<br />

v x … at each point x in the domain <strong>of</strong> the vec<strong>to</strong>r fields. We call<br />

v<br />

α<br />

α = … a local basis for D . We say that D is smooth at some point x<br />

0<br />

if there exists a<br />

local basis defined on a neighborhood <strong>of</strong> x<br />

0<br />

formed by smooth vec<strong>to</strong>r fields v<br />

α<br />

contained in D .<br />

Naturally, D is said <strong>to</strong> be smooth if it is smooth at each point <strong>of</strong> its domain. We shall be<br />

interested in smooth distributions only.<br />

We say that D is integrable at a point x<br />

0<br />

if there exists a local coordinate system ˆx<br />

defined on a neighborhood <strong>of</strong> x<br />

0<br />

such that the vec<strong>to</strong>r fields { g<br />

α<br />

, α = 1, …,D}<br />

form a local basis<br />

for D . Equivalently, this condition means that the hypersurfaces characterized by the conditions<br />

D 1<br />

N<br />

x<br />

+ = const, …, x = const<br />

(50.1)<br />

are integral hypersurfaces <strong>of</strong> D . Since the natural basis vec<strong>to</strong>rs g<br />

i<br />

<strong>of</strong> any coordinate system are<br />

smooth, by the very definition D must be smooth at x0<br />

if it is integrable there. Naturally, D is<br />

said <strong>to</strong> be integrable if it is integrable at each point <strong>of</strong> its domain. Consequently, every<br />

integrable distribution must be smooth.


Sec. 50 • The Frobenius Theorem 369<br />

Now we are ready <strong>to</strong> state <strong>and</strong> prove the Frobenius theorem, which characterizes<br />

integrable distributions.<br />

Theorem 50.1. A smooth distribution D is integrable if <strong>and</strong> only if it is closed with respect <strong>to</strong><br />

the Lie bracket, i.e.,<br />

[ uv]<br />

uv , ∈D ⇒ , ∈D (50.2)<br />

Pro<strong>of</strong>. Necessity can be verified by direct calculation. If uv , ∈D <strong>and</strong> supposing that ˆx is a local<br />

coordinate system such that { g<br />

α<br />

, α = 1, …,D}<br />

forms a local basis for D , then u<strong>and</strong><br />

v have the<br />

component forms<br />

u = u α g , v =υ<br />

α g (50.3)<br />

α<br />

α<br />

where the Greek index α is summed from 1 <strong>to</strong> D. Substituting (50.3) in<strong>to</strong> (49.21), we see that<br />

α<br />

α<br />

⎛∂υ<br />

β ∂u<br />

β ⎞<br />

uv = ⎜ u − υ<br />

β<br />

β ⎟g α<br />

(50.4)<br />

⎝ ∂x<br />

∂x<br />

⎠<br />

[ , ]<br />

Thus (50.2) holds.<br />

Conversely, suppose that (50.2) holds. Then for any local basis { , 1, ,D}<br />

v<br />

α<br />

α = … for D<br />

we have<br />

⎡<br />

⎣v , C γ<br />

α<br />

vβ⎤ ⎦ =<br />

αβv γ<br />

(50.5)<br />

where C γ αβ<br />

are some smooth functions. We show first that there exists a local basis<br />

{ , 1, ,D}<br />

u<br />

α<br />

α = … which satisfies the somewhat stronger condition<br />

⎣<br />

⎡uα, u<br />

β⎤ ⎦ = 0, αβ , = 1, …,D<br />

(50.6)<br />

To construct such a basis, we choose a local coordinate system ŷ <strong>and</strong> represent the basis { v α }<br />

by the component form<br />

v = υ k + + υ k + υ k + +<br />

υ k<br />

1 D D+<br />

1<br />

N<br />

α α 1 α D α D+<br />

1<br />

α N<br />

≡ υ k + υ k<br />

β Δ<br />

α β α<br />

Δ<br />

(50.7)


370 Chap. 10 • VECTOR FIELDS<br />

where { k i } denotes the natural basis <strong>of</strong> ŷ , <strong>and</strong> whre the repeated Greek indices β <strong>and</strong> Δ are<br />

summed from 1 <strong>to</strong> D <strong>and</strong> D + 1 <strong>to</strong> N , respectively. Since the local basis { v } is linearly<br />

independent, without loss <strong>of</strong> generality we can assume that the D× D matrix ⎣<br />

nonsingular, namely<br />

Now we define the basis { u α } by<br />

α<br />

β<br />

⎡υ α<br />

⎤<br />

⎦ is<br />

β<br />

det ⎡<br />

⎣υ α<br />

⎤<br />

⎦ ≠ 0<br />

(50.8)<br />

u ≡ υ v , α = 1, …,D<br />

(50.9)<br />

−1β<br />

α α β<br />

1β<br />

where υ −<br />

β<br />

⎡<br />

⎣ α<br />

⎤<br />

⎦ denotes the inverse matrix <strong>of</strong> ⎡<br />

⎣υ α<br />

⎤<br />

⎦ ,<br />

see that the component representation <strong>of</strong> u<br />

α<br />

is<br />

as usual. Substituting (50.7) in<strong>to</strong> (50.9), we<br />

β Δ<br />

Δ<br />

u = δ k + u k = k + u k (50.10)<br />

α α β α Δ α α Δ<br />

We now show that the basis { u<br />

α}<br />

has the property (50.6). From (50.10), by direct<br />

calculation based on (49.21), we can verify easily that the first D components <strong>of</strong> ⎡<br />

⎣uα<br />

, u<br />

β<br />

⎤<br />

⎦ are<br />

zero, i.e., ⎡<br />

⎣uα,<br />

u<br />

β⎤<br />

⎦ has the representation<br />

Δ<br />

⎡<br />

⎣uα, uβ⎤ ⎦ = Kαβk Δ<br />

(50.11)<br />

but, by assumption, D is closed with respect <strong>to</strong> the Lie bracket; it follows that<br />

⎡<br />

⎣u , K γ<br />

α<br />

uβ⎤ ⎦ =<br />

αβu γ<br />

(50.12)<br />

Substituting (50.10) in<strong>to</strong> (50.12), we then obtain<br />

γ<br />

Δ<br />

⎡<br />

⎣uα, uβ⎤ ⎦ = Kαβkγ + Kαβk Δ<br />

(50.13)<br />

Comparing this representation with (50.11), we see that K γ αβ<br />

must vanish <strong>and</strong> hence, by (50.12),<br />

(50.6) holds.<br />

V <strong>and</strong> that<br />

Now we claim that the local basis { u<br />

α}<br />

for D can be extended <strong>to</strong> a field <strong>of</strong> basis { i }<br />

u for


Sec. 50 • The Frobenius Theorem 371<br />

⎡<br />

⎣ui, u<br />

j⎤ ⎦ = 0 , i, j = 1, …,<br />

N<br />

(50.14)<br />

This fact is more or less obvious. From (50.6), by the argument presented in the preceding<br />

section, the integral curves <strong>of</strong> { uα}<br />

form a “coordinate net” on a D-dimensional hypersurface<br />

defined in the neighborhood <strong>of</strong> any reference point x . To define 0<br />

u , D+ 1<br />

we simply choose an<br />

arbitrary smooth curve λ ( t, x<br />

0 ) passing through x , 0<br />

having a nonzero tangent, <strong>and</strong> not belonging<br />

<strong>to</strong> the hyper surface generated by { u , , 1<br />

… u } . D<br />

We regard the points <strong>of</strong> the curve λ ( t, x<br />

0 ) <strong>to</strong><br />

have constant coordinates in the coordinate net on the D-dimensional hypersurfaces, say<br />

1 D<br />

,…, .<br />

λ t, x for all neighboring points x on the hypersurface<br />

( xo<br />

x<br />

o ) Then we define the curves ( )<br />

D<br />

<strong>of</strong> x<br />

o<br />

, by exactly the same condition with constant coordinates ( x x )<br />

the flow generated by the curves ( t, )<br />

u <strong>to</strong> be the tangent vec<strong>to</strong>r field <strong>of</strong> the curves λ ( t, )<br />

define<br />

D+ 1<br />

1 ,…, . Thus, by definition,<br />

λ x preserves the integral curves <strong>of</strong> any u . Hence if we<br />

[ ]<br />

x , then<br />

u , , 1, ,<br />

D+ 1<br />

uα<br />

= 0 α = … D<br />

(50.15)<br />

α<br />

where we have used the necessary <strong>and</strong> sufficient condition (49.29) for the condition (49.30).<br />

Having defined the vec<strong>to</strong>r fields { u … u }<br />

1, ,<br />

D+ 1<br />

which satisfy the conditions (50.6) <strong>and</strong><br />

(50.15), we repeat that same procedure <strong>and</strong> construct the fields uD+ 2, u<br />

D+<br />

3, …,<br />

until we arrive at a<br />

field <strong>of</strong> basis { u , , 1<br />

… u } . N<br />

Now from a theorem proved in the preceding section [cf. (49.39)] the<br />

condition (50.14) is necessary <strong>and</strong> sufficient that { u i } be the natural basis <strong>of</strong> a coordinate system<br />

x ˆ. Consequently, D is integrable <strong>and</strong> the pro<strong>of</strong> is complete.<br />

From the pro<strong>of</strong> <strong>of</strong> the preceding theorem it is clear that an integrable distribution D can<br />

be characterized by the opening remark: In the neighborhood <strong>of</strong> any point x<br />

0<br />

in the domain <strong>of</strong><br />

D there exists a D-dimensional hypersurface S such that D ( x)<br />

is the D-dimensional tangent<br />

hyperplane <strong>of</strong> S at each point x in S .<br />

Exercises<br />

We shall state <strong>and</strong> prove a dual version <strong>of</strong> the Frobenius theorem in Section 52.<br />

50.1 Let D be an integrable distribution defined on U . Then we define the following<br />

equivalence relation on U: x0 ∼ x1<br />

⇔ there exists a smooth curve λ in U joining x0<strong>to</strong>x<br />

1<br />

<strong>and</strong> tangent <strong>to</strong> D at each point, i.e., λ ( t) ∈D<br />

( λ( t)<br />

), for all t. Suppose that S is an<br />

equivalence set relative <strong>to</strong> the preceding equivalence relation. Show that S is an


372 Chap. 10 • VECTOR FIELDS<br />

integral surface <strong>of</strong> D . (Such an integral surface is called a maximal integral surface or a<br />

leaf <strong>of</strong> D .)


Sec. 51 • Differential Forms, Exterior Derivative 373<br />

Section 51. Differential Forms <strong>and</strong> Exterior Derivative<br />

In this section we define a differential opera<strong>to</strong>r on skew-symmetric covariant tensor fields.<br />

This opera<strong>to</strong>r generalizes the classical curl opera<strong>to</strong>r defined in Section 47.<br />

Let U be an open set in E <strong>and</strong> let A be a skew-symmetric covariant tensor field <strong>of</strong> order<br />

r on U ,i.e.,<br />

:<br />

r<br />

ˆ ( )<br />

A U → T V<br />

(51.1)<br />

x U , ( )<br />

Then for any point ∈ A x is a skew-symmetric tensor <strong>of</strong> order r (cf. Chapter 8).<br />

Choosing a coordinate chart xˆ onU as before, we can express A in the component form<br />

i1<br />

ir<br />

A = A i 1…<br />

i<br />

g ⊗⊗g<br />

(51.2)<br />

r<br />

where { g i<br />

} denotes the natural basis <strong>of</strong> x ˆ. Since A is skew-symmetric, its components obey the<br />

identify<br />

A<br />

i1 j k i<br />

= −A<br />

(51.3)<br />

……… r i1……… k j ir<br />

for any pair <strong>of</strong> indices ( j,<br />

k ) in ( )<br />

by<br />

i ,..., i<br />

1 r<br />

. As explained in Section 39, we can then represent A<br />

i1 i 1<br />

r<br />

i1<br />

ir<br />

A = ∑ Ai 1... ig ∧∧ g = A<br />

r<br />

i1...<br />

ig ∧∧g<br />

(51.4)<br />

r<br />

r!<br />

i1<br />

< <<br />

ir<br />

where ∧ denotes the exterior product.<br />

Now if A is smooth, then it is called a differential form <strong>of</strong> order r, or simply an r-form.<br />

In the theory <strong>of</strong> differentiable manifolds, differential forms are important geometric entities,<br />

since they correspond <strong>to</strong> linear combinations <strong>of</strong> volume tensors on various hypersurfaces (cf. the<br />

book by Fl<strong>and</strong>ers 1 ). For our purpose, however, we shall consider only some elementary results<br />

about differential forms.<br />

1 H. Fl<strong>and</strong>ers, Differential Forms with Applications <strong>to</strong> the Physical Sciences, Academic Press, New York-London,<br />

1963.


374 Chap. 10 • VECTOR FIELDS<br />

We define first the notion <strong>of</strong> the exterior derivative <strong>of</strong> a differential form. Let A be the r-<br />

r + 1 –form given<br />

form represented by (51.2) or (51.4). Then the exterior derivative dA is an ( )<br />

by any one <strong>of</strong> the following three equivalent formulas:<br />

N ∂A<br />

dA<br />

= ∑∑ g ∧g ∧ ∧g<br />

∂x<br />

i1<br />

< < ir<br />

k=<br />

1<br />

( )<br />

i1…<br />

ir<br />

k i1<br />

ir<br />

<br />

k<br />

1 ∂Ai 1…<br />

ir<br />

k i1<br />

ir<br />

= g ∧g ∧∧g<br />

k<br />

r!<br />

∂x<br />

1 ∂A<br />

=<br />

r! r+ 1 ! ∂x<br />

∧ ∧<br />

ki1…<br />

i i1…<br />

i<br />

r<br />

r j1 jr+<br />

1<br />

δ<br />

j1…<br />

j<br />

g g<br />

r+<br />

1 k<br />

(51.5)<br />

Of course, we have <strong>to</strong> show that dA as defined by (51.5), is independent <strong>of</strong> the choice <strong>of</strong> the<br />

chart ˆ. x If this result is granted, then (51.5) can be written in the coordinate-free form<br />

( 1) r<br />

( 1 )! ( gradA)<br />

dA<br />

= − r + K<br />

+<br />

(51.6)<br />

r 1<br />

where K<br />

r+ 1<br />

is the skew-symmetric opera<strong>to</strong>r introduced in Section 37.<br />

system, then<br />

i<br />

We shall now show that dA is well-defined by (51.5), i.e., if ( x ) is another coordinate<br />

or, equivalently,<br />

∂A<br />

∂x<br />

∂A<br />

g ∧g ∧∧ g = g ∧g ∧ g<br />

(51.7)<br />

∂x<br />

i1…<br />

ir<br />

k i1 ir<br />

i1…<br />

ir<br />

k i1<br />

ir<br />

k<br />

k<br />

l1 l<br />

∂A r 1<br />

1 r i1 i<br />

∂A +<br />

ki i r ki1<br />

ir<br />

i1<br />

i ∂x ∂x<br />

<br />

<br />

<br />

r<br />

δ<br />

j1… j<br />

= δ<br />

r+ 1 k l1… l<br />

(51.8)<br />

r+ 1 k j1 jr<br />

+ 1<br />

∂x ∂x ∂x ∂x<br />

for all j1, …, j<br />

r + 1,<br />

where<br />

i1 i<br />

g<br />

i denote the components <strong>of</strong> A <strong>and</strong> the natural basis<br />

r<br />

i<br />

corresponding <strong>to</strong> ( x ). To prove (51.8), we recall first the transformation rule<br />

A …<br />

<strong>and</strong> { }<br />

j1<br />

jr<br />

∂x<br />

∂x<br />

Ai 1 i<br />

= A<br />

r j1 jr i1<br />

ir<br />

∂x<br />

∂x<br />

(51.9)<br />

k<br />

for any covariant tensor components. Now differentiating (51.9) with respect <strong>to</strong> x , we obtain


Sec. 51 • Differential Forms, Exterior Derivative 375<br />

∂ ∂ ∂ ∂ ∂<br />

∂x ∂x ∂x ∂x ∂x<br />

l j1<br />

j<br />

A<br />

r<br />

i1 i<br />

A<br />

r j1<br />

j x x x<br />

<br />

… r<br />

= <br />

k l k i1<br />

ir<br />

+ A<br />

2 j1 j2<br />

jr<br />

∂ x ∂x ∂x<br />

j1…<br />

j<br />

<br />

r k i1 i2<br />

ir<br />

+ +<br />

A<br />

∂x ∂x ∂x ∂x<br />

j1 jr−1<br />

2 jr<br />

∂x ∂x ∂ x<br />

<br />

∂x ∂x ∂x ∂x<br />

j1…<br />

jr i1 ir−1<br />

k ir<br />

(51.10)<br />

2 j k i<br />

Since the second derivative ∂ x ∂x ∂ x is symmetric with respect <strong>to</strong> the pair ( ki , ), when we<br />

form the contraction <strong>of</strong> (51.10) with the skew-symmetric opera<strong>to</strong>r K<br />

r+ 1<br />

the result is (51.8). Here<br />

we have used the fact that K<br />

r+ 1<br />

is a tensor <strong>of</strong> order 2( r + 1 ),<br />

so that we have the identities<br />

δ<br />

∂x ∂x ∂x ∂x ∂x<br />

∂x ∂x ∂x ∂x ∂x<br />

∂x ∂x ∂x ∂x ∂x<br />

∂x ∂x ∂x ∂x ∂x<br />

k i1 ir<br />

p1 pr+<br />

1<br />

ki1i r lm1m<br />

r<br />

j1…<br />

j<br />

= δ<br />

r+ 1 p1…<br />

p<br />

<br />

r+ 1 1 m1 mr j1 jr+<br />

1<br />

= δ<br />

k i1 ir<br />

p1 pr+<br />

1<br />

lm1<br />

mr<br />

p1…<br />

p<br />

<br />

r+ 1 l m1 mr j1 jr+<br />

1<br />

(51.11)<br />

From (38.2) <strong>and</strong> (51.5), if A is a 1-form, say A = u,<br />

then<br />

⎛ ∂u<br />

∂u<br />

i j ⎞ j i<br />

du = ⎜ − ⊗<br />

j i<br />

∂x<br />

∂x<br />

⎟g g (51.12)<br />

⎝ ⎠<br />

Comparing this representation with (47.52), we see that the exterior derivative <strong>and</strong> the curl<br />

opera<strong>to</strong>r are related by<br />

2curl<br />

u= − du (51.13)<br />

for any 1-form u . Equation (51.13) also follows from (51.6) <strong>and</strong> (47.50). In the sense <strong>of</strong> (51.13)<br />

, the exterior derivative is a generalization <strong>of</strong> the curl opera<strong>to</strong>r from 1-forms <strong>to</strong> r-forms in<br />

general. We shall now consider some basic properties <strong>of</strong> the exterior derivative.<br />

I. If f is a smooth function (i.e., a 0-form), then the exterior derivative <strong>of</strong> f coincides<br />

with the gradient <strong>of</strong> f,<br />

df<br />

= grad f<br />

(51.14)<br />

This result follows readily from (51.5) <strong>and</strong> (47.14).<br />

II. If w is a smooth covariant vec<strong>to</strong>r field (i.e., a 1-form), <strong>and</strong> if u <strong>and</strong> v are smooth<br />

vec<strong>to</strong>r fields, then


376 Chap. 10 • VECTOR FIELDS<br />

( ) ( ) [ , ] ( , )<br />

u⋅d v⋅w −v⋅d u⋅ w = u v ⋅ w+<br />

dw u v (51.15)<br />

We can prove this formula by direct calculation. Let the component representations <strong>of</strong><br />

uvwin , , a chart ˆx be<br />

i i i<br />

u = u g , v = υ g , w = w g (51.16)<br />

i i i<br />

Then<br />

From I it follows that<br />

i<br />

i<br />

u⋅ w = uw,<br />

v⋅ w =υ w<br />

(51.17)<br />

i<br />

i<br />

i<br />

( )<br />

∂ uw<br />

i<br />

i k ⎛ i ∂wi<br />

∂u<br />

⎞ k<br />

d( u⋅ w)<br />

= g = u + w<br />

k k i k<br />

∂x ⎜<br />

∂x ∂x<br />

⎟g (51.18)<br />

⎝<br />

⎠<br />

<strong>and</strong> similarly<br />

d<br />

( )<br />

i<br />

( υ w )<br />

∂<br />

i<br />

i k ⎛ i ∂wi<br />

∂υ<br />

⎞ k<br />

v⋅ w = g = υ + w<br />

k k i k<br />

∂x ⎜<br />

∂x ∂x<br />

⎟g (51.19)<br />

⎝<br />

⎠<br />

Consequently,<br />

i<br />

i<br />

⎛ k ∂υ<br />

k ∂u<br />

⎞<br />

k i i k ∂wi<br />

u⋅d( v⋅w) −v⋅d( u⋅ w ) = ⎜u − υ w<br />

k k ⎟ i<br />

+ ( u υ −uυ<br />

)<br />

(51.20)<br />

k<br />

⎝ ∂x ∂x ⎠<br />

∂x<br />

Now from (49.21) the first term on the right-h<strong>and</strong> side <strong>of</strong> (51.20) is simply [ uv]<br />

, ⋅ w . Since the<br />

coefficient <strong>of</strong> the second tensor on the right-h<strong>and</strong> side <strong>of</strong> (51.20) is skew-symmetric, that term<br />

can be rewritten as<br />

k i wi wk<br />

u υ ⎛ ∂ ∂ ⎞<br />

⎜ −<br />

k i ⎟<br />

⎝∂x<br />

∂x<br />

⎠<br />

or, equivalently,<br />

dw( u,<br />

v )<br />

when the representation (51.12) is used. Thus the identity (51.15) is proved.


Sec. 51 • Differential Forms, Exterior Derivative 377<br />

III If A is an r-form <strong>and</strong> B is an s-form, then<br />

( ) ( )<br />

1 r<br />

d A∧ B = dA∧ B+ − A∧dB (51.21)<br />

This formula can be proved by direct calculation also, so we leave it as an exercise.<br />

IV. If A <strong>and</strong> B are any r-forms, then<br />

( )<br />

d A + B = dA+<br />

dB (51.22)<br />

The pro<strong>of</strong> <strong>of</strong> this result is obvious from the representation (51.5). Combining (51.21) <strong>and</strong> (51.22)<br />

, we have<br />

( )<br />

d αA + βB = αdA+<br />

βdB (51.23)<br />

for any r-forms A <strong>and</strong> B <strong>and</strong> scalars α <strong>and</strong> β.<br />

V. For any r-form A<br />

2<br />

d<br />

( )<br />

A≡ d dA = 0<br />

(51.24)<br />

This result is a consequence <strong>of</strong> the symmetry <strong>of</strong> the second derivative.<br />

Indeed, from (51.5) for the ( r )<br />

+ 1 −formdA we have<br />

∂ A<br />

2<br />

d A = ∧ ∧<br />

! 2 ! 1!<br />

2<br />

1 1 1 lj1j r+ 1 ki1i r i1…<br />

ir<br />

p1 pr+<br />

2<br />

δ<br />

2 p1…<br />

p<br />

δ<br />

r+ 2 j1…<br />

j<br />

g g<br />

r+<br />

1 l k<br />

r r+ ∂x ∂x<br />

( ) ( r + )<br />

( )<br />

1 1 1 ∂ A<br />

= δ<br />

∧ ∧<br />

r! r+ 1 ! r+ 2 ! ∂x ∂x<br />

= 0<br />

( ) ( )<br />

2<br />

lki1<br />

ir<br />

i1…<br />

ir<br />

p1 pr+<br />

2<br />

p1…<br />

p<br />

g g<br />

r+<br />

2 l k<br />

(51.25)<br />

where we have used the identity (20.14).<br />

VI. Let f be a smooth mapping from an open set U in E in<strong>to</strong> an open set U in E,<br />

<strong>and</strong><br />

suppose that A is an r-form on U . Then<br />

where<br />

d<br />

∗<br />

∗<br />

( f ( )) = f ( d )<br />

A A (51.26)<br />

∗<br />

f denotes the induced linear map (defined below) corresponding <strong>to</strong> ,<br />

∗<br />

f so that f ( )<br />

r-form on U , <strong>and</strong> where d denotes the exterior derivative on U . To establish (51.26), let<br />

A is an


378 Chap. 10 • VECTOR FIELDS<br />

dimE = N <strong>and</strong> dim = M<br />

( x … M)<br />

E . We choose coordinate systems ( x i , i 1, , N)<br />

= … <strong>and</strong><br />

α , α = 1, , on U <strong>and</strong> U , respectively. Then the mapping f can be characterized by<br />

( )<br />

i<br />

Suppose that A has the representation (51.4) 2 in ( ).<br />

representation<br />

α<br />

where { g } denotes the natural basis <strong>of</strong> ( x α<br />

).<br />

∗<br />

derivatives <strong>of</strong> A <strong>and</strong> f ( )<br />

representation<br />

x i = f i x 1 ,…, x M , i = 1, …,<br />

N<br />

(51.27)<br />

∗<br />

x Then by definition f ( )<br />

A has the<br />

i1<br />

i<br />

f r<br />

∗ 1<br />

1<br />

r<br />

( A ∂x<br />

∂x<br />

α<br />

α<br />

) = Ai<br />

1 i<br />

∧ ∧<br />

r α1<br />

αr<br />

r!<br />

∂x ∂x<br />

g g<br />

…<br />

(51.28)<br />

Now by direct calculation <strong>of</strong> the exterior<br />

A , <strong>and</strong> by using the symmetry <strong>of</strong> the second derivative, we obtain the<br />

d<br />

∗<br />

∗<br />

( f ( A)<br />

) = f ( dA)<br />

k i1<br />

i<br />

1 ∂A r<br />

i1…<br />

i ∂x ∂x ∂x<br />

r<br />

β α1<br />

= g ∧g ∧∧g<br />

k β α1<br />

αr<br />

r!<br />

∂x ∂x ∂x ∂x<br />

α<br />

r<br />

(51.29)<br />

Thus (51.26) is proved.<br />

Applying the result (51.26) <strong>to</strong> the special case when f is the flow ρt<br />

generated by a vec<strong>to</strong>r<br />

field v as defined in Section 49, we obtain the following property.<br />

VII. The exterior derivative <strong>and</strong> the Lie derivative commute, i.e., for any differential form<br />

A <strong>and</strong> any smooth vec<strong>to</strong>r field v<br />

d<br />

( L ) = L ( d )<br />

v<br />

A A (51.30)<br />

v<br />

We leave the pro<strong>of</strong> <strong>of</strong> (51.30), by direct calculation based on the representations (51.5) <strong>and</strong><br />

(49.42), as an exercise.<br />

The results I-VII summarized above are the basic properties <strong>of</strong> the exterior derivative. In<br />

fact, results I <strong>and</strong> III-V characterize the exterior derivative completely. This converse assertion<br />

can be stated formally in the following way.<br />

Coordinate-free definition <strong>of</strong> the exterior derivative. The exterior derivative d is an opera<strong>to</strong>r<br />

r + 1 −form<strong>and</strong> satisfies the conditions, I, III-V above.<br />

from an r-form <strong>to</strong> an ( )


Sec. 51 • Differential Forms, Exterior Derivative 379<br />

To see that this definition is equivalent <strong>to</strong> the representations (51.5), we consider any r-<br />

form A given by the representation (51.4) 2 . Applying the opera<strong>to</strong>r d <strong>to</strong> both sides <strong>of</strong> that<br />

representation <strong>and</strong> making use <strong>of</strong> conditions I <strong>and</strong> III-V, we get<br />

i1<br />

i<br />

A = d( Ai<br />

)<br />

1…<br />

i<br />

g ∧∧g<br />

r<br />

r<br />

( … ) <br />

r d<br />

!<br />

r<br />

= dA ∧g ∧ ∧ g + A dg ∧g ∧∧g<br />

i1 i i1 i2<br />

ir<br />

i1 ir<br />

i1…<br />

ir<br />

i1 i2<br />

i3<br />

ir<br />

−Ai<br />

1…<br />

i<br />

g ∧dg ∧g ∧∧ g +<br />

r<br />

<br />

⎛∂Ai 1…<br />

ir<br />

k ⎞ i1<br />

ir<br />

= ⎜ g<br />

k ⎟∧g ∧∧ g + 0− 0+<br />

<br />

⎝ ∂x<br />

⎠<br />

∂Ai 1…<br />

ir<br />

k i1<br />

ir<br />

= g ∧g ∧∧g<br />

k<br />

∂x<br />

(51.31)<br />

where we have used the fact that the natural basis vec<strong>to</strong>r<br />

i<br />

function x , so that, by I,<br />

i<br />

g is the gradient <strong>of</strong> the coordinate<br />

i i i<br />

g = grad x = dx<br />

(51.32)<br />

<strong>and</strong> thus, by V,<br />

i 2 i<br />

dg = d x = 0 (51.33)<br />

Consequently (51.5) is equivalent <strong>to</strong> the coordinate-free definition just stated.<br />

Exercises<br />

51.1 Prove the product rule (51.21) for the exterior derivative.<br />

51.2 Given (47.53), the classical definition <strong>of</strong> the curl <strong>of</strong> a vec<strong>to</strong>r field, prove that<br />

2<br />

( dv)<br />

curl v=<br />

D (51.34)<br />

where D is the duality opera<strong>to</strong>r introduced in Section 41.<br />

51.3 Let f be a 0-form. Prove that<br />

( )<br />

Δ f =D d D df<br />

(51.35)<br />

3 1<br />

in the case where dim E =3.<br />

51.4 Let f be the smooth mapping from an open set U in E on<strong>to</strong> an open set U in E<br />

discussed in VI. First, show that


380 Chap. 10 • VECTOR FIELDS<br />

i<br />

∂x<br />

grad f = g<br />

α i<br />

⊗g<br />

∂x<br />

α<br />

∗<br />

Next show that f , which maps r-forms on U in<strong>to</strong> r-forms on U , can be defined by<br />

1 r<br />

1<br />

r<br />

( A)( ,…, ) = A (( grad ) ,…,( grad ) )<br />

∗<br />

f u u f u f u<br />

for all<br />

1 r<br />

u ,…,<br />

u in the translation space <strong>of</strong> E . Note that for r = 1,<br />

grad f .<br />

51.5 Check the commutation relation (51.30) by component representations.<br />

∗<br />

f is the transpose <strong>of</strong>


Sec. 52 • Dural Form <strong>of</strong> Frobenius Theorem; Poincaré Lemma 381<br />

Section 52. The Dual Form <strong>of</strong> the Frobenius Theorem; the Poincaré Lemma<br />

The Frobenius theorem as stated <strong>and</strong> proved in Section 50 characterizes the integrability<br />

<strong>of</strong> a distribution by a condition on the Lie bracket <strong>of</strong> the genera<strong>to</strong>rs <strong>of</strong> the distribution. In this<br />

section we shall characterize the same by a condition on the exterior derivative <strong>of</strong> the genera<strong>to</strong>rs<br />

<strong>of</strong> the orthogonal complement <strong>of</strong> the distribution. This condition constitutes the dual form <strong>of</strong> the<br />

Frobenius theorem.<br />

As in Section 50, let D be a distribution <strong>of</strong> dimension D defined on a domain U in E .<br />

Then at each point x ∈U , ( x)<br />

D ⊥ x <strong>to</strong> be the<br />

orthogonal complement <strong>of</strong> ( x)<br />

D is a D -dimensional subspace <strong>of</strong> V . We put ( )<br />

⊥<br />

D . Then dimD ( x)<br />

= N − D <strong>and</strong> the field<br />

⊥<br />

D on U defined in<br />

this way is itself a distribution <strong>of</strong> dimension N − D.<br />

In particular, locally, there exist 1-forms<br />

{ Γ , Γ= 1, ,N −<br />

⊥<br />

z … D}<br />

which generate D . The dual form <strong>of</strong> the Frobenius theorem is the<br />

following theorem.<br />

Theorem 52.1. The distribution D is integrable if <strong>and</strong> only if<br />

Pro<strong>of</strong>. As in Section 50, let { , 1, ,D}<br />

Γ 1<br />

N−D<br />

dz ∧z ∧ ∧ z = 0, Γ= 1, , N −D<br />

… (52.1)<br />

v<br />

α<br />

α = … be a local basis for D . Then<br />

Γ<br />

v ⋅ z = 0, α = 1, …, D, Γ= 1, …,<br />

N −D<br />

(52.2)<br />

α<br />

By the Frobenius theorem D is integrable if <strong>and</strong> only if (50.5) holds. Substituting (50.5) <strong>and</strong><br />

Γ<br />

(52.2) in<strong>to</strong> (51.15) with u= v , v = v , <strong>and</strong> w = z , we see that (50.5) is equivalent <strong>to</strong><br />

α<br />

( )<br />

β<br />

Γ<br />

dz v , v = 0, , = 1, …, D, Γ= 1, …,<br />

N −D<br />

(52.3)<br />

α β<br />

αβ<br />

We now show that this condition is equivalent <strong>to</strong> (52.1).<br />

To prove the said equivalence, we extend { } α<br />

dual basis { i<br />

}<br />

v in<strong>to</strong> a basis { }<br />

v in such a way that its<br />

D+Γ Γ<br />

v satisfies v = z for all Γ= 1, …, N −D.<br />

This extension is possibly by virtue <strong>of</strong><br />

(52.2). Relative <strong>to</strong> the basis { v i }, the condition (52.3) means that in the representation <strong>of</strong> d z<br />

Γ<br />

by<br />

Γ Γ i j Γ i j<br />

2d z = ζ v ∧ v = 2ζ<br />

v ⊗v (52.4)<br />

ij<br />

ij<br />

i<br />

the components ζ Γ ij<br />

with 1 ≤i,<br />

j ≤ D must vanish. Thus


382 Chap. 10 • VECTOR FIELDS<br />

i<br />

2d z Γ = ζ Γ Δ<br />

i D<br />

v ∧ z (52.5)<br />

Δ+<br />

which is clearly equivalent <strong>to</strong> (52.1).<br />

As an example <strong>of</strong> the integrability condition (52.1), we can derive a necessary <strong>and</strong><br />

sufficient condition for a vec<strong>to</strong>r field z <strong>to</strong> be orthogonal <strong>to</strong> a family <strong>of</strong> surfaces. That is, there<br />

exist smooth functions h <strong>and</strong> f such that<br />

z = h grad f<br />

(52.6)<br />

Such a vec<strong>to</strong>r field is called complex-lamellar in the classical theory. Using the terminology<br />

⊥<br />

here, we see that a complex-lamellar vec<strong>to</strong>r field z is a vec<strong>to</strong>r field such that z is integrable.<br />

Hence by (52.1), z is complex-lamellar if <strong>and</strong> only if dz∧ z = 0 . In a three-dimensional space<br />

this condition can be written as<br />

( )<br />

curl z ⋅ z = 0<br />

(52.7)<br />

which was first noted by Kelvin (1851). In component form, if z is represented by<br />

i i<br />

z = zig = zdx<br />

i<br />

(52.8)<br />

then dz is represented by<br />

1 ∂zi<br />

k i<br />

dz = dx ∧dx<br />

(52.9)<br />

k<br />

2 ∂x<br />

<strong>and</strong> thus dz∧ z = 0 means<br />

∂zi<br />

k i j<br />

zj<br />

dx ∧ dx ∧ dx = 0<br />

k<br />

∂x<br />

(52.10)<br />

or, equivalently,<br />

δ<br />

∂zi<br />

z = 0, p, q, r = 1, …,<br />

N<br />

(52.11)<br />

∂x<br />

kij<br />

pqr j k<br />

Since it suffices <strong>to</strong> consider the special cases with p < q< r in (52.11), when N=3 we have<br />

kij ∂zi<br />

0 = δ123<br />

zj<br />

= z<br />

k jε<br />

∂x<br />

kij<br />

∂z<br />

∂x<br />

i<br />

k<br />

(52.12)


Sec. 52 • Dural Form <strong>of</strong> Frobenius Theorem; Poincaré Lemma 383<br />

which is equivalent <strong>to</strong> the component version <strong>of</strong> (52.7). In view <strong>of</strong> this special case we see that<br />

the dual form <strong>of</strong> the Frobenius theorem is a generalization <strong>of</strong> Kelvin’s condition from 1-forms <strong>to</strong><br />

r-forms in general.<br />

In the classical theory a vec<strong>to</strong>r field (1-form) z is called lamellar if locally z is the gradient<br />

<strong>of</strong> a smooth function (0-form) f, namely<br />

z = grad f = df<br />

(52.13)<br />

Then it can be shown that z is lamellar if <strong>and</strong> only if dz vanishes, i.e.,<br />

curl z=<br />

0 (52.14)<br />

The generalization <strong>of</strong> lamellar fields from 1-forms <strong>to</strong> r-forms in general is obvious: We say that<br />

an r-form A is closed if its exterior derivative vanishes<br />

d A = 0<br />

(52.15)<br />

<strong>and</strong> we say that A is exact if it is the exterior derivatives <strong>of</strong> an (r-1)-form, say<br />

A = dB (52.16)<br />

From (51.24) any exact form is necessarily closed. The theorem that generalizes the classical<br />

result is the Poincaré lemma, which implies that (52.16) <strong>and</strong> (52.15) are locally equivalent.<br />

Before stating the Poincaré lemma, we defined first the notion <strong>of</strong> a star-shaped (or<br />

retractable) open set U in E : U is star-shaped if there exists a smooth mapping<br />

ρ : U × R →U (52.17)<br />

such that<br />

( , t)<br />

ρ x<br />

⎧x<br />

= ⎨<br />

⎩x<br />

0<br />

when t ≤ 0<br />

when t ≥1<br />

(52.18)<br />

where x<br />

0<br />

is a particular point in U . In a star-shaped open set U all closed hypersurfaces <strong>of</strong><br />

dimensions 1 <strong>to</strong> N − 1 are retractable in the usual sense. For example, an open ball is starshaped<br />

but the open set between two concentric spheres is not, since a closed sphere in the latter<br />

is not retractable.


384 Chap. 10 • VECTOR FIELDS<br />

The equivalence <strong>of</strong> (52.13) <strong>and</strong> (52.14) requires only that the domain U <strong>of</strong> z be simply<br />

connected, i.e., every closed curve in U be retractable. When we generalize the result <strong>to</strong> the<br />

equivalence <strong>of</strong> (52.15) <strong>and</strong> (52.16), the domain U <strong>of</strong> A must be retractable relative <strong>to</strong> all r-<br />

dimensional closed hypersurfaces. For simplicity, we shall assume that U is star-shaped.<br />

Theorem 52.2. If A is a closed r-form defined on a star-shaped domain, then A is exact, i.e.,<br />

there exists an (r-1)-form B on U such that (52.16) holds.<br />

i<br />

Pro<strong>of</strong>. We choose a coordinate system ( x )on U . Then a coordinate system<br />

( y<br />

α , α = 1, , N + 1)<br />

… on U × R is given by<br />

1 1 1<br />

( , , N +<br />

N<br />

y y ) = ( x , , x , t)<br />

… … (52.19)<br />

From (52.18) the mapping ρ has the property<br />

<strong>and</strong><br />

( )<br />

ρ ⋅ , t = idU when t ≤0<br />

(52.20)<br />

( )<br />

ρ ⋅ , t = x when t ≥0<br />

(52.21)<br />

o<br />

Hence if the coordinate representation <strong>of</strong> ρ is<br />

i i α<br />

x = ρ ( y ), i = 1, …,<br />

N<br />

(52.22)<br />

Then<br />

i<br />

i<br />

∂ρ<br />

i ∂ρ<br />

= δ<br />

j, = 0 when t ≤0<br />

N + 1<br />

∂y<br />

∂y<br />

j<br />

(52.23)<br />

<strong>and</strong><br />

i<br />

i<br />

∂ρ<br />

∂ρ<br />

= 0 = 0 when t ≥1<br />

j N+<br />

1<br />

∂y<br />

∂y<br />

(52.24)<br />

i<br />

As usual we can express A in component form relative <strong>to</strong> ( x )<br />

1<br />

i1<br />

ir<br />

A = A i 1…<br />

i<br />

g ∧∧g<br />

(52.25)<br />

r<br />

r!


Sec. 52 • Dural Form <strong>of</strong> Frobenius Theorem; Poincaré Lemma 385<br />

ρ A is an r-form on U × R defined by<br />

∗<br />

Then ( )<br />

i1<br />

ir<br />

∗ 1 ∂ρ<br />

∂ρ<br />

α1<br />

( ) = A i 1… i<br />

∧∧<br />

r α1<br />

αr<br />

ρ αr<br />

A<br />

r!<br />

∂y ∂y<br />

h h<br />

(52.26)<br />

where { α , α = 1, , N + 1}<br />

h … denotes the natural basis <strong>of</strong> ( y α<br />

),<br />

i.e.,<br />

α α<br />

h = dy<br />

(52.27)<br />

where d denotes the exterior derivative on U × R .<br />

Now from (52.19) we can rewrite (52.26) as<br />

1 r<br />

1 r 1<br />

( ) <br />

<br />

−<br />

r! ρ A = X h ∧ ∧ h + rY<br />

h ∧ ∧h ∧h<br />

(52.28)<br />

∗ i i i i N + 1<br />

i1…<br />

ir<br />

i1…<br />

ir−1<br />

where<br />

X<br />

j1<br />

jr<br />

∂ρ<br />

∂ρ<br />

= A<br />

∂y<br />

… … ∂y<br />

(52.29)<br />

i1 ir j1 jr i1<br />

ir<br />

<strong>and</strong><br />

j1 jr−1<br />

jr<br />

∂ρ ∂ρ ∂ρ<br />

Yi 1… i<br />

= A<br />

r−1 j1… j<br />

(52.30)<br />

r i1 ir<br />

− 1 N + 1<br />

∂y ∂y ∂y<br />

We put<br />

1<br />

i1<br />

ir<br />

X = X i 1…<br />

i<br />

h ∧∧h<br />

(52.31)<br />

r<br />

r!<br />

<strong>and</strong><br />

1<br />

1!<br />

j1 jr<br />

−1<br />

Y = Y j 1…<br />

j<br />

h ∧∧h<br />

(52.32)<br />

r−1<br />

( r − )<br />

Then X is an r-form <strong>and</strong> Y is an ( r −1) − form on U × R . Further, from (52.28) we have<br />

ρ<br />

N 1<br />

( ) h<br />

A = X+ Y∧ = X+ Y ∧dt<br />

(52.33)<br />

∗ +


386 Chap. 10 • VECTOR FIELDS<br />

From (52.23), (52.24), (52.29), <strong>and</strong> (52.30), X <strong>and</strong> Y satisfy the end conditions<br />

<strong>and</strong><br />

( x ) ( x) ( x )<br />

X<br />

…<br />

,t = A<br />

…<br />

, Y<br />

…<br />

,t = 0 when t ≤0<br />

(52.34)<br />

i1 ir i1 ir i1 ir−1<br />

( x ) ( x )<br />

X ,t = 0, Y ,t = 0 when t ≥ 1<br />

… …<br />

(52.35)<br />

i1 ir<br />

i1 ir−1<br />

for all x ∈U . We define<br />

( −1)<br />

( r )<br />

1<br />

1<br />

( Yi<br />

( , )<br />

0 1…<br />

i<br />

t dt<br />

r−1<br />

)<br />

≡ ⋅ ∧ ∧<br />

−1!<br />

∫ g g<br />

i<br />

ir−1<br />

B <br />

(52.36)<br />

<strong>and</strong> we claim that this ( r −1)<br />

-form B satisfies the condition (52.16).<br />

To prove this, we take the exterior derivative <strong>of</strong> (52.33). By (52.26) <strong>and</strong> the fact that A is<br />

closed, the result is<br />

dX+ dY ∧ dt =0<br />

(52.37)<br />

where we have used also (51.21) <strong>and</strong> (51.24) on the term<br />

exterior derivatives <strong>of</strong> X<strong>and</strong><br />

Yhave the representations<br />

Y ∧ dt . From (52.31) <strong>and</strong> (52.32) the<br />

∂Xi 1 i<br />

∂X<br />

r j i1 ir<br />

i1ir<br />

i1<br />

ir<br />

rd ! X = h ∧h ∧∧ h + dt∧h ∧∧h<br />

(52.38)<br />

j<br />

∂x<br />

∂t<br />

<strong>and</strong><br />

∂Y<br />

∂Y<br />

r− d = ∧ ∧ ∧ + dt∧ ∧ ∧<br />

∂x<br />

∂t<br />

i1ir−1 j<br />

r−<br />

i1ir−1<br />

( 1! ) <br />

Substituting these in<strong>to</strong> (52.37), we then get<br />

i1 i 1 i1<br />

ir<br />

Y h h h h h (52.39)<br />

j<br />

∂X<br />

i1 ir<br />

j i1<br />

ir<br />

h ∧ h ∧ ∧ h = 0<br />

j<br />

∂x<br />

(52.40)<br />

<strong>and</strong>


Sec. 52 • Dural Form <strong>of</strong> Frobenius Theorem; Poincaré Lemma 387<br />

⎛ r ∂Xi 1 i<br />

∂Y<br />

r i2 i ⎞<br />

r i1<br />

ir<br />

⎜( − 1) + r ∧ ∧ ∧ dt =<br />

i1<br />

⎟h h 0<br />

(52.41)<br />

⎝ ∂t<br />

∂x<br />

⎠<br />

Consequently<br />

−<br />

1 ∂Y<br />

1 r i<br />

( )<br />

2 i r ∂X<br />

i i<br />

r<br />

j<br />

1<br />

1 jr<br />

δ<br />

j1… j<br />

= −<br />

(52.42)<br />

r i1<br />

r−1!<br />

∂x ∂t<br />

( )<br />

result is<br />

Now from (52.36) if we take the exterior derivative <strong>of</strong> the ( r −1)<br />

-form B on U , the<br />

r<br />

( 1)<br />

( r−<br />

)<br />

∂Y<br />

dB<br />

= ∧ ∧<br />

1! ⎝ ∂x<br />

⎠<br />

− ⎛ 1<br />

i2 i ⎞<br />

r i1<br />

ir<br />

⎜∫<br />

dt<br />

0 i ⎟g<br />

g<br />

1<br />

r<br />

( )<br />

( − )<br />

−1 1 ⎛ ∂Y<br />

⎞<br />

= δ<br />

∧ ∧<br />

r 1! r!<br />

⎝ ∂x<br />

⎠<br />

1<br />

i1 ir<br />

i2ir<br />

j1<br />

jr<br />

j1 … j<br />

dt<br />

r ⎜∫0<br />

i1<br />

⎟g<br />

g<br />

(52.43)<br />

Substituting (52.42) in<strong>to</strong> this equation yields<br />

1 ⎛ 1 ∂X<br />

⎞<br />

dB<br />

= ⎜− dt<br />

r!<br />

∫0<br />

⎟g ⎝ ∂t<br />

⎠<br />

∧∧g<br />

1<br />

( X<br />

j j ( ,0) X<br />

j j<br />

r!<br />

( ,1))<br />

1<br />

j1<br />

jr<br />

= Aj<br />

1 j<br />

g ∧∧ g = A<br />

r<br />

r!<br />

j1<br />

jr<br />

j1<br />

jr<br />

j1<br />

= ⋅ − ⋅ g ∧∧<br />

1<br />

r<br />

1<br />

r<br />

g<br />

jr<br />

(52.44)<br />

where we have used the end conditions (52.34) <strong>and</strong> (52.35) on X .<br />

course. Since<br />

The ( r −1)<br />

-form B , whose existence has just been proved by (52.44), is not unique <strong>of</strong><br />

ˆ<br />

( )<br />

B is unique <strong>to</strong> within an arbitrary closed ( r −1)<br />

-form on U .<br />

Exercises<br />

52.1 In calculus a “differential”<br />

dB= dBˆ<br />

⇔d<br />

B− B =0<br />

(52.45)


388 Chap. 10 • VECTOR FIELDS<br />

( , , ) ( , , ) ( , , )<br />

P x y z dx + Q x y z dy + R x y z dz<br />

(52.46)<br />

is called exact if there exists a function U( x, y,<br />

z ) such that<br />

∂U ∂U ∂U<br />

dU = dx + dy + dz<br />

∂x ∂y ∂z<br />

= Pdx+ Qdy+<br />

Rdz<br />

(52.47)<br />

Use the Poincaré lemma <strong>and</strong> show that (52.46) is exact if <strong>and</strong> only if<br />

∂P ∂Q ∂P ∂R ∂Q ∂R<br />

= , = , =<br />

∂y ∂x ∂z ∂x ∂z ∂y<br />

(52.48)<br />

52.2 The “differential” (52.46) is called integrable if there exists a nonvanishing function<br />

μ x, yz , , called an integration fac<strong>to</strong>r, such that the differential<br />

( )<br />

μ ( Pdx+ Qdy+ Rdz)<br />

(52.49)<br />

is exact. Show that (52.46) is integrable if <strong>and</strong> only if the two-dimensional distribution<br />

orthogonal <strong>to</strong> the 1-form (52.46) is integrable in the sense defined in this section. Then<br />

use the dual form <strong>of</strong> the Frobenius theorem <strong>and</strong> show that (52.46) is integrable if <strong>and</strong><br />

only if<br />

⎛∂R ∂Q⎞ ⎛ ∂P ∂R⎞<br />

⎛∂Q ∂P⎞<br />

P ⎜ − + Q⎜<br />

− ⎟ + R − = 0<br />

∂y ∂z ⎟<br />

∂z ∂x ⎜<br />

∂x ∂y<br />

⎟<br />

⎝ ⎠ ⎝ ⎠ ⎝ ⎠<br />

(52.50)


Sec. 53 • Three-Dimensional Euclidean Manifold, I 389<br />

Section 53. Vec<strong>to</strong>r Fields in a Three-Dimensional Euclidean Manifold, I.<br />

Invariants <strong>and</strong> Intrinsic Equations<br />

The preceding four sections <strong>of</strong> this chapter concern vec<strong>to</strong>r fields, distributions, <strong>and</strong><br />

differential forms, defined on domains in an N-dimensional Euclidean manifold E in general. In<br />

applications, <strong>of</strong> course, the most important case is when E is three-dimensional. Indeed,<br />

classical vec<strong>to</strong>r analysis was developed just for this special case. In this section we shall review<br />

some <strong>of</strong> the most important results in the classical theory from the more modern point <strong>of</strong> view, as<br />

we have developed so far in this text.<br />

We recall first that when E is three-dimensional the exterior product may be replaced by<br />

the cross product [cf.(41.19)]. Specially, relative <strong>to</strong> a positive orthonormal basis<br />

e , e , e for V , the component representation <strong>of</strong> u×<br />

v for any uv , ,∈V is<br />

{ }<br />

1 2 3<br />

u× v = ε e<br />

uv i j k<br />

ijk<br />

( u 2 υ 3 u 3 υ 2 ) ( u 3 υ 1 u 1 υ 3 ) ( u 1 υ 2 u<br />

2 υ<br />

1<br />

)<br />

= − e + − e + − e<br />

1 2 3<br />

(53.1)<br />

Where<br />

are the usual component representations <strong>of</strong><br />

i<br />

j<br />

u= u e , v = υ e (53.2)<br />

i<br />

j<br />

u <strong>and</strong> v<strong>and</strong> where the reciprocal basis { e<br />

i<br />

} coincides<br />

with { e i } . It is important <strong>to</strong> note that the representation (53.1) is valid relative <strong>to</strong> a positive<br />

orthonormal basis only; if the orthonormal basis { e i } is negative, the signs on the right-h<strong>and</strong> side<br />

<strong>of</strong> (53.1) must be reversed. For this reason, u×<br />

v is called an axial vec<strong>to</strong>r in the classical theory.<br />

We recall also that when E is three-dimensional, then the curl <strong>of</strong> a vec<strong>to</strong>r field can be<br />

represented by a vec<strong>to</strong>r field [cf. (47.53) or (51.34)]. Again, if a positive rectangular Cartesian<br />

1 2 3<br />

, ,<br />

e is used, then curl v has the components<br />

coordinate system ( )<br />

x x x induced by { }<br />

i<br />

∂υ<br />

j<br />

curl v = εijk<br />

e<br />

i k<br />

∂x<br />

⎛∂υ ∂υ ⎞ ⎛∂υ ∂υ ⎞ ⎛∂υ ∂υ<br />

⎞<br />

= ⎜ − ⎟e + ⎜ − ⎟e + ⎜ − ⎟e<br />

⎝∂x ∂x ⎠ ⎝∂x ∂x ⎠ ⎝ ∂x ∂x<br />

⎠<br />

3 2 1 3 2 1<br />

2 3 1 3 1 2 1 2 3<br />

(53.3)<br />

where v is now a smooth vec<strong>to</strong>r field having the representation<br />

i i<br />

v = υie = υ e<br />

i<br />

(53.4)


390 Chap. 10 • VECTOR FIELDS<br />

i<br />

Since ( x ) is rectangular Cartesian, the natural basis vec<strong>to</strong>rs <strong>and</strong> the component fields satisfy the<br />

usual conditions<br />

i<br />

i<br />

υ = υ i<br />

, e = e i<br />

, i = 1,2,3<br />

(53.5)<br />

By the same remark as before, curl v is an axial vec<strong>to</strong>r field, so the signs on the right-h<strong>and</strong> side<br />

i<br />

must be reversed when ( x ) is negative.<br />

Now consider a nonvanishing smooth vec<strong>to</strong>r field v . We put<br />

s=<br />

v/<br />

v (53.6)<br />

Then s is a unit vec<strong>to</strong>r field in the same direction as v . In Section 49 we have introduced the<br />

notions <strong>of</strong> internal curves <strong>and</strong> flows corresponding <strong>to</strong> any smooth vec<strong>to</strong>r field. We now apply<br />

these <strong>to</strong> the vec<strong>to</strong>r field s . Sinces is a unit vec<strong>to</strong>r, its integral curves are parameterized by the<br />

arc length 1 s. A typical integral curve <strong>of</strong> s is<br />

( s)<br />

λ = λ (53.7)<br />

where<br />

( ( ))<br />

at each point λ ( s)<br />

<strong>of</strong> the curve. From (53.6) <strong>and</strong> (53.8) we have<br />

dλ / ds = s λ s<br />

(53.8)<br />

dλ<br />

ds<br />

( ( s)<br />

)<br />

parallel <strong>to</strong> v λ (53.9)<br />

but generally dλ / ds is not equal <strong>to</strong> v ( λ ( s)<br />

), so λ is not an integral curve <strong>of</strong> v as defined in<br />

Section 49. In the classical theory the locus <strong>of</strong> λ without any particular parameterization is<br />

called a vec<strong>to</strong>r line <strong>of</strong> v .<br />

Now assuming that λ is not a straight line, i.e., s is not invariant on λ , we can take the<br />

covariant derivative <strong>of</strong> s on λ (cf. Section 48 <strong>and</strong> write the result in the form<br />

1 Arc length shall be defined in a general in Section 68. Here s can be regarded as a parameter such that<br />

dλ<br />

/ ds = 1.


Sec. 53 • Three-Dimensional Euclidean Manifold, I 391<br />

ds/<br />

ds = κn (53.10)<br />

where κ <strong>and</strong> n are called the curvature <strong>and</strong> the principal normal <strong>of</strong> λ , <strong>and</strong> they are characterized<br />

by the condition<br />

ds<br />

κ = ><br />

ds<br />

0<br />

(53.11)<br />

It should be noted that (53.10) defines both κ <strong>and</strong> n : κ is the norm <strong>of</strong> ds<br />

/ ds<strong>and</strong> n is the unit<br />

vec<strong>to</strong>r in the direction <strong>of</strong> the nonvanishing vec<strong>to</strong>r field ds / ds. The reciprocal <strong>of</strong> κ ,<br />

r = 1/ κ<br />

(53.12)<br />

is called the radius <strong>of</strong> curvature <strong>of</strong> λ .<br />

Since s is a vec<strong>to</strong>r, we have<br />

or, equivalently,<br />

( s⋅s)<br />

d ds<br />

0= = 2s⋅ = 2κ<br />

s⋅n (53.13)<br />

ds ds<br />

s⋅ n = 0<br />

(53.14)<br />

Thus n is normal <strong>to</strong> s , as it should be. In view <strong>of</strong> (53.14) the cross product <strong>of</strong><br />

vec<strong>to</strong>r<br />

s with n is a unit<br />

b≡ s×<br />

n (53.15)<br />

which is called the binormal <strong>of</strong> λ . The triad { snb , , } now forms a field <strong>of</strong> positive orthonormal<br />

basis in the domain <strong>of</strong> v . In general { snb , , } is anholonomic, <strong>of</strong> course.<br />

Now we compute the covariant derivative <strong>of</strong> n <strong>and</strong> balong the curve λ . Since b is a<br />

unit vec<strong>to</strong>r, by the same argument as (53.13) we have<br />

d b 0<br />

ds ⋅ b =<br />

(53.16)<br />

Similarly, since bs ⋅ = 0, on differentiating with respect <strong>to</strong> s along λ , we obtain


392 Chap. 10 • VECTOR FIELDS<br />

db<br />

ds<br />

⋅ s=−b⋅ =−κ<br />

b⋅ n = 0<br />

(53.17)<br />

ds ds<br />

where we have used (53.10). Combining (53.16) <strong>and</strong> (53.17), we see that db / ds is parallel <strong>to</strong> n ,<br />

say<br />

db<br />

= −τ n (53.18)<br />

ds<br />

where τ is called the <strong>to</strong>rsion <strong>of</strong> the curve λ . From (53.10) <strong>and</strong> (53.18) the gradient <strong>of</strong> n along<br />

λ can be computed easily by the representation<br />

n= b×<br />

s (53.19)<br />

so that<br />

dn ds db<br />

= b× + × s= b× κn− τn× s=− κs+<br />

τb (53.20)<br />

ds ds ds<br />

The results (53.10), (53.18), <strong>and</strong> (53.20) are the Serret-Frenet formulas for the curve λ .<br />

So far we have introduced a field <strong>of</strong> basis { snb , , } associated with the nonzero <strong>and</strong><br />

nonrectilinear vec<strong>to</strong>r field v . Moreover, the Serret-Frenet formulas give the covariant derivative<br />

<strong>of</strong> the basis along the vec<strong>to</strong>r lines <strong>of</strong> v . In order <strong>to</strong> make full use <strong>of</strong> the basis, however, we need<br />

a complete representation <strong>of</strong> the covariant derivative <strong>of</strong> that basis along all curves. Then we can<br />

express the gradient <strong>of</strong> the vec<strong>to</strong>r fields snb , , in component forms relative <strong>to</strong> the anholonomic<br />

basis { snb , , }. These components play the same role as the Chris<strong>to</strong>ffel symbols for a holonomic<br />

basis. The component forms <strong>of</strong> grad s , grad n <strong>and</strong> grad b have been derived by Bjørgum. 2 We<br />

shall summarize his results without pro<strong>of</strong>s here.<br />

Bjørgum shows first that the components <strong>of</strong> the vec<strong>to</strong>r fields curls , curln , <strong>and</strong> curlb<br />

snb , , are given by<br />

relative <strong>to</strong> the basis { }<br />

curls= Ω s+<br />

κb<br />

s<br />

( div )<br />

( κ div )<br />

curln=− b s+Ω n+<br />

θb<br />

curlb= + n s+ ηn+Ω<br />

b<br />

n<br />

b<br />

(53.21)<br />

where Ωs, Ωn,<br />

Ωb are given by<br />

2 O. Bjørgum, “On Beltrami Vec<strong>to</strong>r Fields <strong>and</strong> Flows, Part I. A Comparative Study <strong>of</strong> Some Basic Types <strong>of</strong> Vec<strong>to</strong>r<br />

Fields,” Universitetet I Bergen, Arbok 1951, Naturvitenskapelig rekke Nr. 1.


Sec. 53 • Three-Dimensional Euclidean Manifold, I 393<br />

s⋅ curl s=Ω , n⋅ curl n=Ω , b⋅ curlb =Ω<br />

(53.22)<br />

s n b<br />

<strong>and</strong> are called the abnormality <strong>of</strong> snb, , , respectively. From Kelvin’s theorem (cf. Section 52)<br />

we know that the abnormality measures, in some sense, the departure <strong>of</strong> a vec<strong>to</strong>r field from a<br />

complex-lamellar field. Since b is given by (53.15), the abnormalities Ωs, Ωn,<br />

Ωbare not<br />

independent. Bjørgum shows that<br />

Ω +Ω =Ω −2τ<br />

n b s<br />

(53.23)<br />

where τ is the <strong>to</strong>rsion <strong>of</strong> λ as defined by (53.18). The quantities θ <strong>and</strong> η in (53.21) are defined<br />

by<br />

<strong>and</strong> Bjørgum shows that<br />

b⋅ curl n= θ, n⋅ curlb = η<br />

(53.24)<br />

θ − η = div s (53.25)<br />

Notice that (53.21) implies that<br />

n⋅ curls = 0<br />

(53.26)<br />

but the remaining eight components in (53.21) are generally nonzero.<br />

Next Bjørgum shows that the components <strong>of</strong> the second-order tensor fields grad s ,<br />

grad n , <strong>and</strong> grad b relative <strong>to</strong> the basis { snb , , } are given by<br />

κ θ ( n<br />

τ)<br />

( b<br />

τ)<br />

b n ηb b<br />

κ θ ( n<br />

τ)<br />

τ ( div ) ( κ div )<br />

( b<br />

τ)<br />

η τ<br />

( div b) n n ( κ div n)<br />

n b<br />

grad s= n⊗ s+ n− Ω + n⊗b<br />

+ Ω + ⊗ − ⊗<br />

grad n=− s⊗s− s⊗ n+ Ω + s⊗b<br />

+ b⊗s− b b⊗ n+ + n b⊗b<br />

grad b=− Ω + s⊗ n+ s⊗b− n⊗s<br />

+ ⊗ − + ⊗<br />

(53.27)<br />

These representations are clearly consistent with the representations (53.21) through the general<br />

formula (47.53). Further, from (48.15) the covariant derivatives <strong>of</strong> sn , , <strong>and</strong> b along the integral<br />

curve λ <strong>of</strong> s are


394 Chap. 10 • VECTOR FIELDS<br />

ds<br />

= ( grad ss ) = κn<br />

ds<br />

dn<br />

= ( grad ns ) =− κs+<br />

τb<br />

ds<br />

db<br />

= ( grad bs ) = −τ<br />

n<br />

ds<br />

(53.28)<br />

which are consistent with the Serret-Frenet formulas (53.10), (53.20), <strong>and</strong> (53.18).<br />

The representations (53.27) tell us also the gradients <strong>of</strong> { snb , , } along any integral curves<br />

( n) <strong>of</strong> <strong>and</strong> ( b)<br />

<strong>of</strong> .<br />

μ= μ n ν = ν b Indeed, we have<br />

( ) θ ( b<br />

τ)<br />

( ) θ ( )<br />

( ) ( τ ) ( )<br />

ds/ dn = grads n= n+ Ω + b<br />

dn/ dn = gradn n= − s−<br />

divb b<br />

db/ dn = gradb n= − Ω + s+<br />

divb n<br />

b<br />

(53.29)<br />

<strong>and</strong><br />

( ) ( n<br />

τ)<br />

η<br />

( ) ( n<br />

τ) ( κ )<br />

( ) η ( κ )<br />

ds/ db= grads b= − Ω + n−<br />

b<br />

dn/ db= gradn b= Ω + s+ + divn b<br />

db/ db= gradb b= s− + divn n<br />

(53.30)<br />

Since the basis { snb , , } is anholonomic in general, the parameters ( , , )<br />

snb are not local<br />

coordinates. In particular, the differential opera<strong>to</strong>rs d / ds, d / dn, d / db do not commute. We<br />

derive first the commutation formulas 3 for a scalar function f.<br />

From (47.14) <strong>and</strong> (46.10) we verify easily that the anholonomic representation <strong>of</strong> grad f is<br />

grad<br />

f<br />

df df df<br />

= s+ n+<br />

b (53.31)<br />

ds dn db<br />

for any smooth function f defined on the domain <strong>of</strong> the basis { snb , , }. Taking the gradient <strong>of</strong><br />

(53.31) <strong>and</strong> using (53.27), we obtain<br />

3 See A.W. Marris <strong>and</strong> C.-C. Wang, “Solenoidal Screw Fields <strong>of</strong> Constant Magnitude,” Arch. Rational Mech. Anal.<br />

39, 227-244 (1970).


Sec. 53 • Three-Dimensional Euclidean Manifold, I 395<br />

⎛ df ⎞ df ⎛ df ⎞<br />

grad ( grad f ) = s⊗ ⎜grad ⎟+ ( grad s)<br />

+ n⊗⎜grad<br />

⎟<br />

⎝ ds ⎠ ds ⎝ dn ⎠<br />

df ⎛ df ⎞ df<br />

+ ( grad n) + b⊗ ⎜grad ⎟+<br />

( grad b)<br />

dn ⎝ db ⎠ db<br />

⎡ d df df ⎤<br />

=<br />

⎢<br />

−κ<br />

ds ds dn ⎥<br />

s⊗s<br />

⎣<br />

⎦<br />

⎡ d df df df ⎤<br />

+<br />

⎢<br />

−θ<br />

−( Ω<br />

b<br />

+ τ)<br />

dn ds dn db⎥<br />

s⊗n<br />

⎣<br />

⎦<br />

⎡ d df df df ⎤<br />

+<br />

⎢<br />

+ ( Ω<br />

n<br />

+ τ)<br />

+ η ⊗<br />

⎣db ds dn db⎥<br />

s b<br />

⎦<br />

⎡ df d df df ⎤<br />

+<br />

⎢<br />

κ + −τ<br />

⎣ ds ds dn db⎥<br />

n⊗<br />

s<br />

⎦<br />

⎡ df d df df ⎤<br />

+<br />

⎢<br />

θ + + ( div b)<br />

dn dn dn db⎥<br />

n⊗n<br />

⎣<br />

⎦<br />

⎡ df d df df ⎤<br />

+<br />

⎢<br />

−( Ω<br />

n<br />

+ τ) + − ( κ + div n)<br />

ds db dn db⎥<br />

n⊗b<br />

⎣<br />

⎦<br />

⎡ df d df ⎤<br />

+<br />

⎢<br />

τ +<br />

dn ds db⎥<br />

b⊗s<br />

⎣<br />

⎦<br />

⎡ df df d df ⎤<br />

+<br />

⎢( Ω<br />

b<br />

+ τ ) − ( div b)<br />

+ ⊗<br />

⎣ ds dn dn db⎥<br />

b n<br />

⎦<br />

⎡ df df d df ⎤<br />

+<br />

⎢<br />

− η + ( κ + div n)<br />

+ ⊗<br />

⎣ ds dn db db⎥<br />

b b<br />

⎦<br />

(53.32)<br />

Now since grad ( grad f ) is symmetric, (53.32) yields<br />

d df d df df df df<br />

− = κ + θ +Ωb<br />

ds ds ds dn ds dn db<br />

d df d df df df<br />

− =Ω<br />

n<br />

+ η<br />

ds db db ds dn db<br />

d df d df df df df<br />

− =Ωs<br />

− ( div b) + ( κ + div n)<br />

db dn dn db ds dn db<br />

(53.33)<br />

where we have used (53.23). The Formulas (53.33) 1-3 are the desired commutation rules.<br />

Exactly the same technique can be applied <strong>to</strong> the identities<br />

( ) ( ) ( )<br />

curl grad s = curl grad n = curl grad b = 0 (53.34)


396 Chap. 10 • VECTOR FIELDS<br />

<strong>and</strong> the results are the following nine intrinsic equations 4 for the basis { snb , , }:<br />

d dθ<br />

− ( Ω<br />

n<br />

+ τ) − + ( κ + div n)( Ωb −Ωn) − ( θ + η)<br />

div b+Ω sκ<br />

= 0<br />

dn db<br />

dη<br />

d<br />

− − ( Ω<br />

b<br />

+ τ) −( Ωb −Ωn) div b− ( θ + η)( κ + div n)<br />

= 0<br />

dn db<br />

dκ<br />

d<br />

+ ( Ω<br />

n<br />

+ τ) −η( Ωs −Ω<br />

b)<br />

+Ω<br />

nθ<br />

= 0<br />

db ds<br />

dτ<br />

d<br />

− ( κ + div n) −Ω<br />

n<br />

div b+ η( 2κ<br />

+ div n)<br />

= 0<br />

db ds<br />

dη η<br />

2 κ<br />

2<br />

− + − ( κ + div n) − τ −Ωn ( Ωs −Ω<br />

n)<br />

= 0<br />

ds<br />

dκ<br />

dθ<br />

2 2<br />

− −κ − θ + ( 2Ωs<br />

− 3τ) +Ωn( Ωs −Ωn<br />

− 4τ)<br />

= 0<br />

dn ds<br />

dτ<br />

d<br />

+ div b−κ( Ωs −Ω<br />

n ) + θ div b−Ω b( κ + div n)<br />

= 0<br />

dn ds<br />

d<br />

d<br />

2 2<br />

( κ + div n) + div b− θn<br />

+ ( div b) + ( κ + div n)<br />

dn<br />

db<br />

+Ω τ + Ω + τ Ω + τ = 0<br />

( )( )<br />

s n b<br />

dΩ s<br />

dκ<br />

+ +Ω<br />

s ( θ − η)<br />

+ κdiv b = 0<br />

ds db<br />

(53.35)<br />

Having obtained the representations (53.21) <strong>and</strong> (53.27), the commutation rules (53.33),<br />

<strong>and</strong> the intrinsic equations (53.35) for the basis { snb , , }, we can now return <strong>to</strong> the original<br />

relations (53.6) <strong>and</strong> derive various representations for the invariants <strong>of</strong> v . For brevity, we shall<br />

now denote the norm <strong>of</strong> v by υ.<br />

Then (53.6) can be rewritten as<br />

v = υs (53.36)<br />

Taking the gradient <strong>of</strong> this equation yields<br />

grad v = s⊗grad υ+ υ grad s<br />

dυ dυ dυ<br />

= s⊗ s+ s⊗ n+ s⊗b<br />

ds dn db<br />

+ υκn⊗ s+ υθn⊗n−υ Ω + τ n⊗b<br />

( )<br />

( )<br />

+ υ Ω + τ b⊗n−υηb⊗b<br />

b<br />

n<br />

(53.37)<br />

4 A. W. Marris <strong>and</strong> C.-C. Wang, see footnote 3.


Sec. 53 • Three-Dimensional Euclidean Manifold, I 397<br />

where we have used (53.27) 1 . Notice that the component <strong>of</strong> grad v in<br />

identically, i.e.,<br />

(( ) ) [ ]( )<br />

b⊗<br />

s vanishes<br />

b⋅ grad v s = grad v b, s = 0<br />

(53.38)<br />

or, equivalently,<br />

dv<br />

b ⋅ = 0<br />

(53.39)<br />

ds<br />

so that the covariant derivative <strong>of</strong> v along its vec<strong>to</strong>r lines stays on the plane spanned by s <strong>and</strong> n .<br />

In differential geometry this plane is called the osculating plane <strong>of</strong> the said vec<strong>to</strong>r line.<br />

Next we can read <strong>of</strong>f from (53.37) the representations 5<br />

dυ div v = + υθ − υη<br />

ds<br />

dυ = + υ div s<br />

ds<br />

dυ ⎛ dυ<br />

⎞<br />

curl v = υ( Ω<br />

b<br />

+Ω<br />

n<br />

+ 2τ)<br />

s+ n+ ⎜υκ<br />

− ⎟b<br />

db ⎝ dn ⎠<br />

dυ<br />

⎛ dυ<br />

⎞<br />

= υΩ ss+ n+ ⎜υκ<br />

− ⎟b<br />

db ⎝ dn ⎠<br />

(53.40)<br />

where we have used (53.23) <strong>and</strong> (53.25). From (53.40) 4 the scalar field<br />

Ω =<br />

v⋅curl<br />

vis given by<br />

2<br />

Ω = υ ⋅ curl = υ Ω s<br />

s v (53.41)<br />

The representation (53.37) <strong>and</strong> its consequences (53.40) <strong>and</strong> (53.41) have many<br />

important applications in hydrodynamics <strong>and</strong> continuum mechanics. It should be noted that<br />

(53.21) 1 <strong>and</strong> (53.27) 1 now can be regarded as special cases <strong>of</strong> (53.40) 4 <strong>and</strong> (53.37) 2, respectively,<br />

with υ = 1.<br />

Exercises<br />

53.1 Prove the intrinsic equations (53.35).<br />

53.2 Show that the Serret-Frenet formulas can be written<br />

5 O. Bjørgum, see footnote 2. See also J.L. Ericksen, “Tensor Fields,” H<strong>and</strong>buch der Physik, <strong>Vol</strong>. III/1, Appendix,<br />

Edited by Flügge, Springer-Verlag (1960).


398 Chap. 10 • VECTOR FIELDS<br />

ds / ds = ω × s, dn/ ds = ω × n, db/<br />

ds = ω × b (53.42)<br />

where ω = τs+<br />

κb.


Sec. 54 • Three-Dimensional Euclidean Manifold, II 399<br />

Section 54. Vec<strong>to</strong>r Fields in a Three-Dimensional Euclidean Manifold, II.<br />

Representations for Special Classes <strong>of</strong> Vec<strong>to</strong>r Fields<br />

In Section 52 we have proved the Poincaré lemma, which asserts that, locally, a<br />

differential form is exact if <strong>and</strong> only if it is closed. This result means that we have the local<br />

representation<br />

f = dg (54.1)<br />

for any f such that<br />

d f = 0<br />

(54.2)<br />

In a three-dimensional Euclidean manifold the local representation (54.1) has the following two<br />

special cases.<br />

(i) Lamellar Fields<br />

v = grad f<br />

(54.3)<br />

is a local representation for any vec<strong>to</strong>r field v such that<br />

curl v = 0<br />

(54.4)<br />

Such a vec<strong>to</strong>r field v is called a lamellar field in the classical theory, <strong>and</strong> the scalar function f is<br />

called the potential <strong>of</strong> v . Clearly, the potential is locally unique <strong>to</strong> within an arbitrary additive<br />

constant.<br />

(ii) Solenoidal Fields<br />

v = curl u (54.5)<br />

is a local representation for any vec<strong>to</strong>r field v such that<br />

div v = 0<br />

(54.6)<br />

Such a vec<strong>to</strong>r field v is called a solenoidal field in the classical theory, <strong>and</strong> the vec<strong>to</strong>r field u is<br />

called the vec<strong>to</strong>r potential <strong>of</strong> v . The vec<strong>to</strong>r potential is locally unique <strong>to</strong> within an arbitrary<br />

additive lamellar field.<br />

In the representation (54.3), f is regarded as a 0-form <strong>and</strong> v is regarded as a 1-form,<br />

while in the representation (54.5), u is regarded as a 1-form <strong>and</strong> v is regarded as the dual <strong>of</strong> a 2-<br />

form; duality being defined by the canonical positive unit volume tensor <strong>of</strong> the 3-dimensional<br />

Euclidean manifold E .


400 Chap. 10 • VECTOR FIELDS<br />

In Section 52 we remarked also that the dual form <strong>of</strong> the Frobenius theorem implies the<br />

following representation.<br />

(iii) Complex-Lamellar Fields<br />

v = h grad f<br />

(54.7)<br />

is a local representation for any vec<strong>to</strong>r field v such that<br />

v⋅ curl v = 0<br />

(54.8)<br />

Such a vec<strong>to</strong>r field v is called complex-lamellar in the classical theory. In the representation<br />

(54.7) the surfaces defined by<br />

f ( x ) = const<br />

(54.9)<br />

are orthogonal <strong>to</strong> the vec<strong>to</strong>r field v .<br />

We shall now derive some other well-known representations in the classical theory.<br />

A. Euler’s Representation for Solenoidal Fields<br />

Every solenoidal vec<strong>to</strong>r field v may be represented locally by<br />

( grad h) ( grad f )<br />

v = ×<br />

(54.10)<br />

Pro<strong>of</strong>. We claim that v has a particular vec<strong>to</strong>r potential û which is complex-lamellar. From the<br />

remark on (54.5), we may choose û by<br />

uˆ<br />

= u + grad k<br />

(54.11)<br />

where u is any vec<strong>to</strong>r potential <strong>of</strong> v . In order for û <strong>to</strong> be complex-lamellar, it must satisfy the<br />

condition (54.8), i.e.,<br />

( u k) ( u k)<br />

( u grad k) curlu v ( u grad k)<br />

0 = + grad ⋅ curl + grad<br />

= + ⋅ = ⋅ +<br />

(54.12)<br />

Clearly, this equation possesses infinitely many solutions for the scalar function k, since it is a<br />

first-order partial differential equation with smooth coefficients. Hence by the representation<br />

(54.7) we may write


Sec. 54 • Three-Dimensional Euclidean Manifold, II 401<br />

u ˆ = h grad f<br />

(54.13)<br />

Taking the curl <strong>of</strong> this equation, we obtain the Euler representation (54.10):<br />

( h f ) ( h) ( f )<br />

v = curlu ˆ = curl grad = grad × grad<br />

(54.14)<br />

It should be noted that in the Euler representation (54.10) the vec<strong>to</strong>r v is orthogonal <strong>to</strong><br />

grad h as well as <strong>to</strong> grad f , namely<br />

v⋅ grad h = v ⋅ grad f = 0<br />

(54.15)<br />

Consequently, v is tangent <strong>to</strong> the surfaces<br />

h ( x ) = const<br />

(54.16)<br />

or<br />

f ( x ) = const<br />

(54.17)<br />

For this reason, these surfaces are then called vec<strong>to</strong>r sheets <strong>of</strong> v . If v ≠ 0 , then from (54.10),<br />

grad h <strong>and</strong> grad f are not parallel, so that h <strong>and</strong> f are functionally independent. In this case the<br />

intersections <strong>of</strong> the surfaces (54.16) <strong>and</strong> (54.17) are the vec<strong>to</strong>r lines <strong>of</strong> v .<br />

Euler’s representation for solenoidal fields implies the following results.<br />

B. Monge’s Representation for Arbitrary Smooth Vec<strong>to</strong>r Fields<br />

Every smooth vec<strong>to</strong>r field v may be represented locally by<br />

v = grad h+<br />

kgrad<br />

f<br />

(54.18)<br />

where the scalar functions, h, k, f are called the Monge potentials (not unique) <strong>of</strong> v .<br />

Pro<strong>of</strong>. Since (54.10) is a representation for any solenoidal vec<strong>to</strong>r field, from (54.5) we can write<br />

curl v as<br />

( k) ( f )<br />

curl v = grad × grad<br />

(54.19)<br />

It follows that (54.19) that<br />

( k f )<br />

curl v − grad = 0<br />

(54.20)


402 Chap. 10 • VECTOR FIELDS<br />

Thus v − k grad f is a lamellar vec<strong>to</strong>r field. From (54.3) we then we have the local<br />

representation<br />

v − k grad f = grad h<br />

(54.21)<br />

which is equivalent <strong>to</strong> (54.18).<br />

Next we prove another well-known representation for arbitrary smooth vec<strong>to</strong>r fields in<br />

the classical theory.<br />

C. S<strong>to</strong>kes’ Representation for Arbitrary Smooth Vec<strong>to</strong>r Fields<br />

Every smooth vec<strong>to</strong>r field v may be represented locally by<br />

v = grad h + curlu (54.22)<br />

where<br />

h <strong>and</strong> u are called the S<strong>to</strong>kes potential (not unique) <strong>of</strong> v .<br />

Pro<strong>of</strong>. We show that there exists a scalar function h such that<br />

Equivalently, this condition means<br />

( h)<br />

v − grad h is solenoidal.<br />

div v − grad = 0<br />

(54.23)<br />

Exp<strong>and</strong>ing (54.23), we get<br />

Δ h = div v (54.24)<br />

where Δ denotes the Laplacian [cf. (47.49)]. Thus h satisfies the Poisson equation (54.24). It is<br />

well known that, locally, there exist infinitely many solutions (54.24). Hence the representation<br />

(54.22) is valid.<br />

Notice that, from (54.19), S<strong>to</strong>kes’ representation (54.22) also can be put in the form<br />

( ) ( )<br />

v = grad h+ grad k × grad f<br />

(54.25)<br />

Next we consider the intrinsic conditions for the various special classes <strong>of</strong> vec<strong>to</strong>r fields.<br />

First, from (53.40) 2 a vec<strong>to</strong>r field v is solenoidal if <strong>and</strong> only if<br />

dυ<br />

/ ds = −υ<br />

div s (54.26)


Sec. 54 • Three-Dimensional Euclidean Manifold, II 403<br />

where s , υ, <strong>and</strong> s are defined in the preceding section. Integrating (54.26) along any vec<strong>to</strong>r line<br />

( s)<br />

λ = λ defined before, we obtain<br />

υ<br />

s<br />

( s) υ0exp( - div ds<br />

s<br />

)<br />

= ∫ s (54.27)<br />

0<br />

where υ<br />

0<br />

is the value <strong>of</strong> υ at any reference point λ ( s 0 ) . Thus1 in a solenoidal vec<strong>to</strong>r field v<br />

the vec<strong>to</strong>r magnitude is determined <strong>to</strong> within a constant fac<strong>to</strong>r along any vec<strong>to</strong>r line = ( s)<br />

the vec<strong>to</strong>r line pattern <strong>of</strong> v .<br />

From (53.40) 4 , a vec<strong>to</strong>r field v is complex-lamellar if <strong>and</strong> only if<br />

λ λ by<br />

or, equivalently,<br />

υΩ = 0<br />

(54.28)<br />

s<br />

Ω = 0<br />

(54.29)<br />

s<br />

since in the intrinsic representation v is assumed <strong>to</strong> be nonvanishing. The result (54.29) is<br />

entirely obvious, because it defines the unit vec<strong>to</strong>r field s , parallel <strong>to</strong> v , <strong>to</strong> be complex-lamellar.<br />

From (53.40) 4 again, a vec<strong>to</strong>r field v is lamellar if <strong>and</strong> only if, in addition <strong>to</strong> (54.28) or<br />

(54.29), we have also<br />

dυ / db= 0, dυ/<br />

dn = υκ<br />

(54.30)<br />

It should be noted that, when v is lamellar, it can be represented by (54.3), <strong>and</strong> thus the potential<br />

surfaces defined by<br />

f ( x ) = const<br />

(54.31)<br />

are formed by the integral curves <strong>of</strong> n <strong>and</strong> b . From (54.30) 1 along any b − line<br />

( ) υ( )<br />

υ b = b 0<br />

= const<br />

(54.32)<br />

while from (54.30) 2 along any<br />

n − line<br />

n<br />

( n) ( n0 ) exp( dn<br />

n<br />

)<br />

υ = υ ∫ κ<br />

(54.33)<br />

0<br />

1 O. Bj∅rgum, see footnote 2 in Section 53.


404 Chap. 10 • VECTOR FIELDS<br />

Finally, in the classical theory a vec<strong>to</strong>r field v is called a screw field or a Beltrami field if<br />

v is parallel <strong>to</strong> its curl, namely<br />

v× curl v = 0 (54.34)<br />

or, equivalently,<br />

curl v = Ωsv (54.35)<br />

where Ω<br />

s<br />

is the abnormality <strong>of</strong> s , defined by (53.22) 1 . In some sense a screw field is just the<br />

opposite <strong>of</strong> a complex-lamellar field, which is defined by the condition that the vec<strong>to</strong>r field is<br />

orthogonal <strong>to</strong> its curl [cf. (54.8)]. Unfortunately, there is no known simple direct representation<br />

for screw fields. We must refer the reader <strong>to</strong> the three long articles by Bjørgum <strong>and</strong> Godal (see<br />

footnote 1 above <strong>and</strong> footnotes 4 <strong>and</strong> 5 below), which are devoted entirely <strong>to</strong> the study <strong>of</strong> these<br />

fields.<br />

We can, <strong>of</strong> course, use some general representations for arbitrary smooth vec<strong>to</strong>r fields,<br />

such as Monge’s representation or S<strong>to</strong>kes’ representation, <strong>to</strong> express a screw field first. Then we<br />

impose some additional restrictions on the scalar fields involved in the said representations. For<br />

example, if we use Monge’s representation (54.18) for v , then curl v is given by (54.19). In<br />

this case v is a screw field if <strong>and</strong> only if the constant potential surfaces <strong>of</strong> k <strong>and</strong> f are vec<strong>to</strong>r<br />

sheets <strong>of</strong> v , i.e.,<br />

v⋅ grad k = v ⋅ grad f = 0<br />

(54.36)<br />

From (53.40) 4 the intrinsic conditions for a screw field are easily fround <strong>to</strong> be simply the<br />

conditions (54.30). So the integrals (54.32) <strong>and</strong> (54.33) remain valid in this case, along the<br />

b − lines <strong>and</strong> the n − lines. When the abnormality Ω<br />

s<br />

is nonvanishing, the integral <strong>of</strong> υ along<br />

any s − line can be found in the following way: From (53.40) we have<br />

dυ<br />

/ ds = divv−υ<br />

divs (54.37)<br />

Now from the basic conditions (54.35) for a screw field we obtain<br />

d<br />

s<br />

0= div( Ω<br />

sv) =Ω<br />

sdivv +υ Ω<br />

(54.38)<br />

ds<br />

So div v can be represented by<br />

div<br />

υ dΩ<br />

=− Ω ds<br />

s<br />

v (54.39)<br />

s


Sec. 54 • Three-Dimensional Euclidean Manifold, II 405<br />

Substituting (54.39) in<strong>to</strong> (54.37) yields<br />

dυ<br />

1<br />

=− υ ⎜<br />

⎛ div +<br />

ds ⎝ Ω<br />

s s<br />

(54.40)<br />

s<br />

d Ω<br />

⎟<br />

⎞<br />

ds ⎠<br />

or, equivalently,<br />

d<br />

( υ )<br />

The last equation can be integrated at once, <strong>and</strong> the result is<br />

υ<br />

( s)<br />

Ω<br />

s =− υ Ω<br />

s<br />

div s (54.41)<br />

ds<br />

( s0) Ωs<br />

( s0)<br />

( s )<br />

υ<br />

= −<br />

s<br />

( ( div ) ds<br />

s<br />

)<br />

exp<br />

Ω<br />

∫ s (54.42)<br />

0<br />

s 0<br />

which may be rewritten as<br />

υ<br />

υ<br />

( s)<br />

( s )<br />

0<br />

( s0<br />

)<br />

( s)<br />

Ω<br />

= −<br />

s<br />

( ( div ) ds<br />

s<br />

)<br />

s<br />

exp<br />

Ω<br />

∫ s (54.43)<br />

0<br />

s<br />

since υ is nonvanishing. From (54.43), (54.33), <strong>and</strong> (54.32) we see that the magnitude <strong>of</strong> a<br />

screw field, except for a constant fac<strong>to</strong>r, is determined along any s− line, n−line, or b −line<br />

by<br />

the vec<strong>to</strong>r line pattern <strong>of</strong> the field. 2<br />

A screw field whose curl is also a screw field is called a Trkalian field. According <strong>to</strong> a<br />

theorem <strong>of</strong> Mémenyi <strong>and</strong> Prim (1949), a screw field is a Trkalian field if <strong>and</strong> only if its<br />

abnormality is a constant. Further, all Trkalian fields are solenoidal <strong>and</strong> successive curls <strong>of</strong> the<br />

field are screw fields, all having the same abnormality. 3<br />

The pro<strong>of</strong> <strong>of</strong> this theorem may be found also in Bjørgum’s article. Trkalian fields are<br />

considered in detail in the subsequent articles <strong>of</strong> Bjørgum <strong>and</strong> Godal 4 <strong>and</strong> Godal 5<br />

2 O. Bjørgum, see footnote 2, Section 53.<br />

3 P. Nemenyi <strong>and</strong> R. Prim, “Some Properties <strong>of</strong> Rotational Flow <strong>of</strong> a Perfect Gas,” Proc. Nat. Acad. Sci. 34, 119-<br />

124; Erratum 35, 116 (1949).<br />

4 O. Bjørgum <strong>and</strong> T. Godal, “On Beltrami Vec<strong>to</strong>r Fields <strong>and</strong> Flows, Part II. The Case when Ω is Constant in Space,”<br />

Universitetet i Bergen, Arbok 1952, Naturvitenskapelig rekke Nr. 13.<br />

5 T. Godal, “On Beltrami Vec<strong>to</strong>r Fields <strong>and</strong> Flows, Part III. Some Considerations on the General Case,” Universitete<br />

i Bergen, Arbok 1957, Naturvitenskapelig rekke Nr.12.


________________________________________________________________<br />

Chapter 11<br />

HYPERSURFACES ON A EUCLIDEAN MANIFOLD<br />

In this chapter we consider the theory <strong>of</strong> (N-1)-dimensional hypersurfaces embedded in an N-<br />

dimensional Euclidean manifold E . We shall not treat hypersurfaces <strong>of</strong> dimension less than N-<br />

1, although many results <strong>of</strong> this chapter can be generalized <strong>to</strong> results valid for those<br />

hypersurfaces also.<br />

Section 55. Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface metric<br />

A hypersurface <strong>of</strong> dimension N-1 in E a set S <strong>of</strong> points in E which can be<br />

characterized locally by an equation<br />

x∈N ⊂S ⇔ f ( x) = 0<br />

(55.1)<br />

where f is a smooth function having nonvanishing gradient. The unit vec<strong>to</strong>r field on N<br />

grad f<br />

n =<br />

(55.2)<br />

grad f<br />

is called a unit normal <strong>of</strong> S , since from (55.1) <strong>and</strong> (48.15) for any smooth curve λ = λ( t)<br />

in S<br />

we have<br />

1<br />

0 = df λ<br />

(grad f )<br />

dt<br />

= ⋅ λ = n<br />

grad f<br />

⋅ λ (55.3)<br />

The local representation (55.1) <strong>of</strong> S is not unique, or course. Indeed, if f satisfies (55.1),<br />

so does –f, the induced unit normal <strong>of</strong> –f being –n. If the hypersurface S can be represented<br />

globally by (55.1), i.e., there exists a smooth function whose domain contains the entire<br />

hypersurface such that<br />

x∈S ⇔ f ( x) = 0<br />

(55.4)<br />

then S is called orientable. In this case S can be equipped with a smooth global unit normal<br />

field n. (Of course, -n is also a smooth global unit normal field.) We say that S is oriented if a<br />

particular smooth global unit normal field n has been selected <strong>and</strong> designated as the positive unit<br />

normal <strong>of</strong> S . We shall consider oriented hypersurfaces only in this chapter.<br />

407


408 Chap. 11 • HYPERSURFACES<br />

Since grad f is nonvanishing, S can be characterized locally also by<br />

1 1<br />

x = x ( y ,…, y N − )<br />

(55.5)<br />

1 N −1<br />

in such a way that ( y ,…, y , f)<br />

forms a local coordinate system in S . If n is the positive unit<br />

1 1<br />

normal <strong>and</strong> f satisfies (55.2), then the parameters ( y ,…, y N − ) are said <strong>to</strong> form a positive local<br />

1 N −1<br />

coordinate system in S when ( y ,…, y , f)<br />

is a positive local coordinate system in E . Since<br />

the coordinate curves <strong>of</strong> y<br />

Γ , Γ= 1, …, N −1,<br />

are contained in S the natural basis vec<strong>to</strong>rs<br />

h = ∂x / ∂ Γ<br />

Γ<br />

y , Γ= 1, … , N − 1<br />

(55.6)<br />

are tangent <strong>to</strong> S . Moreover, the basis { h , Γ<br />

n}<br />

is positive in E . We call the ( 1)<br />

S spanned by { }<br />

dimensional hyperplane<br />

x<br />

Since<br />

The reciprocal basis <strong>of</strong> { } Γ ,<br />

N − -<br />

h ( ) Γ<br />

x the tangent plane <strong>of</strong> S at the point x ∈S.<br />

h ⋅ n = Γ<br />

0, Γ= 1, … , N − 1<br />

(55.7)<br />

h n has the form { h<br />

Γ , }<br />

n where<br />

Γ<br />

h are also tangent <strong>to</strong> S , namely<br />

h Γ ⋅ n = 0, Γ= 1, …, N −1<br />

(55.8)<br />

<strong>and</strong><br />

In view <strong>of</strong> (55.9) we call { } Γ<br />

Γ<br />

Γ<br />

h ⋅ h = δ , Γ, Δ = 1, …, N −1<br />

(55.9)<br />

Δ<br />

Δ<br />

h <strong>and</strong> { h Γ<br />

} reciprocal natural bases <strong>of</strong> ( )<br />

y Γ on S .<br />

Let v be a vec<strong>to</strong>r field on S , i. e., a function v; S →V . Then, for each x ∈S , v can<br />

h<br />

Γ , n by<br />

be represented in terms <strong>of</strong> the bases { } Γ ,<br />

h n <strong>and</strong> { }<br />

Γ N<br />

Γ<br />

υ<br />

Γ<br />

υ υΓ<br />

υ N<br />

v = h + n= h + n (55.10)<br />

where, from (55.7)-(55.9),<br />

Γ Γ<br />

N<br />

υ = v⋅ h , υ = v⋅ h , υ = υ = v⋅n (55.11)<br />

Γ<br />

Γ<br />

N


Sec. 55 • Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface Metric 409<br />

We call the vec<strong>to</strong>r field<br />

Γ<br />

Γ<br />

vS ≡ υ hΓ<br />

= υΓh = v−( v⋅n)<br />

n<br />

(55.12)<br />

the tangential projection <strong>of</strong> v , <strong>and</strong> we call the vec<strong>to</strong>r field<br />

v υ n υ n v v S<br />

(55.13)<br />

N<br />

n<br />

≡ =<br />

N<br />

= −<br />

the normal projection <strong>of</strong> v . Notice that in (55.10)-(55.12) the repeated Greek index is summed<br />

from 1 <strong>to</strong> N-1. We say that v is a tangential vec<strong>to</strong>r field on S if v = n<br />

0 <strong>and</strong> a normal vec<strong>to</strong>r<br />

field if vS = 0.<br />

be written<br />

i<br />

If we introduce a local coordinate system ( x ) in E , then the representation (55.5) may<br />

i i 1 N−1<br />

x = x ( y , , y ), i = 1, , N<br />

… … (55.14)<br />

From (55.14) the surface basis { h Γ } is related <strong>to</strong> the natural basis { i }<br />

x by<br />

i<br />

g <strong>of</strong> ( )<br />

i<br />

i ∂x<br />

hΓ = hΓgi<br />

= g<br />

i, Γ = 1, …, N −1<br />

(55.15)<br />

Γ<br />

∂y<br />

while from (55.2) the unit normal n has the component form<br />

i<br />

∂f<br />

/ ∂x<br />

i<br />

n =<br />

g (55.16)<br />

ab a b 1/2<br />

( g ( ∂f / ∂x )( ∂f / ∂x<br />

))<br />

where, as usual{ i<br />

}<br />

metric, namely<br />

g , is the reciprocal basis <strong>of</strong> { }<br />

ab<br />

g i<br />

<strong>and</strong> { g }<br />

is the component <strong>of</strong> the Euclidean<br />

ab a b<br />

g = g ⋅g (55.17)<br />

The component representation for the surface reciprocal basis { h Γ<br />

} is somewhat harder<br />

<strong>to</strong> find. We find first the components <strong>of</strong> the surface metric.<br />

a<br />

i i<br />

∂x<br />

∂x<br />

≡ h ⋅ h = g<br />

∂y<br />

∂y<br />

(55.18)<br />

ΓΔ Γ Δ ij Γ Δ


410 Chap. 11 • HYPERSURFACES<br />

where g<br />

ij<br />

is a component <strong>of</strong> the Euclidean metric<br />

g = g ⋅g (55.19)<br />

ij i j<br />

Now let ⎡<br />

⎣a ΓΔ ⎤<br />

⎦<br />

be the inverse <strong>of</strong> [ ]<br />

a ΓΔ ,i.e., , , 1, , 1<br />

ΓΔ<br />

Γ<br />

a a = δ ΓΣ= … N −<br />

(55.20)<br />

ΔΣ<br />

Σ<br />

The inverse matrix exists because from (55.18), [ a ΓΔ ] is positive-definite <strong>and</strong> symmetric. In<br />

fact, from (55.18) <strong>and</strong> (55.9)<br />

a ΓΔ = h Γ ⋅ h Δ<br />

(55.21)<br />

so that a ΓΔ is also a component <strong>of</strong> the surface metric. From (55.21) <strong>and</strong> (55.15) we then have<br />

i<br />

Γ ΓΔ ∂x<br />

h = a g<br />

i, Γ= 1, …, N −1<br />

(55.22)<br />

Δ<br />

∂y<br />

Γ<br />

which is the desired representation for h .<br />

At each point x ∈S the components a ΓΔ<br />

( x ) <strong>and</strong> a ΓΔ ( x ) defined by (55.18) <strong>and</strong> (55.21)<br />

are those <strong>of</strong> an inner product on S<br />

x<br />

relative <strong>to</strong> the surface coordinate system ( y Γ ) , the inner<br />

product being the one with induced by that <strong>of</strong> V since S<br />

x<br />

is a subspace <strong>of</strong> V . In other words<br />

if u <strong>and</strong> v are tangent <strong>to</strong> S at x , say<br />

Γ<br />

Δ<br />

u= u h ( x) v = υ h ( x )<br />

(55.23)<br />

Γ<br />

Δ<br />

then<br />

u⋅ v = a ( x ) u Γ υ<br />

Δ<br />

(55.24)<br />

ΓΔ<br />

This inner product gives rise <strong>to</strong> the usual operations <strong>of</strong> rising <strong>and</strong> lowering <strong>of</strong> indices for tangent<br />

vec<strong>to</strong>rs <strong>of</strong> S . Thus (55.23) 1<br />

is equivalent <strong>to</strong><br />

Γ Δ Δ<br />

u= a ( x) u h ( x) = u h ( x )<br />

(55.25)<br />

ΓΔ<br />

Δ


Sec. 55 • Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface Metric 411<br />

Obviously we can also extend the operations <strong>to</strong> tensor fields on S having nonzero components<br />

in the product basis <strong>of</strong> the surface bases { h Γ } <strong>and</strong> { h Γ<br />

} only. Such a tensor field A may be<br />

called a tangential tensor field <strong>of</strong> S <strong>and</strong> has the representation<br />

Then<br />

A<br />

A<br />

h<br />

<br />

Γ1...<br />

Γr<br />

=<br />

Γ<br />

⊗ ⊗<br />

1 Γ2<br />

= A h ⊗h ⊗⊗h<br />

h<br />

Γ2...<br />

Γr<br />

Γ1<br />

Γ1 Γ2<br />

Γr<br />

, etc.<br />

(55.26)<br />

A = a A , etc.<br />

(55.27)<br />

Γ2... Γr<br />

ΔΓ2...<br />

Γr<br />

Γ1 Γ1Δ<br />

There is a fundamental difference between the surface metric a on S <strong>and</strong> the Euclidean<br />

metric g on E however. In the Euclidean space E there exist coordinate systems in which the<br />

components <strong>of</strong> g are constant. Indeed, if the coordinate system is a rectangular Cartesian one,<br />

then gij<br />

is δij<br />

at all points <strong>of</strong> the domain <strong>of</strong> the coordinate system. On the hypersurface S ,<br />

generally, there need not be any coordinate system in which the components a ΓΔ<br />

or a ΓΔ are<br />

constant unless S happens <strong>to</strong> be a hyperplane. As we shall see in a later section, the departure<br />

<strong>of</strong> a from g in this regard can be characterized by the curvature <strong>of</strong> S .<br />

Another important difference between S <strong>and</strong> E is the fact that in S the tangent planes at<br />

different points generally are different (N-1)-dimensional subspaces <strong>of</strong> V . Hence a vec<strong>to</strong>r in V<br />

may be tangent <strong>to</strong> S at one point but not at another point. For this reason there is no canonical<br />

parallelism which connects the tangent planes <strong>of</strong> S at distinct points. As a result, the notions <strong>of</strong><br />

gradient or covariant derivative <strong>of</strong> a tangential vec<strong>to</strong>r or tensor field on S must be carefully<br />

defined, as we shall do in the next section.<br />

The notions <strong>of</strong> Lie derivative <strong>and</strong> exterior derivative introduced in Sections 49 <strong>and</strong> 51,<br />

however, can be readily defined for tangential fields <strong>of</strong> S . We consider first the Lie derivative.<br />

Let v be a smooth tangent field defined on a domain in S . Then as before we say that a<br />

smooth curve λ = λ ( t)<br />

in the domain <strong>of</strong> v is an integral curve if<br />

dλ()/ t dt = v( λ ()) t<br />

(55.28)<br />

at all points <strong>of</strong> the curve. If we represent λ <strong>and</strong> v in component forms<br />

1 N −1<br />

ˆ( ( )) ( ( ), , ( )) <strong>and</strong><br />

y λ t = λ t … λ t v = υ Γ h<br />

(55.29)<br />

Γ<br />

relative <strong>to</strong> a surface coordinate system ( y Γ ) , then (55.28) can be expressed by<br />

Γ<br />

Γ<br />

dλ<br />

()/ t dt = υ ( λ ()) t<br />

(55.30)


412 Chap. 11 • HYPERSURFACES<br />

Hence integral curves exist for any smooth tangential vec<strong>to</strong>r field v <strong>and</strong> they generate flow, <strong>and</strong><br />

hence a parallelism, along any integral curve. By the same argument as in Section 49, we define<br />

the Lie derivative <strong>of</strong> a smooth tangential vec<strong>to</strong>r field u relative <strong>to</strong> v by the limit (49.14), except<br />

that now ρt<br />

<strong>and</strong> Pt<br />

are the flow <strong>and</strong> the parallelism in S . Following exactly the same derivation<br />

as before, we then obtain<br />

Γ<br />

Γ<br />

⎛∂u<br />

Δ ∂υ<br />

Δ⎞<br />

L u= ⎜ υ − u<br />

Δ<br />

Δ ⎟hΓ<br />

(55.31)<br />

v<br />

⎝∂y<br />

∂y<br />

⎠<br />

which generalizes (49.21). Similarly if A is a smooth tangential tensor field on S , then L A is<br />

v<br />

defined by (49.41) with ρt<br />

<strong>and</strong> Pt<br />

as just explained, <strong>and</strong> the component representation for L A is<br />

v<br />

( S A)<br />

v<br />

∂A<br />

Γ1...<br />

Γr<br />

Γ1<br />

Γ1...<br />

Γr<br />

Δ1... Δs<br />

Σ ΣΓ2.<br />

... Γ ∂υ<br />

r<br />

Δ1... Δ<br />

= υ − A<br />

s<br />

Σ<br />

Δ1...<br />

Δs<br />

Σ<br />

∂y<br />

∂υ<br />

−− A<br />

+<br />

∂y<br />

+ +<br />

A<br />

Γr<br />

Σ<br />

Γ1... Γr−1Σ Γ1...<br />

Γr<br />

Δ<br />

A<br />

1…<br />

Δs<br />

Σ<br />

ΣΔ2…<br />

Δs<br />

Δ1<br />

∂υ<br />

∂y<br />

Σ<br />

Γ1...<br />

Γr<br />

Δ1…<br />

Δs−1Σ Δs<br />

∂y<br />

∂υ<br />

∂y<br />

(55.32)<br />

which generalizes (49.42).<br />

Next we consider the exterior derivative. Let A be a tangential differential form on S ,<br />

i.e., A is skew-symmetric <strong>and</strong> has the representations<br />

A<br />

A<br />

Γ1<br />

=<br />

Γ1...<br />

Γ<br />

⊗ ⊗<br />

r<br />

Γ1<br />

= Σ<br />

Γ1...<br />

Γ<br />

∧ ∧<br />

Γ r<br />

1


Sec. 55 • Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface Metric 413<br />

N −1<br />

A<br />

dA<br />

= Σ Σ h ∧h ∧∧h<br />

Γ 1


414 Chap. 11 • HYPERSURFACES<br />

i j<br />

ij x x i j<br />

g ( x ΓΔ ∂ ∂<br />

) = a ( x ) + n ( ) n ( )<br />

Δ Γ<br />

∂y<br />

∂y<br />

x x<br />

55.4 Show that<br />

l<br />

∂x<br />

Σ<br />

Lxg j( x) = g<br />

jl( x) h ( x )<br />

Σ<br />

∂y<br />

<strong>and</strong><br />

j<br />

∂x<br />

j<br />

Σ<br />

Lxg ( x) = h ( x ) ∂<br />

Σ<br />

y<br />

55.5 Compute the components a ΓΔ<br />

for (a) the spherical surface defined by<br />

x = csin<br />

y cos y<br />

1 1 2<br />

x = csin<br />

y sin y<br />

2 1 2<br />

x = ccos<br />

y<br />

3 1<br />

<strong>and</strong> (b) the cylindrical surface<br />

1 1 2 1 3 2<br />

x = c cos y , x = c sin y , x = y<br />

where c is a constant.<br />

55.6 For N=3 show that<br />

a = a / a, a = − a / a, a = a / a<br />

11 12 22<br />

22 12 11<br />

a = a a − a = a ΔΓ<br />

.<br />

2<br />

where [ ]<br />

11 22 12<br />

det<br />

55.7 Given an ellipsoid <strong>of</strong> revolution whose surface is determined by<br />

x = bcos<br />

y sin y<br />

1 1 2<br />

x = bsin<br />

y sin y<br />

2 1 2<br />

x = ccos<br />

y<br />

3 1<br />

where b <strong>and</strong> c are constants <strong>and</strong><br />

b<br />

> c show that<br />

2 2


Sec. 55 • Normal Vec<strong>to</strong>r, Tangent Plane, <strong>and</strong> Surface Metric 415<br />

a = b sin y , a = 0<br />

2 2 2<br />

11 12<br />

<strong>and</strong><br />

a22 = b cos y + c sin y<br />

2 2 2 2 2 2


416 Chap. 11 • HYPERSURFACES<br />

Section 56. Surface Covariant Derivatives<br />

As mentioned in the preceding section, the tangent planes at different points <strong>of</strong> a<br />

hypersurface generally do not coincide as subspaces <strong>of</strong> the translation space V <strong>of</strong> E. Consequently,<br />

it is no longer possible <strong>to</strong> define the gradient or covariant derivative <strong>of</strong> a tangential tensor field<br />

by a condition that formally generalizes (47.17). However, we shall see that it is possible <strong>to</strong><br />

define a type <strong>of</strong> differentiation which makes (47.17) formally unchanged. For a tangential tensor<br />

field A represented by (55.26) we define the surface gradient ∇A by<br />

∇ A = ⊗ ⊗ ⊗<br />

(56.1)<br />

1<br />

A Γ … Γr<br />

,<br />

Δ<br />

ΔhΓ h<br />

1<br />

Γ<br />

h<br />

r<br />

where<br />

Γ1…<br />

Γr<br />

A<br />

Γ<br />

1 r 2 r 1<br />

Γ<br />

Γ … Γ ∂<br />

ΣΓ … Γ ⎧ ⎫ Γ1…<br />

Γr−1Σ⎧ r ⎫<br />

A ,<br />

Δ<br />

= + A + + A<br />

Δ ⎨ ⎬ ⎨ ⎬<br />

(56.2)<br />

∂y<br />

⎩ΣΔ⎭ ⎩ΣΔ⎭<br />

<strong>and</strong><br />

⎧ Ω ⎫ 1 ΩΣ ⎛∂aΓΣ ∂aΔΣ ∂aΓΔ<br />

⎞<br />

⎨ ⎬= a + −<br />

Δ Γ Σ<br />

ΓΔ 2<br />

⎜<br />

∂y ∂y ∂y<br />

⎟<br />

⎩ ⎭ ⎝ ⎠<br />

(56.3)<br />

The quantities<br />

⎧ Ω ⎫<br />

⎨ ⎬are the surface Chris<strong>to</strong>ffel symbols. They obey the symmetric condition<br />

⎩ΓΔ⎭ ⎧ Ω ⎫ ⎧ Ω ⎫<br />

⎨ ⎬=<br />

⎨ ⎬<br />

⎩ΓΔ⎭ ⎩ΔΓ⎭<br />

(56.4)<br />

We would like <strong>to</strong> have available a formula like (47.26) which would characterize the surface<br />

Chris<strong>to</strong>ffel symbols as components <strong>of</strong> ∂h<br />

y Γ<br />

Δ<br />

∂ . However, we have no assurance that<br />

∂h y Γ<br />

Δ<br />

∂ is a tangential vec<strong>to</strong>r field. In fact, as we shall see in Section 58, ∂h<br />

y Γ<br />

Δ<br />

∂ is not<br />

generally a tangential vec<strong>to</strong>r field. However, given the surface Chris<strong>to</strong>ffel symbols, we can<br />

formally write<br />

Dh<br />

Dy<br />

Δ<br />

Γ<br />

⎧ Ω ⎫<br />

= ⎨ ⎬h Ω<br />

(56.5)<br />

⎩ΓΔ⎭<br />

With this definition, we can combine (56.2) with (56.1) <strong>and</strong> obtain


Sec. 56 • Surface Covariant Derivatives 417<br />

Γ1…<br />

Γr<br />

∂A<br />

∇ A = h<br />

Δ Γ<br />

⊗⊗h 1<br />

Γ<br />

⊗h<br />

r<br />

∂y<br />

⎧Dh<br />

Dh<br />

⎫<br />

+ A ⊗⊗ h + + h ⊗⊗ ⊗h<br />

⎩ Dy<br />

Dy ⎭<br />

Γ1…<br />

Γr<br />

Γ1<br />

Γr<br />

Δ<br />

⎨ Δ<br />

Γr<br />

Γ1<br />

Δ ⎬<br />

Δ<br />

(56.6)<br />

If we adopt the notional convention<br />

DA<br />

Dy<br />

∂A<br />

= ∂ y<br />

Γ1…<br />

Γr<br />

Γ1…<br />

Γr<br />

Δ<br />

Δ<br />

(56.7)<br />

then (56.6) takes the suggestive form<br />

DA<br />

Δ<br />

∇ A = ⊗h<br />

(56.8)<br />

Δ<br />

Dy<br />

1 r<br />

The components A Γ … Γ ,<br />

Δ<br />

<strong>of</strong> ∇A represents the surface covariant derivative. If the mixed<br />

components <strong>of</strong> A are used, say with<br />

A = A ⊗ ⊗ ⊗ ⊗ ⊗<br />

Γ1…<br />

Γr<br />

Δ1<br />

Δs<br />

Δ1… Δ<br />

h<br />

s Γ<br />

h<br />

1<br />

Γ<br />

h h<br />

(56.9)<br />

r<br />

then the components <strong>of</strong> ∇A are given by<br />

A<br />

∂A<br />

Γ1…<br />

Γr<br />

Γ1…<br />

Γ<br />

Δ<br />

r<br />

1…<br />

Δs<br />

Δ<br />

,<br />

1…<br />

Δs<br />

Σ=<br />

Σ<br />

∂y<br />

Γ<br />

2 1<br />

Γ<br />

ΩΓ … Γ ⎧ ⎫<br />

r<br />

Γ1…<br />

Γr−1Ω<br />

⎧ r ⎫<br />

+ A<br />

Δ<br />

A<br />

1…<br />

Δ ⎨ ⎬+ +<br />

s<br />

Δ1<br />

Δ ⎨ ⎬<br />

s<br />

⎩ΩΣ⎭ ⎩ΩΔ⎭<br />

⎧ Ω ⎫<br />

⎧ Ω ⎫<br />

−A<br />

−−<br />

A<br />

Γ1…<br />

Γr<br />

Γ1…<br />

Γr<br />

ΩΔ2…<br />

Δ ⎨ ⎬<br />

s<br />

Δ1…<br />

Δs−1Ω<br />

⎨ ⎬<br />

⎩ΔΣ<br />

1 ⎭ ⎩ΔΣ<br />

s ⎭<br />

(56.10)<br />

which generalizes (47.39). Equation (56.10) can be formally derived from (56.8) if we adopt the<br />

definition<br />

Dh<br />

Dy<br />

Γ<br />

Σ<br />

⎧ Γ ⎫<br />

=−⎨ ⎬<br />

⎩ΣΔ⎭<br />

Δ<br />

h (56.11)<br />

When we apply (56.10) <strong>to</strong> the surface metric, we obtain


418 Chap. 11 • HYPERSURFACES<br />

a = a , = δ = 0<br />

(56.12)<br />

ΓΔ Γ<br />

ΓΔ, Σ Σ Δ,<br />

Σ<br />

which generalizes (47.40).<br />

The formulas (56.2) <strong>and</strong> (56.10) give the covariant derivative <strong>of</strong> a tangential tensor field<br />

1 … r<br />

1…<br />

r<br />

in component form. To show that , <strong>and</strong> , are the components <strong>of</strong> some<br />

A Γ Γ Δ<br />

A Γ Γ Δ1…<br />

Δs<br />

Σ<br />

tangential tensor fields, we must show that they obey the tensor transformation rule, e.g., if<br />

y Γ is another surface coordinate system, then<br />

( )<br />

A<br />

∂y ∂y ∂y<br />

, = ,<br />

Ω<br />

(56.13)<br />

∂ y ∂ y ∂ y<br />

Γ1<br />

Γr<br />

Ω<br />

Γ1…<br />

Γr<br />

Σ1…<br />

Σr<br />

Δ<br />

A<br />

Σ1<br />

Σr<br />

Δ<br />

1 … r<br />

where ,<br />

A Γ Γ Δ<br />

are obtained from (56.2) when all the fields on the right-h<strong>and</strong> side are referred <strong>to</strong><br />

( y Γ<br />

) . To prove (56.13), we observe first that from (56.3) <strong>and</strong> the fact that a ΓΔ<br />

<strong>and</strong> a ΓΔ are<br />

components <strong>of</strong> the surface metric, so that<br />

Γ Δ Σ Ω<br />

ΓΔ ∂y ∂y ΣΘ ∂y ∂y<br />

a = a , aΓΔ<br />

= a<br />

Σ Θ Γ Δ<br />

∂y ∂y ∂y ∂y<br />

ΣΩ<br />

(56.14)<br />

we have the transformation rule<br />

Σ 2 Ω Σ Θ Φ<br />

⎧ Σ ⎫ ∂y ∂ y ∂y ∂y ∂y<br />

⎧ Ω ⎫<br />

⎨ ⎬= +<br />

Ω Γ Δ Ω Γ Δ ⎨ ⎬<br />

⎩ΓΔ⎭ ∂y ∂y ∂y ∂y ∂y ∂y<br />

⎩ΘΦ⎭<br />

(56.15)<br />

which generalizes (47.35). Now using (56.15), (56.3), <strong>and</strong> the transformation rule<br />

A<br />

∂y<br />

∂y<br />

= (56.16)<br />

∂ y ∂ y<br />

Γ1<br />

Γr<br />

Γ1…<br />

Γr<br />

Δ1…<br />

Δr<br />

A<br />

Δ1<br />

Δr<br />

we can verify that (56.13) is valid. Thus ∇A , as defined by (56.1), is indeed a tangential tensor<br />

field.<br />

The surface covariant derivative ∇A just defined is not the same as the covariant<br />

derivative defined in Section 47. First, the domain <strong>of</strong> A here is contained in the hypersurface<br />

⎧ Σ ⎫<br />

S , which is not an open set in E . Second, the surface Chris<strong>to</strong>ffel symbols ⎨ ⎬ are generally<br />

⎩ΓΔ⎭ nonvanishing relative <strong>to</strong> any surface coordinate system ( y Γ<br />

) unless S happens <strong>to</strong> be a<br />

hyperplane on which the metric components a ΓΔ <strong>and</strong> a ΓΔ<br />

are constant relative <strong>to</strong> certain


Sec. 56 • Surface Covariant Derivatives 419<br />

“Cartesian” coordinate systems. Other than these two points the formulas for the surface<br />

covariant derivative are formally the same as those for the covariant derivative on E .<br />

In view <strong>of</strong> (56.3) <strong>and</strong> (56.10), we see that the surface covariant derivative <strong>and</strong> the surface<br />

exterior derivative are still related by a formula formally the same as (51.6), namely<br />

r<br />

dA = ( − 1) ( r + 1)! K<br />

+<br />

( ∇A )<br />

(56.17)<br />

r 1<br />

for any surface r-form. Here Kr+<br />

1<br />

denotes the surface skew-symmetric opera<strong>to</strong>r. That is, in<br />

terms <strong>of</strong> any surface coordinate system ( y Γ<br />

)<br />

K<br />

p<br />

1<br />

=<br />

<br />

⊗⊗ ⊗ ⊗ ⊗<br />

(56.18)<br />

1 p 1<br />

p<br />

1 p<br />

1<br />

p<br />

p!<br />

δ Γ Γ Δ<br />

Δ<br />

Δ Δ<br />

h h hΓ hΓ<br />

for any integer p from 1 <strong>to</strong> N-1. Notice that from (56.18) <strong>and</strong> (56.10) the opera<strong>to</strong>r K<br />

p<br />

, like the<br />

surface metric a , is a constant tangential tensor field relative <strong>to</strong> the surface covariant derivative,<br />

i.e.,<br />

δ Γ 1…<br />

Γ =<br />

Δ1<br />

Δ p Σ<br />

p<br />

…<br />

, 0<br />

(56.19)<br />

As a result, the skew-symmetric opera<strong>to</strong>r as well as the opera<strong>to</strong>rs <strong>of</strong> raising <strong>and</strong> lowering <strong>of</strong><br />

indices for tangential fields both commute with the surface covariant derivative.<br />

In Section 48 we have defined the covariant derivative along a smooth curve λ in E .<br />

That concept can be generalized <strong>to</strong> the surface covariant derivative along a smooth curve λ in<br />

t<br />

y Γ<br />

S . Specifically, let λ be represented by ( )<br />

( λ Γ ) relative <strong>to</strong> a surface coordinate system ( )<br />

S , <strong>and</strong> suppose that A is a tangential tensor field on λ represented by<br />

Γ1…<br />

Γ<br />

() () ( )<br />

r<br />

Γ1<br />

( ) Γ ( ( ))<br />

A t = A t h λ t ⊗ ⊗h λ t<br />

(56.20)<br />

r<br />

in<br />

Then we define the surface covariant derivative by the formula<br />

⎡ ⎛ ⎧Γ<br />

⎫ ⎧Γ<br />

⎫⎞<br />

⎤<br />

= ⎢ + ⎜A<br />

⎨ ⎬+ + A ⎨ ⎬⎟<br />

⎥hΓ<br />

⊗ ⊗h<br />

1<br />

Γ<br />

(56.21)<br />

r<br />

Dt ⎢⎣<br />

dt ⎝ ⎩ΔΣ⎭ ⎩ΔΣ⎭⎠<br />

dt ⎥⎦<br />

Γ1…<br />

Γr<br />

Σ<br />

DA<br />

dA ΔΓ2…<br />

Γr<br />

1 Γ1…<br />

Γr<br />

1 dλ<br />

which formally generalizes (48.6). By use <strong>of</strong> (56.15) <strong>and</strong> (56.16) we can show that the surface<br />

covariant derivative DA<br />

/ Dtalong λ is a tangential tensor field on λ independent <strong>of</strong> the choice<br />

<strong>of</strong> the surface coordinate system ( ) y Γ employed in the representation (56.21). Like the surface<br />

covariant derivative <strong>of</strong> a field in S , the surface covariant derivative along a curve commutes


420 Chap. 11 • HYPERSURFACES<br />

with the operations <strong>of</strong> raising <strong>and</strong> lowering <strong>of</strong> indices. Consequently, when the mixed<br />

component representation is used, say with A given by<br />

… r<br />

() () ( )<br />

Δs<br />

( ) ( ( )) ( ( )) ( ())<br />

A t = A<br />

…<br />

t h λ t ⊗⊗h λ t ⊗h λ t ⊗ ⊗h λ t (56.22)<br />

Γ1 Γ Δ1<br />

Δ1 Δs<br />

Γ1 Γ1<br />

the formula for DA<br />

/ Dtbecomes<br />

Γ1…<br />

Γr<br />

⎛ dA<br />

Δ1… Δ<br />

Γ<br />

s 2 r 1<br />

Γ<br />

ΣΓ … Γ Γ1…<br />

Γr−1Σ<br />

r<br />

( A<br />

Δ<br />

A<br />

1…<br />

Δ<br />

<br />

s<br />

Δ1…<br />

Δs<br />

DA<br />

⎧ ⎫ ⎧ ⎫<br />

= + ⎨ ⎬+ +<br />

⎨ ⎬<br />

Dt ⎜<br />

⎝ dt<br />

⎩ΣΩ⎭ ⎩ΣΩ⎭<br />

Ω<br />

Σ<br />

Σ<br />

Γ1…<br />

Γ ⎧ ⎫<br />

r<br />

Γ1…<br />

Γ ⎧ ⎫ dλ<br />

⎞<br />

r<br />

Δ1<br />

−A<br />

ΣΔ<br />

A<br />

)<br />

2…<br />

Δ ⎨ ⎬−− s<br />

Δ1…<br />

Δs−1Σ ⎨ ⎬ ⎟hΓ ⊗⊗h 1 Γ<br />

⊗h ⊗⊗h<br />

1<br />

⎩ΔΩ<br />

1 ⎭ ⎩ΔΩ<br />

s ⎭ dt ⎠<br />

Δs<br />

(56.23)<br />

which formally generalizes (48.7).<br />

As before, if A is a tangential tensor field <strong>and</strong> λ is a curve in the domain <strong>of</strong> A , then the<br />

restriction <strong>of</strong> A on λ is a tensor <strong>of</strong> the form (56.20) or (56.22). In this case the covariant<br />

derivative <strong>of</strong> A along λ is given by<br />

( ( )) = ⎡∇<br />

( )<br />

( ) ⎤ ( )<br />

DA λ t Dt ⎣<br />

A λ t ⎦<br />

λ t<br />

(56.24)<br />

which generalizes (48.15). The component form <strong>of</strong> (56.24) is formally the same as (48.14). A<br />

special case <strong>of</strong> (56.24) is equation (56.5), which is equivalent <strong>to</strong><br />

[ ]<br />

Dh Dy Δ = ∇h h (56.25)<br />

Γ Γ Δ<br />

since from (56.10)<br />

⎧ Σ ⎫<br />

∇ = ⎨ ⎬ ⊗<br />

⎩ΓΩ⎭<br />

Ω<br />

hΓ<br />

hΣ<br />

h (56.26)<br />

If v is a tangential vec<strong>to</strong>r field defined on λ such that<br />

Dv Dt = 0 (56.27)<br />

then v may be called a constant field or a parallel field on λ . From (56.21), v is a parallel field<br />

if <strong>and</strong> only if its components satisfy the equations <strong>of</strong> parallel transport:<br />

Γ<br />

dυ<br />

dt<br />

Δ ⎧ Γ ⎫ Σ<br />

+ υ ⎨ ⎬ = 0, Γ= 1, , N −1<br />

⎩ΔΣ⎭ λ … (56.28)


Sec. 56 • Surface Covariant Derivatives 421<br />

Since ⎧ Γ ⎫<br />

λ Σ<br />

() t <strong>and</strong> ⎨ ⎬( λ () t ) are smooth functions <strong>of</strong> t, it follows from a theorem in ordinary<br />

⎩ΔΣ⎭ differential equations that (56.28) possesses a unique solution<br />

Δ<br />

υ<br />

Δ<br />

= υ ( t), Δ= 1, …, N −1<br />

(56.29)<br />

provided that a suitable initial condition<br />

Δ<br />

υ<br />

Δ<br />

( ) υ0<br />

0 = , Δ= 1, …, N −1<br />

(56.30)<br />

is specified. Since (56.28) is linear in v , the solution v ( t)<br />

<strong>of</strong> (56.27) depends linearly on ( 0)<br />

Thus there exists a linear isomorphism<br />

v .<br />

defined by<br />

ρ : S → S (56.31)<br />

0, t λ 0<br />

( ) λ( t)<br />

( ( )) = ( t)<br />

ρ v v (56.32)<br />

0, t<br />

0<br />

for all parallel fields v on λ . Naturally, we call ρ0,t<br />

the surface covariant derivative.<br />

the parallel transport along λ induced by<br />

The parallel transport ρ<br />

0,t<br />

preserves the surface metric in the sense that<br />

for all u ( 0)<br />

, ( 0)<br />

( ( )) t ( )<br />

0, t<br />

0,<br />

( ) ( ) ( )<br />

ρ v 0 ⋅ ρ u 0 = v 0 ⋅u 0<br />

(56.33)<br />

v in S<br />

λ( 0)<br />

but generally ρ0,t<br />

does not coincide with the Euclidean parallel<br />

transport on E through the translation space V . In fact since S<br />

λ( 0)<br />

<strong>and</strong> S<br />

λ()<br />

need not be the<br />

t<br />

same subspace in V , it is not always possible <strong>to</strong> compare ρ0,t<br />

with the Euclidean parallelism. As<br />

we shall prove later, the parallel transport ρ<br />

0,t<br />

depends not only on the end points λ ( 0)<br />

<strong>and</strong><br />

λ () t but also on the particular curve joining the two points. When the same two end points<br />

λ ( 0)<br />

<strong>and</strong> λ () t are joined by another curve μ in S , generally the parallel transport along μ from<br />

λ ( 0)<br />

<strong>to</strong> () t<br />

λ need not coincide with that along λ .<br />

Dv<br />

Dt<br />

If the parallel transport<br />

tt ,<br />

t λ t is used, the covariant derivative<br />

<strong>of</strong> a vec<strong>to</strong>r field on λ can be defined also by the limit <strong>of</strong> a difference quotient, namely<br />

ρ along λ from λ ( ) <strong>to</strong> ( )


422 Chap. 11 • HYPERSURFACES<br />

Dv<br />

Dt<br />

( )<br />

( t+Δt) − ( t)<br />

v ρtt<br />

, +Δt<br />

v<br />

= lim<br />

Δ→ t 0 Δt<br />

(56.34)<br />

To prove this we observe first that<br />

( t)<br />

Γ<br />

⎛<br />

Γ<br />

dυ<br />

⎞<br />

v( t+Δ t) = ⎜υ<br />

( t)<br />

+ Δ t⎟hΓ<br />

( λ ( t+Δ t)<br />

) + o( Δt)<br />

(56.35)<br />

⎝ dt ⎠<br />

From (56.28) we have also<br />

ρ<br />

( v ())<br />

tt , +Δt t<br />

⎛<br />

Γ Δ ⎧ Γ ⎫<br />

Σ<br />

⎞<br />

= ⎜υ () t −υ () t ⎨ ⎬ () t λ t Δ t⎟<br />

t+Δ t + o Δt<br />

⎝<br />

⎩ΔΣ⎭<br />

⎠<br />

( λ )<br />

() hΓ<br />

( λ ( )) ( )<br />

(56.36)<br />

Substituting these approximations in<strong>to</strong> (56.34), we see that<br />

Δ→ t 0<br />

( ( ))<br />

Γ<br />

v( t+Δt)<br />

−ρtt<br />

, +Δt<br />

v t ⎛dυ<br />

Δ ⎧Γ<br />

⎫ Σ⎞<br />

lim<br />

= + υ ⎨ ⎬<br />

<br />

⎜<br />

λ ⎟hΓ<br />

(56.37)<br />

Δt<br />

⎝ dt ⎩ΔΣ⎭<br />

⎠<br />

which is consistant with the previous formula (56.21) for the surface covariant derivative<br />

Dv D<strong>to</strong>f v along λ .<br />

Since the parallel transport ρtt<br />

,<br />

is a linear isomorphism from S<br />

λ()<br />

<strong>to</strong> S<br />

t λ( t ), it gives rise <strong>to</strong><br />

various induced parallel transport for tangential tensors. We define a tangential tensor field A<br />

on λ a constant field or a parallel field if<br />

DA Dt = 0<br />

(56.38)<br />

Then the equations <strong>of</strong> parallel transport along λ for tensor fields <strong>of</strong> the forms (56.20) are<br />

dA<br />

dt<br />

⎛ ⎧Γ<br />

⎫ ⎧Γ<br />

⎫⎞dλ<br />

+ ⎜A<br />

⎨ ⎬+ + A ⎨ ⎬⎟<br />

= 0<br />

(56.39)<br />

⎝ ⎩ΔΣ⎭ ⎩ΔΣ⎭⎠<br />

dt<br />

Γ1…<br />

Γr<br />

Σ<br />

ΔΓ2…<br />

Γr 1<br />

Γ1…<br />

Γr−1Δ<br />

r<br />

If we donate the induced parallel transport by P<br />

0,t<br />

, or more generally by P<br />

tt ,<br />

, then as before we<br />

have<br />

A A<br />

= lim<br />

D<br />

Dt<br />

Δ→ t 0<br />

( )<br />

( t+Δt) −P<br />

A( t)<br />

Δt<br />

tt , +Δt<br />

(56.40)


Sec. 56 • Surface Covariant Derivatives 423<br />

The coordinate-free definitions (56.34) <strong>and</strong> (56.40) demonstrate clearly the main<br />

difference between the surface covariant derivative on S <strong>and</strong> the ordinary covariant derivative<br />

on the Euclidean manifold E . Specifically, in the former case the parallelism used <strong>to</strong> compute<br />

the difference quotient is the path-dependent surface parallelism, while in the latter case the<br />

parallelism is simply the Euclidean parallelism, which is path-independent between any pair <strong>of</strong><br />

points in E .<br />

In classical differential geometry the surface parallelismρ , or more generally P, defined<br />

by the equation (56.28) or (56.39) is known as the Levi-Civita parallelism or the Riemannian<br />

parallelism. This parallelism is generally path-dependent <strong>and</strong> is determined completely by the<br />

surface Chris<strong>to</strong>ffel symbols.<br />

Exercises<br />

56.1 Use (55.18), (56.5), <strong>and</strong> (56.4) <strong>and</strong> formally derive the formula (56.3).<br />

Γ<br />

56.2 Use (55.9) <strong>and</strong> (56.5) <strong>and</strong> the assumption that Dh<br />

/ Dy<br />

formally derive the formula (56.11).<br />

56.3 Show that<br />

Ω ∂hΔ<br />

Ω Dh<br />

⎧ Ω ⎫<br />

Δ<br />

h ⋅ = h ⋅ =<br />

Γ<br />

Γ ⎨ ⎬<br />

∂y<br />

Dy ⎩ΓΔ⎭<br />

Σ<br />

is a surface vec<strong>to</strong>r field <strong>and</strong><br />

<strong>and</strong><br />

Γ<br />

Γ<br />

∂h<br />

Dh<br />

⎧ Γ ⎫<br />

hΔ⋅ = h<br />

Σ Δ⋅ Σ<br />

=−⎨ ⎬<br />

∂y<br />

Dy ⎩ΣΔ⎭<br />

These formulas show that the tangential parts <strong>of</strong><br />

Dh Dy Γ<br />

Γ<br />

<strong>and</strong> Dh<br />

Dy<br />

Δ<br />

Σ<br />

, respectively.<br />

∂<br />

y Γ<br />

Γ Σ<br />

hΔ<br />

∂ <strong>and</strong> ∂h<br />

∂y<br />

coincide with<br />

56.4 Show that the results <strong>of</strong> Exercise 56.3 can be written<br />

Γ<br />

⎛∂hΔ<br />

⎞ DhΔ<br />

⎛∂h ⎞ Dh<br />

L⎜<br />

= <strong>and</strong> ⎜ ⎟=<br />

∂y ⎟<br />

L<br />

⎝ ⎠ Dy ⎝∂y ⎠ Dy<br />

Γ Γ Σ Σ<br />

Γ<br />

where L is the field whose value at each<br />

55.1.<br />

x ∈S is the projection Lx<br />

defined in Exercise


424 Chap. 11 • HYPERSURFACES<br />

56.5 Use the results <strong>of</strong> Exercise 56.3 <strong>and</strong> show that<br />

∗ ⎛ ∂A<br />

⎞<br />

∇ A= L ⎜ ⊗<br />

Δ<br />

∂y<br />

⎟ h<br />

⎝ ⎠<br />

Δ<br />

where<br />

∗<br />

L is the linear mapping induced by L .<br />

56.6 Adopt the result <strong>of</strong> Exercise 56.5 <strong>and</strong> derive (56.10). This result shows that the above<br />

formula for ∇A can be adopted as the definition <strong>of</strong> the surface gradient. One advantage<br />

<strong>of</strong> this approach is that one does not need <strong>to</strong> introduce the formal operation DA<br />

Dy Δ . If<br />

∗<br />

Δ<br />

L ∂A ∂y<br />

.<br />

needed, it can simply be defined <strong>to</strong> be ( )<br />

56.7 Another advantage <strong>of</strong> adopting the result <strong>of</strong> Exercise 56.5 as a definition is that it can be<br />

used <strong>to</strong> compute the surface gradient <strong>of</strong> field on S which are not tangential. For<br />

example, each gk<br />

can be restricted <strong>to</strong> a field on S but it does not have a component<br />

representation <strong>of</strong> the form (56.9). As an illustration <strong>of</strong> this concept, show that<br />

q l<br />

⎧ j ⎫ ∂x ∂x<br />

Σ<br />

∇<br />

k<br />

= ⎨ ⎬g<br />

jq<br />

⊗<br />

Σ Δ<br />

kl ∂y ∂y<br />

g h h<br />

⎩ ⎭<br />

Δ<br />

<strong>and</strong><br />

2 k k l j s<br />

⎛ ∂ x ⎧ Σ ⎫∂x ⎧k<br />

⎫ ∂x ∂x ⎞ ∂x<br />

Φ Δ Γ<br />

∇ L = ⎜ − g<br />

Γ Δ ⎨ ⎬ +<br />

Σ ⎨ ⎬ Δ Γ ⎟ ks<br />

h ⊗h ⊗ h = 0<br />

Φ<br />

⎝∂y ∂y ⎩ΔΓ⎭∂y ⎩jl⎭∂y ∂y ⎠ ∂y<br />

Note that ∇ L =0 is no more than the result (56.12).<br />

56.8 Compute the surface Chris<strong>to</strong>ffel symbols for the surfaces defined in Excercises 55.5 <strong>and</strong><br />

55.7.<br />

56.9 If A is a tensor field on E, then we can, <strong>of</strong> course, calculate its spatial gradient, grad A.<br />

Also, we can restrict its domain S <strong>and</strong> compute the surface gradient∇A . Show that<br />

56.10 Show that<br />

∂<br />

( det[ aΔΓ]<br />

)<br />

12<br />

∇ =<br />

( grad )<br />

∗<br />

A L A<br />

12⎧<br />

Φ ⎫<br />

= ( det[ a<br />

Σ<br />

ΔΓ]<br />

) ⎨ ⎬<br />

∂y<br />

⎩ΦΣ⎭


Sec. 57 • Surface Geodesics, Exponential Map 425<br />

Section 57. Surface Geodesics <strong>and</strong> the Exponential Map<br />

In Section 48 we pointed out that a straight line λ in E with homogeneous parameter can<br />

be characterized by equation (48.11), which means that the tangent vec<strong>to</strong>r λ <strong>of</strong> λ is constant<br />

along λ , namely<br />

dλ dt = 0<br />

(57.1)<br />

The same condition for a curve in S defines a surface geodesic. In terms <strong>of</strong> any surface<br />

coordinate system ( y Γ<br />

) the equations <strong>of</strong> geodesics are<br />

2 Γ Σ Δ<br />

d d d ⎧ Γ<br />

λ λ λ ⎫<br />

+ 0, 1, , N 1<br />

2 ⎨ ⎬= Γ= … −<br />

(57.2)<br />

dt dt dt ⎩ΣΔ⎭<br />

Since (57.2) is a system <strong>of</strong> second-order differential equations with smooth coefficients, at each<br />

point<br />

<strong>and</strong> in each tangential direction<br />

( )<br />

x = λ ∈S (57.3)<br />

0<br />

0<br />

there exists a unique geodesic = ( t)<br />

( 0)<br />

v = λ ∈S<br />

x<br />

(57.4)<br />

0 0<br />

λ λ satisfying the initial conditions (57.3) <strong>and</strong> (57.4).<br />

In classical calculus <strong>of</strong> variations it is known that the geodesic equations (57.2) represent<br />

the Euler-Lagrange equations for the arc length integral<br />

( )<br />

t1<br />

( λ) = λ() ⋅λ()<br />

∫<br />

t0<br />

12<br />

s t t dt<br />

12<br />

Γ Δ<br />

t1<br />

⎛ dλ<br />

dλ<br />

⎞<br />

= ∫ aΓΔ<br />

( () t )<br />

dt<br />

t<br />

⎜ λ<br />

0<br />

dt dt<br />

⎟<br />

⎝<br />

⎠<br />

(57.5)<br />

between any two fixed points<br />

( t ),<br />

( t )<br />

x = λ y = λ (57.6)<br />

0 1


426 Chap. 11 • HYPERSURFACES<br />

on S . If we consider all smooth curves λ in S joining the fixed end points x <strong>and</strong> y , then the<br />

ones satisfying the geodesic equations are curves whose arc length integral is an extremum in the<br />

class <strong>of</strong> variations <strong>of</strong> curves.<br />

To prove this, we observe first that the arc length integral is invariant under any change<br />

<strong>of</strong> parameter along the curve. This condition is only natural, since the arc length is a geometric<br />

property <strong>of</strong> the point set that constitutes the curve, independent <strong>of</strong> how that point set is<br />

parameterized. From (57.1) it is obvious that the tangent vec<strong>to</strong>r <strong>of</strong> a geodesic must have constant<br />

norm, since<br />

D ( λ ⋅ λ)<br />

Dλ<br />

= 2 ⋅ λ = 0⋅ λ = 0<br />

(57.7)<br />

Dt Dt<br />

Consequently, we seek only those extremal curves for (57.5) on which the integr<strong>and</strong> on the righth<strong>and</strong><br />

<strong>of</strong> (57.5) is constant.<br />

Now if we denote that integr<strong>and</strong> by<br />

L<br />

( λ Γ , λ Γ ) ≡ ( a ( λ Ω ) λ Γ λ<br />

Δ<br />

) 12<br />

ΓΔ<br />

(57.8)<br />

then it is known that the Euler-Lagrange equations for (57.5) are<br />

d ⎛ ∂L ⎞ ∂L<br />

⎜<br />

0, 1, , N 1<br />

Δ ⎟ −<br />

Δ = Δ = −<br />

dt ⎝∂<br />

<br />

…<br />

λ ⎠ ∂λ<br />

(57.9)<br />

From (57.8) we have<br />

∂L<br />

Δ<br />

∂ λ<br />

( )<br />

1 Γ Γ 1 Γ<br />

= a <br />

ΓΔλ + a <br />

ΔΓλ = a <br />

ΓΔλ<br />

2L<br />

L<br />

(57.10)<br />

Hence on the extermal curves we have<br />

d ⎛ ∂L ⎞ 1 ⎧∂aΓΔ<br />

Ω Γ Γ⎫<br />

λ λ aΓΔλ<br />

Δ = ⎨<br />

<br />

Ω + <br />

⎜ ⎟<br />

⎬<br />

dt ⎝∂<br />

λ ⎠ L ⎩∂y<br />

⎭<br />

(57.11)<br />

where we have used the condition that L is constant on those curves. From (57.8) we have also<br />

∂L<br />

Δ<br />

∂λ<br />

1 ∂a<br />

=<br />

2L<br />

∂y<br />

ΓΩ<br />

Δ<br />

Γ<br />

Ω<br />

λ λ<br />

(57.12)<br />

Combining (57.11) <strong>and</strong> (57.12), we see that the Euler-Lagrange equations have the explicit form


Sec. 57 • Surface Geodesics, Exponential Map 427<br />

1 ⎡ 1 a a a<br />

a ⎛∂ ΓΔ<br />

∂<br />

ΩΔ<br />

∂<br />

ΓΩ<br />

⎞ ⎤<br />

λ<br />

λ Γ λ<br />

Ω<br />

ΓΔ<br />

+ + − <br />

⎢<br />

0<br />

Ω Γ Δ ⎥ =<br />

L 2<br />

⎜<br />

∂y ∂y ∂y<br />

⎟<br />

(57.13)<br />

⎣ ⎝<br />

⎠ ⎦<br />

where we have used the symmetry <strong>of</strong> the product λ Γ λ<br />

Ω with respect <strong>to</strong> Γ <strong>and</strong> Ω . Since<br />

L ≠ 0 (otherwise the extremal curve is just one point), (57.13) is equivalent <strong>to</strong><br />

Θ 1 ΘΔ⎛∂aΓΔ ∂aΩΔ ∂aΓΩ<br />

⎞ Γ Ω<br />

λ + a + − λ λ = 0<br />

Ω Γ Δ<br />

2<br />

⎜<br />

∂y ∂y ∂y<br />

⎟<br />

(57.14)<br />

⎝<br />

⎠<br />

which is identical <strong>to</strong> (57.2) upon using the formula (56.3) for the surface Chris<strong>to</strong>ffel symbols.<br />

The preceding result in the calculus <strong>of</strong> variations shows only that a geodesic is a curve<br />

whose arc length is an extremum in a class <strong>of</strong> variations <strong>of</strong> curves. In terms <strong>of</strong> a fixed surface<br />

y Γ we can characterize a typical variation <strong>of</strong> the curve λ by N −1smooth<br />

coordinate system ( )<br />

functions η Γ<br />

() t<br />

such that<br />

Γ<br />

η<br />

Γ<br />

( ) η ( )<br />

t0 = t1 = 0, Γ= 1, …, N −1<br />

(57.15)<br />

A one-parameter family <strong>of</strong> variations <strong>of</strong> λ is then given by the curses λα<br />

with representations.<br />

() t = ( t) + ( t), Γ= 1, , N −1<br />

λ λ αη<br />

Γ Γ Γ<br />

α<br />

… (57.16)<br />

From (57.15) the curves λ<br />

α<br />

satisfy the same end conditions (57.6) as the curve λ , <strong>and</strong><br />

λα<br />

reduces <strong>to</strong> λ when α = 0. The Euler-Lagrange equations express simply the condition that<br />

ds<br />

( λ )<br />

d<br />

a<br />

α<br />

α = 0<br />

= 0<br />

(57.17)<br />

for all choice <strong>of</strong> η Γ satisfying (57.15). We note that (57.17) allows α = 0 <strong>to</strong> be a local minimum,<br />

a local maximum, or a local minimax point in the class <strong>of</strong> variations. In order for the arc length<br />

integral <strong>to</strong> be a local minimum, additional conditions must be imposed on the geodesic.<br />

It can be shown, however, that in a sufficient small surface neighborhood N<br />

x 0<br />

<strong>of</strong> any<br />

point x<br />

0<br />

in S every geodesic is, in fact, a curve <strong>of</strong> minimum arc length. We shall omit the<br />

pro<strong>of</strong> <strong>of</strong> this result since it is not simple. A consequence <strong>of</strong> this result is that between any pair <strong>of</strong><br />

points x <strong>and</strong> y in the surface neighborhood N there exists one <strong>and</strong> only one geodesic λ (aside<br />

x 0


428 Chap. 11 • HYPERSURFACES<br />

from a change <strong>of</strong> parameter) which lies entirely in<br />

kength among all curves in S not just curves in<br />

N<br />

x 0<br />

. That geodesic has the minimum arc<br />

N<br />

x 0<br />

, joining x <strong>to</strong> y .<br />

Now, as we have remarked earlier in this section, at any point x ∈ 0<br />

S , <strong>and</strong> in each<br />

tangential direction v ∈ 0<br />

S<br />

x 0<br />

there exists a unique geodesic λ satisfying the conditions (57.3) <strong>and</strong><br />

(57.4). For definiteness, we donate this particular geodesic by λ<br />

v 0<br />

. Since λ<br />

v 0<br />

is smooth <strong>and</strong>,<br />

when the norm <strong>of</strong> v0<br />

is sufficiently small, the geodesic λ<br />

v ( t)<br />

, t ∈ [ 0,1]<br />

, is contained entirely in<br />

0<br />

the surface neighborhood N<br />

x 0<br />

<strong>of</strong> x<br />

0<br />

. Hence there exists a one-<strong>to</strong>-one mapping<br />

defined by<br />

exp : B → N (57.18)<br />

xo xo xo<br />

x0<br />

( ) ≡ ( )<br />

exp v λ 1<br />

(57.19)<br />

v<br />

for all v belonging <strong>to</strong> a certain small neighborhood<br />

exponential map at x<br />

0<br />

.<br />

Since<br />

v<br />

Bx 0<br />

<strong>of</strong> 0 in<br />

S<br />

x 0<br />

λ is the solution <strong>of</strong> (57.2), the surface coordinates ( y Γ<br />

)<br />

x0<br />

( ) ( )<br />

. We call this injection the<br />

<strong>of</strong> the point<br />

y ≡ exp v = λ 1<br />

(57.20)<br />

depends smoothly on the components υ Γ <strong>of</strong> v relative <strong>to</strong> ( y Γ<br />

) ,i.e., there exists smooth functions<br />

y<br />

1 N 1<br />

( υ υ )<br />

Γ Γ −<br />

0<br />

v<br />

= exp<br />

x<br />

,…, , Γ= 1, …, N −1<br />

(57.21)<br />

where ( υ Γ<br />

) can be regarded as a Cartesian coordinate system on<br />

{ hΓ<br />

( x<br />

0 )}<br />

, namely,<br />

( )<br />

0<br />

Bx 0<br />

induced by the basis<br />

v = υ Γ hΓ<br />

x (57.22)<br />

In the sense <strong>of</strong> (57.21) we say that the exponential map exp x 0<br />

is smooth.<br />

Smoothness <strong>of</strong> the exponential map can be visualized also from a slightly different point<br />

<strong>of</strong> view. Since N<br />

x 0<br />

is contained in S , which is contained in E , exp x<br />

can be regarded also as a<br />

0<br />

mapping from a domain Bx 0<br />

in an Euclidean space S<br />

x 0<br />

<strong>to</strong> the Euclidean space E , namely<br />

exp : B →E (57.23)<br />

x0 x0


Sec. 57 • Surface Geodesics, Exponential Map 429<br />

Now the smoothness <strong>of</strong> exp x<br />

has the usual meaning as defined in Section 43. Since the surface<br />

0<br />

coordinates ( y Γ Γ<br />

) can be extended <strong>to</strong> a local coordinate system ( y , f ) as explained in Section<br />

55, smoothness in the sense <strong>of</strong> (57.21) is consistant with that <strong>of</strong> (57.23).<br />

As explained in Section 43, the smooth mapping<br />

x 0<br />

( )<br />

v 0 ( )<br />

0<br />

exp x<br />

has a gradient at any point v in the<br />

0<br />

domain B . In particular, at = , grad exp x<br />

0 exists <strong>and</strong> corresponds <strong>to</strong> a linear map<br />

( )<br />

grad exp 0 : S →V (57.24)<br />

x0 x0<br />

We claim that the image <strong>of</strong> this linear map is precisely the tangent plane<br />

( )<br />

grad exp x<br />

0 is simply the identity map on<br />

subspace <strong>of</strong> V ; moreover, ( )<br />

0<br />

or less obvious, since by definition the linear map ( )<br />

condition that<br />

for any curve = () t<br />

v v such that v( )<br />

( )<br />

0<br />

S<br />

x 0<br />

S<br />

x 0<br />

, considered as a<br />

. This fact is more<br />

grad exp x<br />

0 is characterized by the<br />

⎡grad( exp ( )) ⎤<br />

d<br />

⎣<br />

x<br />

0 ( ( 0)<br />

) = exp ( () t )<br />

0 ⎦<br />

v x<br />

v<br />

(57.25)<br />

0<br />

dt<br />

=<br />

0 = 0. In particular, for the straight lines<br />

( t)<br />

t<br />

0<br />

v = v t<br />

(57.26)<br />

with<br />

v B , we have v( )<br />

∈ x0<br />

0 = v <strong>and</strong><br />

( t) = ( ) = ( t)<br />

exp v λ 1 λ (57.27)<br />

x0<br />

vt<br />

v<br />

so that<br />

Hence<br />

d<br />

expx ( vt)<br />

= v (57.28)<br />

0<br />

dt<br />

But since ( )<br />

<strong>and</strong> thus<br />

( )<br />

0<br />

( x ( )) ( )<br />

⎡grad exp ⎤<br />

⎣<br />

0 = , ∈<br />

0 ⎦<br />

v v v B<br />

x<br />

(57.29)<br />

0<br />

grad exp x<br />

0 is a linear map, (57.29) implies that the same holds for all<br />

v ∈S x0<br />

,


430 Chap. 11 • HYPERSURFACES<br />

( ( ))<br />

grad expx 0 = idS (57.30)<br />

0 x0<br />

It should be noted, however, that the condition (57.30) is valid only at the origin 0 <strong>of</strong><br />

same condition generally is not true at any other point v ∈B x0<br />

.<br />

Now suppose that { , 1, , N 1}<br />

Cartesian coordinate system ( w Γ<br />

) on<br />

S<br />

x 0<br />

; the<br />

e Γ= … − is any basis <strong>of</strong> Γ<br />

S<br />

x<br />

. Then as usual it gives rise <strong>to</strong> a<br />

0<br />

by the component representation<br />

S<br />

x 0<br />

w = w Γ e<br />

Γ<br />

(57.31)<br />

for any<br />

w ∈S x0<br />

. Since<br />

exp x<br />

is a smooth one-<strong>to</strong>-one map, it carries the Cartesian system<br />

0<br />

( w Γ<br />

) on Bx 0<br />

<strong>to</strong> a surface coordinate system ( z Γ<br />

) on the surface neighborhood N<br />

x 0<br />

<strong>of</strong> x<br />

0<br />

in S .<br />

For definiteness, this coordinate system ( z Γ<br />

) is called a canonical surface coordinate system at<br />

x<br />

0<br />

. Thus a point z ∈N x0<br />

has the coordinates ( z Γ<br />

) if <strong>and</strong> only if<br />

=<br />

x0<br />

( )<br />

z exp z Γ e<br />

Γ<br />

(57.32)<br />

Relative <strong>to</strong> a canonical surface coordinate system a curve λ passing through x<br />

0<br />

at t = 0 is a<br />

surface geodesic if <strong>and</strong> only if its representation ( t)<br />

Γ<br />

λ<br />

λ Γ in terms <strong>of</strong> ( z Γ<br />

)<br />

Γ<br />

( t) = υ t, Γ= 1, , N −1<br />

has the form<br />

… (57.33)<br />

for some constant υ Γ . From (57.32) the geodesic λ is simply the one denoted earlier by λ<br />

v<br />

,<br />

where<br />

v = υ Γ e<br />

Γ<br />

(57.34)<br />

An important property <strong>of</strong> a canonical surface coordinate system at x<br />

0<br />

is that the<br />

corresponding surface Chris<strong>to</strong>ffel symbols are all equal <strong>to</strong> zero at x<br />

0<br />

. Indeed, since any curve λ<br />

given by (57.33) is a surface geodesic, the geodesic equations imply<br />

⎧ Σ ⎫<br />

⎨ ⎬υυ<br />

Γ Δ<br />

⎩ΓΔ⎭<br />

= 0<br />

(57.35)<br />

where the Chris<strong>to</strong>ffel symbols are evaluated at any point <strong>of</strong> the curve λ . In particular, at 0 t =<br />

we get


Sec. 57 • Surface Geodesics, Exponential Map 431<br />

⎧ Σ ⎫<br />

( )<br />

υυ<br />

⎨ ⎬<br />

Γ Δ<br />

0<br />

= 0<br />

⎩ΓΔ⎭ x (57.36)<br />

Now since υ Γ is arbitrary, by use <strong>of</strong> the symmetry condition (56.4) we conclude that<br />

point<br />

0<br />

⎧ Σ ⎫<br />

( )<br />

⎨ ⎬ 0<br />

= 0<br />

⎩ΓΔ⎭ x (57.37)<br />

In general, a surface coordinate system ( y Γ<br />

) is called a geodesic coordinate system at a<br />

x if the Chris<strong>to</strong>ffel symbols corresponding <strong>to</strong> ( y Γ ) vanish at x<br />

0<br />

. From (57.37), we see<br />

that a canonical surface coordinate system at x0<br />

is a geodesic coordinate system at the same<br />

point. The converse, <strong>of</strong> course, is not true. In fact, from the transformation rule (56.15) a<br />

y Δ is geodesic at x<br />

0<br />

if <strong>and</strong> only if its coordinate transformation<br />

surface coordinate system ( )<br />

relative <strong>to</strong> a canonical surface coordinate system ( z Γ<br />

) satisfies the condition<br />

2 Δ<br />

∂ y<br />

Γ Ω<br />

∂z<br />

∂z<br />

x<br />

0<br />

= 0<br />

(57.38)<br />

which is somewhat weaker that the transformation rule<br />

for some nonsingular matrix ⎡e Δ Γ<br />

⎤<br />

the ⎡<br />

⎣e Δ Γ<br />

⎤<br />

⎦<br />

{ } Γ<br />

e for ( )<br />

⎣ ⎦ when ( )<br />

y<br />

= ez<br />

(57.39)<br />

Δ Δ Γ<br />

Γ<br />

y Δ is also a canonical surface coordinate system at x<br />

0<br />

,<br />

being simply the transformation matrix connecting the basis { } Γ<br />

y Γ .<br />

e for ( z Γ<br />

)<br />

<strong>and</strong> thebasis<br />

A geodesic coordinate system at x<br />

0<br />

plays a role similar <strong>to</strong> that <strong>of</strong> a Cartesian coordinate<br />

system in E . In view <strong>of</strong> the condition (57.37), we see that the representation <strong>of</strong> the surface<br />

covariant derivative at x<br />

0<br />

reduces simply <strong>to</strong> the partial derivative at x<br />

0<br />

, namely<br />

( x )<br />

∂A<br />

Δ<br />

x e e e<br />

(57.40)<br />

∂z<br />

Γ1…<br />

Γr<br />

0<br />

∇ A( 0 ) =<br />

Δ Γ<br />

⊗ ⊗<br />

1 Γ<br />

⊗<br />

1


432 Chap. 11 • HYPERSURFACES<br />

It should be noted, however, that (57.40) is valid at the point x<br />

0<br />

only, since a geodesic<br />

coordinate system at x0<br />

generally does not remain a geodesic coordinate system at any<br />

neighboring point <strong>of</strong> x<br />

0<br />

. Notice also that the basis { e Γ } in S<br />

x 0<br />

is the natural basis <strong>of</strong> ( z Δ ) at x<br />

0<br />

,<br />

this fact being a direct consequence <strong>of</strong> the condition (57.30).<br />

In closing, we remark that we can choose the basis { e Γ } in<br />

S<br />

x 0<br />

<strong>to</strong> be an orthonormal basis.<br />

In this case the corresponding canonical surface coordinate system ( z Γ<br />

) satisfies the additional<br />

condition<br />

a<br />

ΓΔ<br />

( )<br />

x = δ<br />

(57.41)<br />

0<br />

ΓΔ<br />

Then we do not even have <strong>to</strong> distinguish the contravariant <strong>and</strong> the covariant component <strong>of</strong> a<br />

tangential tensor at<br />

0<br />

x . In classical differential geometry such a canonical surface coordinate<br />

system is called a normal coordinate system or a Riemannian coordinate system at the surface<br />

point under consideration.


Sec. 58 • Surface Curvature, I 433<br />

Section 58. Surface Curvature, I. The Formulas <strong>of</strong> Weingarten <strong>and</strong> Gauss<br />

In the preceding sections we have considered the surface covariant derivative at<br />

tangential vec<strong>to</strong>r <strong>and</strong> tensor fields on S . We have pointed out that this covariant derivative is<br />

defined relative <strong>to</strong> a particular path-dependent parallelism onS , namely the parallelism <strong>of</strong> Levi-<br />

Civita as defined by (56.28) <strong>and</strong> (56.39). We have remarked repeatedly that this parallelism is<br />

not the same as the Euclidean parallelism on the underlying space E in which S is embedded.<br />

Now a tangential tensor on S , or course, is also a tensor over E being merely a tensor having<br />

, which can be regarded as a subset <strong>of</strong> the<br />

n for V . Consequently, the spatial covariant derivative <strong>of</strong> a tangential tensor field<br />

nonzero components only in the product basis <strong>of</strong> { h Γ }<br />

basis { h }<br />

Γ ,<br />

along a curve in S is defined. Similarly, the special covariant derivative <strong>of</strong> the unit normal n <strong>of</strong><br />

S along a curve in S is also defined. We shall study these spatial covariant derivatives in this<br />

section.<br />

In classical differential geometry the spatial covariant derivative <strong>of</strong> a tangential field is a<br />

special case <strong>of</strong> the <strong>to</strong>tal covariant derivative, which we shall consider in detail later. Since the<br />

spatial covariant derivative <strong>and</strong> the surface covariant derivative along a surface curve <strong>of</strong>ten<br />

appear in the same equation, we use the notion d dtfor the former <strong>and</strong> DDtfor the latter.<br />

However, when the curve is the coordinate curve <strong>of</strong> y Γ , we shall write the covariant derivative as<br />

∂∂y Γ <strong>and</strong> DDy Γ , respectively. It should be noted also that d dtis defined for all tensor fields<br />

on λ , whether or not the field is tangential, while DDtis defined only for tangential fields.<br />

We consider first the covariant derivative dn<br />

d<strong>to</strong>f the unit normal field <strong>of</strong> S on any<br />

curve λ ∈S . Since n is a unit vec<strong>to</strong>r field <strong>and</strong> since d dt preserves the spatial metric on S , we<br />

have<br />

dn dt⋅ n = 0<br />

(58.1)<br />

Thus dn<br />

dtis a tangential vec<strong>to</strong>r field. We claim that there exists a symmetric second-order<br />

tangential tensor field B on S such that<br />

dn dt = −Bλ (58.2)<br />

for all surface curves λ . The pro<strong>of</strong> <strong>of</strong> (58.2) is more or less obvious. From (55.2), the unit<br />

normal n on S is parallel <strong>to</strong> the gradient <strong>of</strong> a certain smooth function f , <strong>and</strong> locally S can be<br />

characterized by (55.1). By normalizing the function f <strong>to</strong> a function<br />

w<br />

( x)<br />

=<br />

f<br />

grad<br />

( x)<br />

f ( x)<br />

(58.3)


434 Chap. 11 • HYPERSURFACES<br />

we find<br />

( ) w( x)<br />

n x = grad , x∈S (58.4)<br />

Now for the smooth vec<strong>to</strong>r field grad w we can apply the usual formula (48.15) <strong>to</strong> compute the<br />

spatial covariant derivative ( d dt)( grad w ) along any smooth curve in E . In particular, along<br />

the surface curve λ under consideration we have the formula (58.2), where<br />

B =−grad ( grad w)<br />

(58.5)<br />

S<br />

As a result, B is symmetric. The fact that B is a tangential tensor has been remarked after<br />

(58.1).<br />

From (58.2) the surface tensor B characterizes the spatial change <strong>of</strong> the unit normal n <strong>of</strong> S .<br />

Hence in some sense B is a measurement <strong>of</strong> the curvature <strong>of</strong> S in E . In classical differential<br />

geometry, B is called the second fundamental form <strong>of</strong> S , the surface metric a defined by<br />

(55.18) being the first fundamental form. In component form relative <strong>to</strong> a surface coordinate<br />

y Γ , B can be represented as usual<br />

system ( )<br />

Γ Δ ΓΔ Γ Δ<br />

B= b h ⊗ h = b h ⊗ h = b h ⊗h (58.6)<br />

ΓΔ Γ Δ Δ Γ<br />

From (58.2) the components <strong>of</strong> B are those <strong>of</strong><br />

∂n<br />

∂<br />

y Γ<br />

taken along the y Γ -curve in S , namely<br />

Γ Δ Δ<br />

∂n ∂ y =− b h =−b<br />

h (58.7)<br />

Γ Δ ΓΔ<br />

This equation is called Weingarten’s formula in classical differential geometry.<br />

Γ<br />

Next we consider the covariant derivatives dh dt<strong>and</strong> dh<br />

d<strong>to</strong>f the surface natural basis<br />

vec<strong>to</strong>rs hΓ<br />

<strong>and</strong><br />

Γ<br />

h along any curve λ in S . From (56.21) <strong>and</strong> (56.23) we have<br />

Γ<br />

<strong>and</strong><br />

Dh<br />

Dt<br />

Dh<br />

Dt<br />

Γ<br />

Γ<br />

⎧ Δ ⎫<br />

= ⎨ ⎬<br />

λ Σ hΔ<br />

(58.8)<br />

⎩ΓΣ⎭<br />

⎧ Γ ⎫ Σ Δ<br />

=−⎨ ⎬<br />

λ h<br />

(58.9)<br />

⎩ΔΣ⎭<br />

By a similar argument as (58.2), we have first


Sec. 58 • Surface Curvature, I 435<br />

dh<br />

dt<br />

Γ<br />

=−<br />

Γ<br />

C λ (58.10)<br />

Γ<br />

where C is a symmetric spatial tensor field on S . In particular, when λ is the coordinate curve<br />

<strong>of</strong> y Δ , (58.10) reduces <strong>to</strong><br />

∂h<br />

∂y<br />

Γ<br />

Δ<br />

=−<br />

Γ<br />

Ch<br />

Δ<br />

(58.11)<br />

Since<br />

Γ<br />

C is symmetric, this equation implies<br />

Γ<br />

Γ<br />

∂h<br />

∂h<br />

Γ<br />

−hΣ⋅ = −hΔ⋅ = C ( hΔ,<br />

h<br />

Δ<br />

Σ<br />

Σ)<br />

(58.12)<br />

∂y<br />

∂y<br />

We claim that the quantity given by this equation is simply the surface Chris<strong>to</strong>ffel symbol,<br />

⎧ Γ ⎫<br />

⎨ ⎬<br />

⎩ΔΣ⎭<br />

Γ<br />

∂h<br />

⎧ Γ ⎫<br />

−h Σ<br />

⋅ =<br />

Δ ⎨ ⎬<br />

(58.13)<br />

∂y<br />

⎩ΔΣ⎭<br />

Indeed, since both d dt<strong>and</strong> DDtpreserve the surface metric, we have<br />

Γ<br />

( Σ<br />

⋅ )<br />

Γ<br />

Γ<br />

∂δ<br />

∂ h h<br />

Σ<br />

∂h ∂hΣ<br />

Γ<br />

0 = = = h<br />

Δ Δ Σ<br />

⋅ + ⋅h (58.14)<br />

Δ Δ<br />

∂y ∂y ∂y ∂y<br />

Thus (58.13) is equivalent <strong>to</strong><br />

Γ ∂h<br />

⎧ Γ ⎫<br />

Σ<br />

h ⋅ =<br />

Δ ⎨ ⎬<br />

(58.15)<br />

∂y<br />

⎩ΣΔ⎭<br />

But as in (58.14) we have also<br />

( ⋅ )<br />

∂a<br />

∂ hΓ<br />

hΔ<br />

∂h ∂h<br />

= = hΓ⋅ + hΔ⋅<br />

∂y ∂y ∂y ∂y<br />

ΓΔ Δ Γ<br />

Σ Σ Σ Σ<br />

⎛ Ω ∂hΔ<br />

⎞ ⎛ Ω ∂hΓ<br />

⎞<br />

= aΓΩ<br />

⎜h<br />

⋅ + a<br />

Σ ΔΩ<br />

⋅<br />

Σ<br />

∂y<br />

⎟ ⎜h<br />

∂y<br />

⎟<br />

⎝ ⎠ ⎝ ⎠<br />

(58.16)


436 Chap. 11 • HYPERSURFACES<br />

Further, from (58.14) <strong>and</strong> (58.12) we have<br />

Ω ∂hΔ<br />

Ω ∂hΣ<br />

h ⋅ = h ⋅<br />

(58.17)<br />

Σ<br />

Δ<br />

∂y<br />

∂y<br />

Comparing (58.17) <strong>and</strong> (58.16) with (56.4) <strong>and</strong> (56.12), respectively, we see that (58.15) holds.<br />

On differentiating (55.8), we get<br />

Γ<br />

( )<br />

∂ n⋅h Γ<br />

∂h Γ ∂n<br />

0 = = n⋅ + h ⋅<br />

(58.18)<br />

Δ Δ Δ<br />

∂y ∂y ∂y<br />

Substituting Weingarten’s formula (58.7) in<strong>to</strong> (58.18), we then obtain<br />

Γ<br />

∂h<br />

Γ<br />

n ⋅ = b<br />

Δ Δ<br />

(58.19)<br />

∂y<br />

The formulas (58.19) <strong>and</strong> (58.13) determine completely the spatial covariant derivative <strong>of</strong><br />

Γ<br />

h along any y Δ -curve:<br />

Γ<br />

Γ<br />

∂h<br />

Γ ⎧ Γ ⎫ Σ Γ Dh<br />

= bΔn− b<br />

Δ ⎨ ⎬h =<br />

Δn +<br />

(58.20)<br />

Δ<br />

∂y<br />

⎩ΔΣ⎭<br />

Dy<br />

where we have used (58.9) for (58.20) 2<br />

. By exactly the same argument we have also<br />

∂h<br />

⎧ Σ ⎫<br />

Γ<br />

DhΓ<br />

= bΓΔn+ Σ<br />

b<br />

Δ ⎨ ⎬h =<br />

ΓΔn +<br />

(58.21)<br />

Δ<br />

∂y<br />

⎩ΓΔ⎭<br />

Dy<br />

As we shall see, (58.20) <strong>and</strong> (58.21) are equivalent <strong>to</strong> the formula <strong>of</strong> Gauss in classical<br />

differential geometry.<br />

The formulas (58.20), (58.21), <strong>and</strong> (58.7) determine completely the spatial covariant<br />

h<br />

Γ , n along any curve λ in S . Indeed, if the coordinates<br />

derivatives <strong>of</strong> the bases { } Γ ,<br />

<strong>of</strong> λ in ( y Γ ) are ( λ Γ<br />

)<br />

h n <strong>and</strong> { }<br />

, then we have


Sec. 58 • Surface Curvature, I 437<br />

dn<br />

Δ Γ Γ Δ<br />

=− b <br />

ΓΔλ<br />

h =−b<br />

<br />

Δλ<br />

hΓ<br />

dt<br />

Γ<br />

Γ<br />

dh<br />

Γ Δ ⎧ Γ ⎫ Δ Σ Γ Δ D<br />

= b h<br />

Δλ n− ⎨ ⎬<br />

λ h = b <br />

Δλ<br />

n+<br />

dt ⎩ΔΣ⎭<br />

Dt<br />

dhΓ<br />

Δ ⎧ Σ ⎫ Δ Δ D<br />

= b h<br />

ΓΔλ n+ ⎨ ⎬<br />

λ hΣ = b <br />

ΓΔλ<br />

n+<br />

dt<br />

⎩ΓΔ⎭<br />

Dt<br />

Γ<br />

(58.22)<br />

From these representations we can compute the spatial covariant derivative <strong>of</strong> any vec<strong>to</strong>r or<br />

tensor fields along λ when their components relative <strong>to</strong> { h , Γ<br />

n}<br />

or { h<br />

Γ , n}<br />

are given. For<br />

example, if v is a vec<strong>to</strong>r field having the component form<br />

() ( ) ( )<br />

n<br />

( ) ( ) ( ( ))<br />

v t = υ t n λ t + υ Γ t hΓ<br />

λ t<br />

(58.23)<br />

along λ , then<br />

dv<br />

dt<br />

Γ Δ<br />

⎛<br />

Γ Γ Δ ⎧ Γ ⎫ Σ Δ⎞<br />

= ( υn<br />

+ b υ <br />

ΓΔ<br />

λ ) n+ ⎜ υ − υnb<br />

<br />

Δλ +⎨ ⎬υ λ ⎟hΓ<br />

(58.24)<br />

⎝<br />

⎩ΣΔ⎭<br />

⎠<br />

In particular, if v is a tangential field, then (58.24) reduces <strong>to</strong><br />

dv<br />

dt<br />

Γ Δ<br />

⎛<br />

Γ ⎧ Γ ⎫ Σ Δ⎞<br />

= bΓΔυλ<br />

n+ υ +⎨ ⎬υλ<br />

<br />

⎜<br />

⎟h<br />

⎝ ⎩ΣΔ⎭<br />

⎠<br />

Γ Δ Dv<br />

= bΓΔυλ<br />

n +<br />

Dt<br />

Γ<br />

(58.25)<br />

where we have used (56.21). Equation (58.25) shows clearly the difference between d dt<strong>and</strong><br />

DDtfor any tangential field.<br />

Applying (58.25) <strong>to</strong> the tangent vec<strong>to</strong>r λ <strong>of</strong> λ , we get<br />

dλ<br />

dt<br />

Dλ<br />

= B( λλ ,<br />

<br />

) n +<br />

(58.26)<br />

Dt<br />

In particular, when λ is a surface geodesic, then (58.26) reduces <strong>to</strong><br />

( , )<br />

dλ<br />

dt = B λλ n<br />

(58.27)


438 Chap. 11 • HYPERSURFACES<br />

Comparing this result with the classical notions <strong>of</strong> curvature <strong>and</strong> principal normal <strong>of</strong> a curve (cf.<br />

Section 53), we see that the surface normal is the principal normal <strong>of</strong> a surface curve λ if <strong>and</strong><br />

only if λ is a surface geodesic; further, in this case the curvature <strong>of</strong> λ is the quadratic form<br />

B s,<br />

s <strong>of</strong> B in the direction <strong>of</strong> the unit tangent s <strong>of</strong> λ .<br />

( )<br />

In general, if λ is not a surface geodesic but it is parameterized by the arc length, then<br />

(58.26) reads<br />

ds<br />

Ds<br />

= B( s,<br />

s)<br />

n +<br />

(58.28)<br />

ds<br />

Ds<br />

where s denotes the unit tangent <strong>of</strong> λ , as usual. We call the norms <strong>of</strong> the three vec<strong>to</strong>rs in<br />

(58.28) the spatial curvature, the normal curvature, <strong>and</strong> the geodesic curvature <strong>of</strong> λ <strong>and</strong> denote<br />

them by κκ ,<br />

n,<br />

<strong>and</strong> κ<br />

g<br />

, respectively, namely<br />

ds<br />

Ds<br />

κ = , κn<br />

= B( s, s ),<br />

κg<br />

=<br />

(58.29)<br />

ds<br />

Ds<br />

Then (58.28) implies<br />

κ = κ + κ<br />

(58.30)<br />

2 2 2<br />

n g<br />

Further, if n0<br />

<strong>and</strong><br />

n are unit vec<strong>to</strong>rs in the directions <strong>of</strong> ds ds<strong>and</strong> Ds<br />

Ds, then<br />

g<br />

κn = κ n+<br />

κ n (58.31)<br />

0 n g g<br />

Or equivalently,<br />

( κ κ) ( κ κ)<br />

n = n+<br />

n (58.32)<br />

0 n g g<br />

which is called Meunier’s equation.<br />

At this point we can define another kind <strong>of</strong> covariant derivative for tensor fields on S .<br />

As we have remarked before, the tangent plane S<br />

x<br />

<strong>of</strong> S is a subspace <strong>of</strong> V . Hence a tangential<br />

vec<strong>to</strong>r v can be regarded either as a surface vec<strong>to</strong>r in S<br />

x<br />

or as a special vec<strong>to</strong>r in V . In the<br />

former sense it is natural <strong>to</strong> use the surface covariant derivative Dv<br />

Dt, while in the latter sense<br />

we may consider the spatial covariant derivative dv<br />

dtalong any smooth curve λ in S . For a<br />

tangential tensor field A , such as the one represented (56.20) or (56.22), we may choose <strong>to</strong><br />

recognize certain indices as spatial indices <strong>and</strong> the remaining ones as surface indices. In this


Sec. 58 • Surface Curvature, I 439<br />

case it becomes possible <strong>to</strong> define a covariant derivative tha is a mixture <strong>of</strong> d dt<strong>and</strong> DDt. In<br />

classical differential geometry this new kind <strong>of</strong> covariant derivative is called the <strong>to</strong>tal covariant<br />

derivative.<br />

More specifically, we first consider a concrete example <strong>to</strong> explain this concept. By<br />

definition, the surface metric tensor a is a constant tangential tensor relative <strong>to</strong> the surface<br />

covariant derivative. In terms <strong>of</strong> any surface coordinate system ( y Γ<br />

) , a can be represented by<br />

<strong>and</strong> we have<br />

Γ Δ Δ ΓΔ<br />

a= a h ⊗ h = h ⊗ h = a h ⊗h (58.33)<br />

ΓΔ Δ Γ Δ<br />

Δ<br />

D ( hΔ<br />

⊗ h )<br />

Δ<br />

Da Dh DhΔ<br />

Δ<br />

0= = = hΔ<br />

⊗ + ⊗h (58.34)<br />

Dt Dt Dt Dt<br />

The tensor a , however, can be regarded also as the inclusion map A <strong>of</strong> S<br />

x<br />

in V , namely<br />

A : S x<br />

→V (58.35)<br />

in the sense that for any surface vec<strong>to</strong>r<br />

∈ x<br />

v S , Av ∈V is given by<br />

Δ Δ Δ<br />

( )( ) ( ) υ<br />

Av = h ⊗ h v = h h ⋅ v = h (58.36)<br />

Δ Δ Δ<br />

Δ<br />

In (58.36) it is more natural <strong>to</strong> regard the first basis vec<strong>to</strong>r hΔ<br />

in the product basis h ⊗ h as a<br />

Δ<br />

Δ<br />

spatial vec<strong>to</strong>r <strong>and</strong> the second basis vec<strong>to</strong>r h as a surface vec<strong>to</strong>r. In fact, in classical differential<br />

geometry the tensor A given by (58.35) is <strong>of</strong>ten denoted by<br />

i<br />

i Δ ∂x<br />

Δ<br />

A = hΔgi<br />

⊗ h = g<br />

Δ i<br />

⊗h (58.37)<br />

∂y<br />

i<br />

where ( )<br />

i<br />

. When the indices are<br />

recognized in this way, it is natural <strong>to</strong> define a <strong>to</strong>tal covariant derivative <strong>of</strong> A , denoted by<br />

δ A δt , by<br />

x is a spatial coordinate system whose natural basis is { g }<br />

δ A dhΔ<br />

Δ Dh<br />

= ⊗ h + hΔ<br />

⊗<br />

δt dt Dt<br />

i<br />

d ⎛ ∂x ⎞ Δ Dh<br />

=<br />

Δ i<br />

Δ<br />

dt<br />

⎜ g ⊗ + ⊗<br />

∂y ⎟ h h<br />

⎝ ⎠<br />

Dt<br />

Δ<br />

Δ<br />

(58.38)


440 Chap. 11 • HYPERSURFACES<br />

Since a mixture <strong>of</strong> d dt<strong>and</strong> DDtis used in definingδ δ t , it is important <strong>to</strong> recognize the<br />

“spatial” or “surface” designation <strong>of</strong> the indices <strong>of</strong> the components <strong>of</strong> a tensor before the <strong>to</strong>tal<br />

covariant derivative is computed.<br />

In classical differential geometry the indices are distinguished directly in the notation <strong>of</strong><br />

the components. Thus a tensor A represented by<br />

A<br />

ik… Γ…<br />

() t A ( t)<br />

= ⊗ ⊗ ⊗ ⊗ ⊗<br />

j<br />

Δ<br />

j… Δ… gi g gk<br />

hΓ<br />

h (58.39)<br />

has an obvious interpretation: the Latin indices i, j, k…are designated as “spatial” <strong>and</strong> the Greek<br />

indices ΓΔ…are , , “surface.” So δ A δt<br />

is defined by<br />

δ A d<br />

( A<br />

ik… Γ…<br />

j<br />

Δ<br />

=<br />

j… Δ…<br />

gi ⊗g ⊗gk<br />

⊗)<br />

⊗hΓ<br />

⊗h<br />

<br />

δt<br />

dt<br />

ik… Γ…<br />

j<br />

D<br />

Δ<br />

+ Aj … Δ…<br />

gi ⊗g ⊗gk<br />

⊗⊗ ( hΓ<br />

⊗h<br />

)<br />

Dt<br />

⎛ d A<br />

ik… Γ…<br />

⎞ j<br />

Δ<br />

= ⎜ j… Δ…<br />

⎟gi ⊗g ⊗gk<br />

⊗⊗hΓ<br />

⊗h<br />

<br />

⎝dt<br />

⎠<br />

ik… Γ…<br />

⎡ d<br />

j<br />

Δ<br />

+ Aj … Δ…<br />

⎢ ( gi ⊗g ⊗gk<br />

⊗)<br />

⊗hΓ<br />

⊗h<br />

<br />

⎣dt<br />

j<br />

D<br />

Δ ⎤<br />

+ gi<br />

⊗g ⊗gk<br />

⊗⊗ ( hΓ<br />

⊗h<br />

)<br />

Dt<br />

⎥<br />

⎦<br />

(58.40)<br />

The derivative<br />

dA<br />

ik… Γ…<br />

j… Δ…<br />

dt is the same as DA<br />

ik… Γ…<br />

j… Δ…<br />

Dt , <strong>of</strong> course, since for scalars there is but<br />

one kind <strong>of</strong> parallelism along any curve. In particular, we can compute explicitly the<br />

representation for δ A δt<br />

as defined by (58.38):<br />

i<br />

i<br />

Δ<br />

δ A d ⎛ ∂x ⎞ Δ ∂x ⎛dgi<br />

Δ Dh<br />

⎞<br />

= ⎜ Δ ⎟gi<br />

⊗ h +<br />

Δ ⎜ ⊗ h + gi<br />

⊗ ⎟<br />

δt dt⎝∂y ⎠ ∂y ⎝ dt Dt ⎠<br />

2 i j k i<br />

⎛ ∂ x ∂x ∂x ⎧ i ⎫ ∂x<br />

⎧ Σ ⎫ Γ<br />

= + λ<br />

Δ Γ Δ Γ ⎨ ⎬− Σ ⎨ ⎬<br />

<br />

⎜<br />

gi<br />

⊗h<br />

⎝∂y ∂y ∂y ∂y ⎩jk<br />

⎭ ∂y<br />

⎩ΓΔ⎭<br />

Δ<br />

⎞<br />

⎟<br />

⎠<br />

(58.41)<br />

As a result, when λ is the coordinate curve <strong>of</strong> ( y Γ<br />

) , (58.41) reduces <strong>to</strong><br />

2 i j k i<br />

δ A ⎛ ∂ x ∂x ∂x ⎧ i ⎫ ∂x<br />

⎧ Σ ⎫⎞<br />

Δ<br />

=<br />

Γ ⎜ +<br />

Δ Γ Δ Γ ⎨ ⎬− Σ ⎨ ⎬⎟gi<br />

⊗h (58.42)<br />

δ y ⎝∂y ∂y ∂y ∂y ⎩jk<br />

⎭ ∂y<br />

⎩ΓΔ⎭⎠


Sec. 58 • Surface Curvature, I 441<br />

Actually, the quantity on the right-h<strong>and</strong> side <strong>of</strong> (58.42) has a very simple representation.<br />

Since δ A δ y<br />

Γ can be expressed also by (58.38) 1<br />

, namely<br />

δ A Δ<br />

∂<br />

Δ Δ D<br />

= h ⊗ h + h<br />

Γ Γ Δ<br />

⊗<br />

h<br />

(58.43)<br />

Γ<br />

δ y ∂y Dy<br />

from (58.21) we have<br />

δ A Δ DhΔ<br />

Δ Dh<br />

= bΔΓn⊗ h + ⊗ h + hΔ<br />

⊗<br />

δ y Dy Dy<br />

Γ Γ Γ<br />

Δ D<br />

= bΔΓn⊗ h + h ⊗h<br />

Γ<br />

Dy<br />

= b n⊗h<br />

ΔΓ<br />

Δ<br />

Δ<br />

( Δ )<br />

Δ<br />

(58.44)<br />

where we have used (58.34). The formula (58.44) 3<br />

is known as Gauss’ formula in classical<br />

differential geometry. It is <strong>of</strong>ten written in component form<br />

x<br />

i<br />

; ΓΔ<br />

i<br />

= bΓΔn<br />

(58.45)<br />

where the semicolon in the subscript on the left-h<strong>and</strong> side denotes the <strong>to</strong>tal covariant derivative.<br />

Exercises<br />

58.1 Show that<br />

(a)<br />

b<br />

1 ⎛ ∂n ∂x ∂x ∂n<br />

⎞<br />

=− ⋅ + ⋅<br />

2<br />

⎜<br />

∂y ∂y ∂y ∂y<br />

⎟<br />

⎝<br />

⎠<br />

ΓΔ Δ Γ Γ Γ<br />

(b) b = g x i ; n<br />

ΔΓ<br />

ij<br />

ΔΓ<br />

j<br />

(c)<br />

b<br />

2<br />

∂ x<br />

= n ⋅ ΔΓ ∂ Δ ∂ Γ<br />

y<br />

y<br />

58.2 Compute the quantities b ΔΓ<br />

for the surfaces defined in Exercises 55.5 <strong>and</strong> 55.7.<br />

58.3 Let A be a tensor field <strong>of</strong> the form<br />

A = A g ⊗⊗g ⊗h ⊗<br />

⊗h<br />

j1<br />

jr<br />

Γ1<br />

Γ1<br />

Γs<br />

j1<br />

jr<br />

Γs


442 Chap. 11 • HYPERSURFACES<br />

show that<br />

δ A j1 jr<br />

Γ1<br />

A<br />

Δ Γ1 Γs;<br />

Δ j1<br />

jr<br />

δ y<br />

= <br />

g ⊗ ⊗ g ⊗ h ⊗ ⊗ h<br />

<br />

Γs<br />

where<br />

A<br />

∂A<br />

j1…<br />

jr<br />

j1<br />

jr<br />

Γ1…<br />

Γs<br />

Γ1<br />

Γs<br />

; Δ<br />

=<br />

<br />

Δ<br />

∂y<br />

r<br />

⎧jβ<br />

⎫ ∂x<br />

+ ∑ ⎨ ⎬A<br />

β = 1 ⎩lk<br />

⎭<br />

∂y<br />

s<br />

⎧ Λ ⎫<br />

j1<br />

jr<br />

−∑<br />

⎨ ⎬A<br />

Γ1Γβ− 1ΛΓβ+<br />

1Γs<br />

β = 1 ⎩ΓΔ<br />

β ⎭<br />

k<br />

j1jβ− 1ljβ+<br />

1jr<br />

Γ1…<br />

Γs<br />

Δ<br />

Similar formulas can be derived for other types <strong>of</strong> mixed tensor fields defined on S .<br />

58.4 Show that (58.7) can be written<br />

n<br />

∂x<br />

∂y<br />

i<br />

i Δ<br />

; Γ<br />

=−bΓ Δ


Sec. 59 • Surface Curvature, II 443<br />

Section 59. Surface Curvature, II. The Riemann-Chris<strong>to</strong>ffel Tensor <strong>and</strong> the Ricci<br />

Identities<br />

In the preceding section we have considered the curvature <strong>of</strong> S by examining the change<br />

<strong>of</strong> the unit normal n <strong>of</strong> S in E . This approach is natural for a hypersurface, since the metric on<br />

S is induced by that <strong>of</strong> E . The results <strong>of</strong> this approach, however, are not entirely intrinsic <strong>to</strong><br />

S , since they depend not only on the surface metric but also on the particular imbedding <strong>of</strong> S<br />

in<strong>to</strong> E . In this section, we shall consider curvature from a more intrinsic point <strong>of</strong> view. We<br />

seek results which depend only on the surface metric. Our basic idea is that curvature on S<br />

corresponds <strong>to</strong> the departure <strong>of</strong> the Levi-Civita parallelism on S from a Euclidean parallelism.<br />

We recall that relative <strong>to</strong> a Cartesian coordinate system ˆx on E the covariant derivative<br />

<strong>of</strong> a vec<strong>to</strong>r field v has the simplest representation<br />

i<br />

i j ∂υ<br />

j<br />

grad v = υ ,<br />

jgi ⊗ g = g<br />

j i<br />

⊗g (59.1)<br />

∂x<br />

Hence if we take the second covariant derivatives, then in the same coordinate system we have<br />

2 i<br />

grad ( grad v i j k ∂ υ<br />

j k<br />

) = υ , g jk i<br />

⊗ g ⊗ g =<br />

j k i<br />

⊗ ⊗<br />

∂x<br />

∂x<br />

g g g (59.2)<br />

In particular, the second covariant derivatives satisfy the same symmetry condition as that <strong>of</strong> the<br />

ordinary partial derivatives:<br />

i i<br />

υ , = υ ,<br />

(59.3)<br />

jk<br />

kj<br />

Note that the pro<strong>of</strong> <strong>of</strong> (59.3) depends crucially on the existence <strong>of</strong> a Cartesian coordinate<br />

system relative <strong>to</strong> which the Chris<strong>to</strong>ffel symbols <strong>of</strong> the Euclidean parallelism vanish identically.<br />

For the hypersurface S in general, the geodesic coordinate system at a reference point is the<br />

closest counterpart <strong>of</strong> a Cartesian system. However, in a geodesic coordinate system the surface<br />

Chris<strong>to</strong>ffel symbols vanish at the reference point only. As a result the surface covariant<br />

derivative at the reference point x<br />

0<br />

still has a simple representation like (59.1)<br />

Γ<br />

Δ<br />

( ) = υ ( ) ( ) ⊗ ( )<br />

grad v x , x e x e x<br />

0 Δ 0 Γ 0 0<br />

Γ<br />

∂υ<br />

= e<br />

Δ Γ<br />

x ⊗e x<br />

∂z<br />

x0<br />

Δ<br />

( ) ( )<br />

0 0<br />

(59.4)<br />

[cf. (57.40)]. But generally, in the same coordinate system, the representation (59.4)(59.4) does<br />

not hold any neighboring point <strong>of</strong> x<br />

0<br />

. This situation has been explained in detail in Section 57.


444 Chap. 11 • HYPERSURFACES<br />

In particular, there is no counterpart for the representation (59.2) 2<br />

on S . Indeed the surface<br />

second covariant derivatives generally fail <strong>to</strong> satisfy the symmetry condition (59.3) valid for the<br />

spatial covariant derivatives.<br />

{ } Γ<br />

To see this fact we choose an arbitrary surface coordinate system ( y Γ<br />

) with natural basis<br />

h <strong>and</strong> { h<br />

Γ<br />

}<br />

vec<strong>to</strong>r field<br />

on S as before. From (56.10) the surface covariant derivative <strong>of</strong> a tangential<br />

v = υ Γ h<br />

Γ<br />

(59.5)<br />

is given by<br />

Γ<br />

υ ,<br />

Δ<br />

Γ<br />

∂υ<br />

Σ⎧<br />

Γ ⎫<br />

= + υ<br />

Δ ⎨ ⎬<br />

∂y<br />

⎩ΣΔ⎭<br />

(59.6)<br />

Applying the same formula again, we obtain<br />

Γ<br />

Φ<br />

Γ ∂ ⎛∂υ<br />

Σ⎧ Γ ⎫⎞ ⎛∂υ<br />

Σ⎧Φ ⎫⎞⎧ Γ ⎫<br />

υ ,<br />

ΔΩ<br />

= v<br />

υ<br />

Ω ⎜ +<br />

Δ ⎨ ⎬⎟+ ⎜ +<br />

Δ ⎨ ⎬⎟⎨ ⎬<br />

∂y ⎝ ∂y ⎩ΣΔ⎭⎠ ⎝ ∂y<br />

⎩ΣΔ⎭⎠⎩ΦΩ⎭<br />

Γ<br />

⎛∂υ<br />

Σ ⎧ Γ ⎫⎞⎧ Ψ ⎫<br />

− ⎜ + υ<br />

Ψ ⎨ ⎬⎟⎨ ⎬<br />

⎝∂y<br />

⎩ΣΨ⎭⎠⎩ΔΩ⎭<br />

2 Γ Σ Φ Γ<br />

∂ υ ∂υ ⎧ Γ ⎫ ∂υ ⎧ Γ ⎫ ∂υ<br />

⎧ Ψ ⎫<br />

= +<br />

Δ Ω Ω ⎨ ⎬+ Δ ⎨ ⎬−<br />

Ψ ⎨ ⎬<br />

∂y ∂y ∂y ⎩ΣΔ⎭ ∂y ⎩ΦΩ⎭ ∂y<br />

⎩ΔΩ⎭<br />

Σ<br />

⎛⎧Φ ⎫⎧ Γ ⎫ ⎧ Γ ⎫⎧ Ψ ⎫ ∂ ⎧ Γ ⎫⎞<br />

+ υ ⎜⎨ ⎬⎨ ⎬−⎨ ⎬⎨ ⎬+ ⎩ΣΔ⎭⎩ΦΩ⎭ ⎩ΣΨ⎭⎩ΔΩ<br />

y Ω ⎨ ⎬⎟<br />

⎝<br />

⎭ ∂ ⎩ΣΔ⎭⎠<br />

(59.7)<br />

In particular, even if the surface coordinate system reduces <strong>to</strong> a geodesic coordinate system<br />

y Γ at a reference point x<br />

0<br />

, (59.7) can only be simplified <strong>to</strong><br />

( )<br />

υ<br />

∂ υ<br />

∂ ⎧ Γ ⎫<br />

x x (59.8)<br />

∂ ∂ ∂ ⎩ ⎭<br />

2 Γ<br />

Γ<br />

Σ<br />

,<br />

ΔΩ ( 0) = + υ<br />

Δ Ω<br />

( 0)<br />

Ω ⎨ ⎬<br />

z z X0 z ΣΔ X0<br />

which contains a term in addition <strong>to</strong> the spatial case (59.2)) 2 . Since that additional term<br />

Δ,<br />

Ω , we have shown that the surface second<br />

generally need not be symmetric in the pair ( )<br />

covariant derivatives do not necessarily obey the symmetry condition (59.3).


Sec. 59 • Surface Curvature, II 445<br />

rule:<br />

Γ<br />

Γ<br />

From (59.7) if we subtract υ, from υ,<br />

ΩΔ ΔΩ<br />

then the result is the following commutation<br />

Γ Γ Σ Γ<br />

υ , − υ , =− υ R<br />

(59.9)<br />

ΔΩ ΩΔ ΣΔΩ<br />

where<br />

R<br />

∂ ⎧ Γ ⎫ ∂ ⎧ Γ⎫ ⎧ Φ ⎫⎧ Γ ⎫ ⎧Φ ⎫ ⎧ Γ ⎫<br />

≡ ⎨ ⎬− ⎨ ⎬+ ⎨ ⎬⎨ ⎬−<br />

⎨ ⎬ ⎨ ⎬<br />

∂y<br />

⎩ΣΩ⎭ ∂y<br />

⎩ΣΔ⎭ ⎩ΣΩ⎭⎩ΦΔ⎭ ⎩ΣΔ⎭ ⎩ΦΩ⎭<br />

Γ<br />

ΣΔΩ Δ Ω<br />

(59.10)<br />

Since the commutation rule is valid for all tangential vec<strong>to</strong>r fields v , (59.9) implies that under a<br />

change <strong>of</strong> surface coordinate system the fields R Γ ΣΔΩ<br />

satisfy the transformation rule for the<br />

components <strong>of</strong> a fourth-order tangential tensor. Thus we define<br />

R Γ Σ Δ Ω<br />

ΣΔΩhΓ<br />

h h h<br />

R ≡ ⊗ ⊗ ⊗<br />

(59.11)<br />

<strong>and</strong> we call the tensor R the Riemann-Chris<strong>to</strong>ffel tensor <strong>of</strong> L . Notice that R depends only on<br />

the surface metrica a , since its components are determined completely by the surface Chris<strong>to</strong>ffel<br />

symbols. In particular, in a geodesic coordinates system ( z Γ ) at x<br />

0<br />

(59.10) simplifies <strong>to</strong><br />

R<br />

∂ ⎧ Γ ⎫ ∂ ⎧ Γ⎫<br />

x = −<br />

(59.12)<br />

⎩ ⎭ ⎩ ⎭<br />

( )<br />

Γ<br />

ΣΔΩ 0 Δ ⎨ ⎬ Ω ⎨ ⎬<br />

∂z<br />

ΣΩ X0 ∂z<br />

ΣΔ X0<br />

In a certain sense the Riemann-Chris<strong>to</strong>ffel tensor R characterizes locally the departure <strong>of</strong><br />

the Levi-Civita parallelism from a Euclidean one. We have the following result.<br />

Theorem 59.1. The Riemann-Chris<strong>to</strong>ffel tensor R vanishes identically on a neighborhood <strong>of</strong> a<br />

point x ∈S<br />

if <strong>and</strong> only if there exists a surface coordinate system 0 ( z Γ ) covering x<br />

0<br />

relativwe<br />

⎧ Γ ⎫<br />

<strong>to</strong> which the surface Chris<strong>to</strong>ffel symbols ⎨ ⎬<br />

⎩ΔΣ⎭ vanishing identically on a neighborhood <strong>of</strong> x<br />

0<br />

.<br />

In view <strong>of</strong> the representation (59.10), the sufficiency part <strong>of</strong> the preceding theorem is<br />

obvious. To prove the necessity part, we observe first the following lemma.<br />

Lemma. The Riemann-Chris<strong>to</strong>ffel tensor R vanishes identically near x<br />

0<br />

if <strong>and</strong> only if the<br />

system <strong>of</strong> first-order partial differential equations<br />

Γ<br />

Γ ∂υ<br />

Ω⎧<br />

Γ ⎫<br />

0 = υ ,<br />

Δ<br />

= + υ , 1, N 1<br />

Δ ⎨ ⎬ Γ= … −<br />

(59.13)<br />

∂y<br />

⎩ΩΔ⎭


446 Chap. 11 • HYPERSURFACES<br />

can be integrated near x<br />

0<br />

for each prescribed initial value<br />

υ Γ<br />

( )<br />

Pro<strong>of</strong>. Necessity. Let { , 1, , N 1}<br />

υ Γ<br />

x , 1, N 1<br />

0<br />

=<br />

0<br />

Γ= … −<br />

(59.14)<br />

v Σ= −<br />

Σ<br />

… be linearly independent <strong>and</strong> satisfy (59.13). Then we<br />

can obtain the surface Chris<strong>to</strong>ffel symbols from<br />

<strong>and</strong><br />

Γ<br />

∂υ<br />

Ω ⎧ Γ ⎫<br />

Σ<br />

0 = + υ<br />

Δ Σ ⎨ ⎬<br />

∂y<br />

⎩ΩΔ⎭<br />

⎧ Γ ⎫ Σ ∂<br />

⎨ ⎬=−uΩ<br />

⎩ΩΔ⎭<br />

∂y<br />

υ Γ<br />

Σ<br />

Δ<br />

(59.15)<br />

(59.16)<br />

where ⎡<br />

⎣u Σ Ω<br />

⎤<br />

⎦<br />

directly<br />

denotes the inverse matrix <strong>of</strong> ⎡<br />

⎣ υ Ω Σ<br />

⎤<br />

⎦ .<br />

Substituting (59.16) in<strong>to</strong> (59.10), we obtain<br />

R Γ = ΣΔΩ<br />

0<br />

(59.17)<br />

Sufficiency. It follows from the Frobenius theorem that the conditions <strong>of</strong> integrability for<br />

the system <strong>of</strong> first-order partial differential equations<br />

Γ<br />

∂υ<br />

Ω ⎧ Γ ⎫<br />

=−υ<br />

Δ ⎨ ⎬<br />

∂y<br />

⎩ΩΔ⎭<br />

(59.18)<br />

are<br />

∂<br />

∂y<br />

Σ<br />

⎛<br />

Ω⎧ Γ ⎫⎞ ∂ ⎛<br />

Ω⎧ Γ ⎫⎞<br />

⎜υ<br />

⎨ ⎬⎟− υ = 0, Γ= 1, N −1<br />

Δ ⎜ ⎨ ⎬⎟<br />

… (59.19)<br />

⎝ ⎩ΩΔ⎭⎠ ∂y<br />

⎝ ⎩ΩΣ⎭⎠<br />

If we exp<strong>and</strong> the partial derivatives in (59.19) <strong>and</strong> use (59.18), then (59.17) follows as a result <strong>of</strong><br />

(59.10). Thus the lemma is proved.<br />

Now we return <strong>to</strong> the pro<strong>of</strong> <strong>of</strong> the necessity part <strong>of</strong> the theorem. From (59.16) <strong>and</strong> the<br />

v obeys the rule<br />

condition (56.4) we see that the basis { } Σ<br />

Γ<br />

Γ<br />

∂υΣ<br />

Δ ∂υΩ<br />

Δ<br />

υΩ<br />

− υΣ<br />

= 0<br />

Δ<br />

Δ<br />

∂y<br />

∂y<br />

(59.20)


Sec. 59 • Surface Curvature, II 447<br />

These equations are the coordinate forms <strong>of</strong><br />

[ ]<br />

As a result there exists a coordinate system ( )<br />

particular, relative <strong>to</strong> ( z Γ<br />

) (59.16) reduces <strong>to</strong><br />

v , v = 0 , ΣΩ= Σ Ω<br />

, 1, … , N − 1<br />

(59.21)<br />

z Γ whose natural basis coincides with { v }<br />

Σ<br />

. In<br />

Γ<br />

⎧ Γ ⎫ Ω ∂δΣ<br />

⎨ ⎬= − δΣ<br />

= 0<br />

Δ<br />

⎩ΩΔ⎭<br />

∂z<br />

(59.22)<br />

Thus the theorem is proved.<br />

It should be noted that tangential vec<strong>to</strong>r fields satisfying (59.13) are parallel fields<br />

relative <strong>to</strong> the Levi-Civita parallelism, since from (56.24) we have the representation<br />

Γ<br />

( ( )) / = ,<br />

Δ<br />

Dv λ t Dt υ λ h<br />

(59.23)<br />

Δ<br />

Γ<br />

along any curve λ . Consequently, the Levi-Civita parallelism becomes locally pathindependent.<br />

For definiteness, we call this a locally Euclidean parallelism or a flat parallelism.<br />

Then the preceding theorem asserts that the Levi-Civita parallelism on S is locally Euclidean if<br />

<strong>and</strong> only if the Riemann-Chris<strong>to</strong>ffel tensor R based on the surface metric vanishes.<br />

The commutation rule (59.9) is valid for tangential vec<strong>to</strong>r fields only. However, we can<br />

easily generalize that rule <strong>to</strong> the following Ricci identities:<br />

A<br />

, −A<br />

,<br />

Γ1... Γr<br />

Γ1...<br />

Γr<br />

Δ1... Δs<br />

ΣΩ Δ1...<br />

Δs<br />

ΩΣ<br />

=−A<br />

R<br />

ΦΓ2...<br />

Γr<br />

Γ1<br />

Δ1...<br />

Δs<br />

ΦΣΩ<br />

−− A R + A R<br />

+ +<br />

A<br />

Γ1... Γr−1Φ Γr Γ1...<br />

Γr<br />

Ψ<br />

Δ1... Δs<br />

ΦΣΩ ΨΔ2...<br />

Δs<br />

Δ1ΣΩ<br />

R<br />

Γ1...<br />

Γr<br />

Ψ<br />

Δ1...<br />

Δs−1Ψ ΔsΣΩ<br />

(59.24)<br />

As a result, (59.17) is also the integrability condition <strong>of</strong> the system<br />

1...<br />

r<br />

A Γ Γ 1... , = Δ Δ<br />

0<br />

(59.25)<br />

s Σ<br />

for each prescribed initial value at any reference point<br />

0<br />

x ; further, a solution <strong>of</strong> (59.25)<br />

corresponds <strong>to</strong> a parallel tangential tensor field on the developable hyper surface S .


448 Chap. 11 • HYPERSURFACES<br />

In classical differential geometry a surface S is called developable if its Riemann-<br />

Chris<strong>to</strong>ffel curvature tensore vanishes identically.


Sec. 60 • Surface Curvature, III 449<br />

Section 60. Surface Curvature, III. The Equations <strong>of</strong> Gauss <strong>and</strong> Codazzi<br />

In the preceding two sections we have considered the curvature <strong>of</strong> S from both the<br />

extrinsic point <strong>of</strong> view <strong>and</strong> the intrinsic point <strong>of</strong> view. In this section we shall unite the results <strong>of</strong><br />

these two approaches.<br />

curve:<br />

Our starting point is the general representation (58.25) applied <strong>to</strong> the y Δ -coordinate<br />

∂v<br />

Δ<br />

∂y<br />

Dv<br />

= b υ Γ<br />

ΓΔ<br />

n +<br />

(60.1)<br />

Δ<br />

Dy<br />

for an arbitrary tangential vec<strong>to</strong>r field v . This representation gives the natural decomposition <strong>of</strong><br />

the spatial vec<strong>to</strong>r field ∂v / ∂y Δ on S in<strong>to</strong> a normal projection b υ Γ n <strong>and</strong> a tangential<br />

projection Dv / Dy Δ . Applying the spatial covariant derivative ∂ / ∂ y Σ<br />

Σ<br />

along the y − coordinate<br />

curve <strong>to</strong> (60.1)(60.1), we obtain<br />

ΓΔ<br />

∂ ⎛ ∂v<br />

⎞ ∂ Γ ∂ ⎛ Dv<br />

⎞<br />

( b υ<br />

Σ )<br />

y<br />

⎜ Δ<br />

y<br />

⎟= ΓΔ<br />

n +<br />

Σ Σ Δ<br />

∂ y y<br />

⎜<br />

Dy<br />

⎟<br />

⎝∂ ⎠ ∂ ∂ ⎝ ⎠<br />

∂ Γ<br />

Γ D ⎛ Dv<br />

⎞<br />

= ( bΓΔυ<br />

n) + bΓΣυ<br />

Σ , Δn+ Σ Δ<br />

∂y Dy<br />

⎜<br />

Dy<br />

⎟<br />

⎝ ⎠<br />

(60.2)<br />

where we have applied (60.1) in (60.2) 2 <strong>to</strong> the tangential vec<strong>to</strong>r field Dv / Dy Δ . Now, since the<br />

spatial parallelism is Euclidean, the lefth<strong>and</strong> side <strong>of</strong> (60.2) is symmetric in the pair ( ΣΔ , ),<br />

namely<br />

∂ ⎛ ∂v<br />

⎞ ∂ ⎛ ∂v<br />

⎞<br />

=<br />

Σ Δ Δ Σ<br />

∂y ⎜<br />

∂y ⎟<br />

∂y ⎜<br />

∂y<br />

⎟<br />

⎝ ⎠ ⎝ ⎠<br />

(60.3)<br />

Hence from (60.2) we have<br />

D ⎛ Dv<br />

⎞ D ⎛ Dv<br />

⎞ ∂<br />

Δ Σ Σ Δ Σ<br />

Dy<br />

⎜ − = −<br />

Dy<br />

⎟<br />

Dy<br />

⎜<br />

Dy<br />

⎟<br />

⎝ ⎠ ⎝ ⎠ ∂y<br />

Γ<br />

Γ<br />

( bΓΔυ<br />

n) ( bΓΣυ<br />

n)<br />

Γ<br />

Γ<br />

( bΓΣυ<br />

, Δ<br />

bΓΔυ<br />

,<br />

Σ )<br />

+ −<br />

n<br />

(60.4)<br />

This is the basic formula from which we can extract the relation between the Riemann-<br />

Chris<strong>to</strong>ffel tensor R <strong>and</strong> the second fundamental form B .


450 Chap. 11 • HYPERSURFACES<br />

form<br />

Specifically, the left-h<strong>and</strong> side <strong>of</strong> (60.4) is a tangaential vec<strong>to</strong>r having the component<br />

D ⎛ Dv<br />

⎞ D ⎛ Dv<br />

⎞<br />

Δ Σ Σ Δ<br />

Dy<br />

⎜ − = −<br />

Dy<br />

⎟<br />

Dy<br />

⎜<br />

Dy<br />

⎟<br />

⎝ ⎠ ⎝ ⎠<br />

Γ Γ<br />

( υ<br />

,<br />

υ<br />

, )<br />

Ω<br />

=−υ<br />

R<br />

ΔΣ ΣΔ Γ<br />

Γ<br />

ΩΣΔ<br />

h<br />

Γ<br />

h<br />

(60.5)<br />

as required by (59.9), while the right-h<strong>and</strong> side is a vec<strong>to</strong>r having the normal projection<br />

⎛∂<br />

⎜<br />

⎝<br />

Γ<br />

Γ<br />

( b υ ) ∂( b υ )<br />

ΓΔ ΓΣ Γ Γ<br />

− + bΓΣυ<br />

, Δ<br />

− bΓΔυ<br />

Σ<br />

Δ<br />

, Σ<br />

∂y<br />

∂y<br />

⎞<br />

n (60.6)<br />

⎟<br />

⎠<br />

<strong>and</strong> the tangential projection<br />

( b υ<br />

Γ b Φ b υ<br />

Γ b<br />

Φ<br />

)<br />

− + h (60.7)<br />

ΓΔ Σ ΓΣ Δ Φ<br />

where we have used Weingarten’s formula (58.7) <strong>to</strong> compute the covariant derivatives<br />

Σ<br />

Δ<br />

∂n/ ∂y<br />

<strong>and</strong> ∂n / ∂y<br />

. Consequently, (60.5) implies that<br />

∂<br />

Γ<br />

Γ<br />

( b υ ) ∂( b υ )<br />

ΓΔ ΓΣ Γ Γ<br />

− + bΓΣυ<br />

, Δ<br />

− bΓΔυ<br />

, Σ<br />

= 0<br />

Σ<br />

Δ<br />

∂y<br />

∂y<br />

(60.8)<br />

<strong>and</strong> that<br />

Γ Φ Γ Φ Γ Φ<br />

υ R = b υ b − b υ b<br />

(60.9)<br />

ΓΣΔ ΓΔ Σ ΓΣ Δ<br />

for all tangential vec<strong>to</strong>r fields v .<br />

If we choose v = h , then (60.9) becomes<br />

Γ<br />

Φ Φ Φ<br />

R = b b − b b<br />

(60.10)<br />

ΓΣΔ ΓΔ Σ ΓΣ Δ<br />

or, equivalently,<br />

RΦΓΣΔ = bΓΔ bΦΣ − bΓΣbΦ Δ<br />

(60.11)


Sec. 60 • Surface Curvature, III 451<br />

Thus the Riemann-Chris<strong>to</strong>ffel tensor R is completely determined by the second<br />

fundamental form. The important result (60.11) is called the equations <strong>of</strong> Gauss in classical<br />

differential geometry. A similar choice for v . in (60.8) yields<br />

∂b<br />

∂y<br />

ΓΔ<br />

Σ<br />

⎧ Φ⎫ ∂b<br />

⎧ Φ⎫<br />

ΓΣ<br />

−bΦΔ<br />

⎨ ⎬− + bΦΣ<br />

= 0<br />

Δ ⎨ ⎬<br />

⎩ΓΣ⎭ ∂y<br />

⎩ΓΔ⎭<br />

(60.12)<br />

By use <strong>of</strong> the symmetry condition (56.4) we can rewrite (60.12) in the more elegant form<br />

b<br />

ΓΔ, Σ<br />

bΓΣ, Δ<br />

0<br />

− = (60.13)<br />

which are the equations <strong>of</strong> Codazzi.<br />

The importance <strong>of</strong> the equations <strong>of</strong> Gauss <strong>and</strong> Codazzi lies not only in their uniting the<br />

second fundamental form with the Riemann Chris<strong>to</strong>ffel tensor for any hypersurfaces S in E ,<br />

but also in their being the conditions <strong>of</strong> integrability as asserted by the following theorem.<br />

Theorem 60.1. Suppose that a ΓΔ<br />

<strong>and</strong> b ΓΔ<br />

are any presceibed smooth functions <strong>of</strong> ( y Ω<br />

)<br />

that [ a ΓΔ ] is positive-definite <strong>and</strong> symmetric, [ b ΓΔ ] is symmetric, <strong>and</strong> <strong>to</strong>gether [ a ΓΔ ] <strong>and</strong> [ ]<br />

b ΓΔ<br />

satisfy the equations <strong>of</strong> Gauss <strong>and</strong> Codazzi. Then locally there exists a hyper surface S with<br />

representation<br />

( )<br />

i i<br />

x x y Ω<br />

such<br />

= (60.14)<br />

on which the prescribed<br />

a<br />

ΓΔ<br />

<strong>and</strong> b<br />

ΓΔ<br />

are the first <strong>and</strong> second fundamental forms.<br />

i<br />

Pro<strong>of</strong>. For simplicity we choose a rectangular Cartesian coordinate system x in E . Then in<br />

component form (58.21) <strong>and</strong> (58.7) are represented by<br />

Δ ⎧ Σ ⎫<br />

Γ Δ<br />

∂h / ∂ y = b n + ⎨ ⎬h , ∂n / ∂ y =−b h<br />

⎩ΓΔ⎭<br />

i i i i i<br />

Γ ΓΔ Σ Γ Δ<br />

(60.15)<br />

We claim that this system can be integrated <strong>and</strong> the solution preserves the algebraic conditions<br />

i j i j i j<br />

δ hh = a , δ hn = 0, δ nn = 1<br />

(60.16)<br />

ij Γ Δ ΓΔ ij Γ<br />

ij<br />

In particular, if (60.16) are imposed at any one reference point, then they hold identically on a<br />

neighborhood <strong>of</strong> the reference point.


452 Chap. 11 • HYPERSURFACES<br />

The fact that the solution <strong>of</strong> (60.15) preserves the conditions (60.16) is more or less<br />

obvious. We put<br />

Then initially at some point ( 1 2<br />

0,<br />

0)<br />

i j i j i j<br />

f ≡δ h h −a , f ≡δ h n , f ≡ δ nn − 1<br />

(60.17)<br />

ΓΔ ij Γ Δ ΓΔ Γ ij Γ<br />

ij<br />

y y we have<br />

1 2 1 2 1 2<br />

( 0 0) Γ ( 0 0) ( 0 0)<br />

f y , y = f y , y = f y , y = 0<br />

(60.18)<br />

ΓΔ<br />

Now from (60.15) <strong>and</strong> (60.17) we can verify easily that<br />

∂fΓΔ<br />

∂f Δ ⎧ Δ ⎫<br />

Γ<br />

∂f<br />

Δ<br />

= b f + b f , =− b f + f b f, 2b f<br />

Σ Σ ⎨ ⎬ + =−<br />

Σ<br />

∂y ∂y ⎩ΓΣ⎭<br />

∂y<br />

ΓΣ Δ ΔΣ Γ Σ ΓΔ Δ ΓΣ Σ Δ<br />

(60.19)<br />

along the coordinate curve <strong>of</strong> any y Σ . From (60.18) <strong>and</strong> (60.19) we see that fΓΔ, fΓ<br />

, <strong>and</strong> f<br />

must vanish identically.<br />

Now the system (60.15) is integrable, since by use <strong>of</strong> the equations <strong>of</strong> Gauss <strong>and</strong> Codazzi we<br />

have<br />

<strong>and</strong><br />

i<br />

i<br />

∂ ⎛∂hΓ<br />

⎞ ∂ ⎛∂hΓ<br />

⎞ Ω Ω Ω<br />

− = ( R − b b + b b ) h<br />

Σ Δ Δ Σ<br />

∂y ⎜<br />

∂y ⎟<br />

∂y ⎜<br />

∂y<br />

⎟<br />

⎝ ⎠ ⎝ ⎠<br />

( ΓΣ Δ ΓΔ Σ )<br />

+ b − b n =<br />

i<br />

, ,<br />

0<br />

i<br />

i<br />

⎛ ⎞ ⎛ ⎞<br />

i<br />

ΓΣΔ ΓΣ Δ ΓΔ Σ Ω<br />

∂ ∂n<br />

∂ ∂n<br />

Ω Ω<br />

( b b ) h<br />

Σ ⎜ Γ ⎟− Γ ⎜ Σ ⎟= − =<br />

∂y ⎝∂y ⎠ ∂y ⎝∂y<br />

⎠<br />

i<br />

ΣΓ , ΓΣ , Ω<br />

0<br />

(60.20)<br />

(60.21)<br />

i Ω<br />

i Ω<br />

Hence locally there exist functions h ( y ) <strong>and</strong> n ( y )<br />

Γ<br />

which verify (60.15) <strong>and</strong> (60.16).<br />

Next we set up the system <strong>of</strong> first-order partial differential equations<br />

Ω<br />

( )<br />

i Δ i<br />

∂x / ∂ y = h y<br />

(60.22)<br />

Δ<br />

where the right-h<strong>and</strong> side is the solution <strong>of</strong> (60.15) <strong>and</strong> (60.16). Then (60.22) is also integrable,<br />

since we have


Sec. 60 • Surface Curvature, III 453<br />

i i i<br />

∂ ⎛ ∂x ⎞ ∂hΔ<br />

i ⎧ Ω⎫ i ∂ ⎛ ∂x<br />

⎞<br />

bΔΣn<br />

h<br />

Σ ⎜ Δ ⎟= = +<br />

Σ ⎨ ⎬ Ω<br />

=<br />

Δ ⎜ Σ ⎟<br />

∂y ⎝∂y ⎠ ∂y ⎩ΔΣ⎭<br />

∂y ⎝∂y<br />

⎠<br />

(60.23)<br />

Consequently the solution (60.14) exists; further, from (60.22), (60.16), <strong>and</strong> (60.15) the first <strong>and</strong><br />

the second fundamental forms on the hypersurface defined by (60.14) are precisely the<br />

prescribed fields a <strong>and</strong> b , respectively. Thus the theorem is proved.<br />

ΓΔ ΓΔ<br />

It should be noted that, since the algebraic conditions (60.16) are preserved by the<br />

solution <strong>of</strong> (60.15), any two solutions ( h i , j ) <strong>and</strong> ( i ,<br />

j<br />

Γ<br />

n hΓ n ) <strong>of</strong> (60.15) can differ by at most a<br />

transformation <strong>of</strong> the form<br />

i i j i i j<br />

h = Q h , n = Q n<br />

(60.24)<br />

Γ<br />

j<br />

Γ<br />

i<br />

where ⎡<br />

⎣Q j<br />

⎤<br />

⎦ is a constant orthogonal matrix. This transformation corresponds <strong>to</strong> a change <strong>of</strong><br />

rectangular Cartesian coordinate system on E used in the representation (60.14). In this sense<br />

the fundamental forms aΓΔ<br />

<strong>and</strong> bΓΔ , <strong>to</strong>gether with the equations <strong>of</strong> Gauss <strong>and</strong>Codazzi, determine<br />

the hyper surface S locally <strong>to</strong> within a rigid displacement in E .<br />

While the equations <strong>of</strong> Gauss show that the Riemann Chris<strong>to</strong>ffel tensor R <strong>of</strong> S is<br />

completely determined by the second fundamental form B , the converse is generally not true.<br />

Thus mathematically the extrinsic characterization <strong>of</strong> curvature is much stronger than the<br />

intrinsic one, as it should be. In particular, if B vanishes, then S reduces <strong>to</strong> a hyperplane which<br />

is trivially developable so that R vanishes. On the other h<strong>and</strong>, if R vanishes, then S can be<br />

developed in<strong>to</strong> a hyperplane, but generally B does not vanish, so S need not itself be a<br />

hyperplane. For example, in a three-dimensional Euclidean space a cylinder is developable, but<br />

it is not a plane.<br />

Exercise<br />

60.1 In the case where N = 3 show that the only independent equations <strong>of</strong> Codazzi are<br />

b<br />

ΔΔ, Γ<br />

− bΔΓ, Δ<br />

= 0<br />

j<br />

for Δ≠Γ <strong>and</strong> where the summation convention has been ab<strong>and</strong>oned. Also show that the<br />

only independent equation <strong>of</strong> Gauss is the single equation<br />

R = b b − b = B<br />

2<br />

1212 11 22 12<br />

det


454 Chap. 11 • HYPERSURFACES<br />

Section 61. Surface Area, Minimal Surface<br />

Since we assume that S is oriented, the tangent plane S<br />

x<br />

<strong>of</strong> each point<br />

x ∈S is<br />

equipped with a volume tensor S. Relative <strong>to</strong> any positive surface coordinate system ( y Γ<br />

)<br />

natural basis { h Γ }, S can be represented by<br />

with<br />

S = ah<br />

∧ ∧h<br />

(61.1)<br />

1 N −1<br />

where<br />

a = det[ ]<br />

(61.2)<br />

a ΓΔ<br />

If U is a domain in S covered by ( y Γ ), then the surface area σ ( U ) is defined by<br />

σ<br />

1 N −1<br />

( U ) = ⋅⋅⋅ ady dy<br />

∫ ∫<br />

U<br />

(61.3)<br />

where the ( N −1)<br />

-fold integral is taken over the coordinates ( y Γ ) on the domain U . Since<br />

obeys the transformation rule<br />

a<br />

a<br />

Γ<br />

⎡ ∂y<br />

⎤<br />

= a det ⎢ Δ<br />

∂y<br />

⎥<br />

⎣ ⎦<br />

(61.4)<br />

relative <strong>to</strong> any positive coordinate systems ( y ) <strong>and</strong> ( y )<br />

Γ Γ , the value <strong>of</strong> ( )<br />

σ U is independent <strong>of</strong><br />

the choice <strong>of</strong> coordinates. Thus σ can be extended in<strong>to</strong> a positive measure on S in the sense <strong>of</strong><br />

measure theory.<br />

We shall consider the theory <strong>of</strong> integration relative <strong>to</strong> σ in detail in Chapter 13. Here we<br />

shall note only a particular result wich connects a property <strong>of</strong> the surface area <strong>to</strong> the mean<br />

curvature <strong>of</strong> S . Namely, we seek a geometric condition for S <strong>to</strong> be a minimal surface, which<br />

is defined by the condition that the surface area σ ( U ) be an extremum in the class <strong>of</strong> variations<br />

<strong>of</strong> hypersurfaces having the same boundary as U . The concept <strong>of</strong> minimal surface is similar <strong>to</strong><br />

that <strong>of</strong> a geodesic whichwe have explored in Section 57, except that here we are interested in the<br />

variation <strong>of</strong> the integral σ given by (61.3) instead <strong>of</strong> the integral s given by (57.5).<br />

As before, the geometric condition for a minimal surface follows from the Euler-<br />

Lagrange equation for (61.3), namely,


Sec. 61 • Surface Area, Minimal Surface 455<br />

∂ ⎛∂ a ⎞ ∂ a<br />

0<br />

Γ ⎜ i ⎟− =<br />

i<br />

∂y ⎝ ∂hΓ<br />

⎠ ∂x<br />

(61.5)<br />

where a is regarded as a function <strong>of</strong><br />

x<br />

i<br />

i<br />

<strong>and</strong> h Γ<br />

through the representation<br />

a<br />

∂x<br />

∂y<br />

∂x<br />

∂y<br />

i j<br />

i i<br />

ΓΔ<br />

= δij<br />

hΓ hΔ = δij<br />

Γ Δ<br />

(61.6)<br />

For simplicity we have chosen the spatial coordinates <strong>to</strong> be rectangular Cartesian, so that<br />

a ΓΔ<br />

does not depend explicitly on x i . From (61.6) the partial derivative <strong>of</strong> a with respect <strong>to</strong><br />

i<br />

h Γ<br />

is given by<br />

∂ a<br />

∂h<br />

i<br />

Γ<br />

= δ =<br />

δ<br />

ΓΔ i<br />

Γj<br />

aa<br />

ij<br />

hΔ<br />

a<br />

ij<br />

h<br />

(61.7)<br />

Substituting this formula in<strong>to</strong> (61.5), we see that the condition for S <strong>to</strong> be a minimal surface is<br />

Γj<br />

⎛⎧ Ω ⎫ Γj<br />

∂h<br />

⎞<br />

aδij<br />

⎜⎨<br />

⎬ h + = 0, i = 1, , N<br />

Γ ⎟<br />

… (61.8)<br />

⎝⎩ΩΓ⎭<br />

∂y<br />

⎠<br />

or, equivalently, in vec<strong>to</strong>r notation<br />

a<br />

Γ<br />

⎛⎧ Ω ⎫ Γ ∂h<br />

⎞<br />

⎜⎨<br />

⎬ h + =<br />

Γ ⎟ 0 (61.9)<br />

⎝⎩ΩΓ⎭<br />

∂y<br />

⎠<br />

In deriving (61.8) <strong>and</strong> (61.9), we have used the identity<br />

∂ a ⎧Ω<br />

⎫<br />

=<br />

y Γ ⎨ ⎬<br />

∂ ⎩ΩΓ⎭<br />

a<br />

(61.10)<br />

which we have noted in Exercise 56.10. Now from (58.20) we can rewrite (61.9) in the simple<br />

form<br />

ab Γ Γ<br />

n = 0<br />

(61.11)


456 Chap. 11 • HYPERSURFACES<br />

As a result, the geometric condition for a minimal surface is<br />

I tr b Γ B<br />

= B =<br />

Γ<br />

= 0<br />

(61.12)<br />

In general, the first invariant I B<br />

<strong>of</strong> the second fundamental form B is called the mean<br />

curvature <strong>of</strong> the hyper surface. Equation (61.12) then asserts that S is a minimal surface if <strong>and</strong><br />

only if its mean curvature vanishes.<br />

Since in general B is symmetric, it has the usual spectral decomposition<br />

N −1<br />

∑<br />

B= β c ⊗ c (61.13)<br />

Γ= 1<br />

Γ Γ Γ<br />

where the principal basis { c Γ } is orthonormal on the surface. In differential geometry the<br />

direction <strong>of</strong> ( ) Γ<br />

c x at each point x ∈S is called a principal direction <strong>and</strong> the corresponding<br />

proper number β Γ ( x ) is called a principal (normal) curvature at x . As usual, the principal<br />

(normal) curvatures are the extrema <strong>of</strong> the normal curvature<br />

n<br />

( , )<br />

κ = B s s (61.14)<br />

in all unit tangents s at any point x [cf. (58.29) 2 ]. The mean curvature I B<br />

, <strong>of</strong> course, is equal <strong>to</strong><br />

the sum <strong>of</strong> the principal curvatures, i.e.,<br />

N −1<br />

IB = ∑ βΓ<br />

(61.15)<br />

Γ= 1


Sec. 62 • Three-Dimensional Euclidean Manifold 457<br />

Section 62. Surfaces in a Three-Dimensional Euclidean Manifold<br />

In this section we shall apply the general theory <strong>of</strong> hypersurfaces <strong>to</strong> the special case <strong>of</strong> a<br />

two-dimensional surface imbedded in a three-dimensional Euclidean manifold. This special case<br />

is the commonest case in application, <strong>and</strong> it also provides a good example <strong>to</strong> illustrate the<br />

general results.<br />

As usual we can represent S by<br />

( )<br />

i i<br />

x x y Γ<br />

= (62.1)<br />

i<br />

where i ranges from 1 <strong>to</strong> 3 <strong>and</strong> Γ ranges from 1 <strong>to</strong> 2. We choose ( ) <strong>and</strong> ( )<br />

x y Γ <strong>to</strong> be positive<br />

spatial <strong>and</strong> surface coordinate systems, respectively. Then the tangent plane <strong>of</strong> S is spanned by<br />

the basis<br />

i<br />

∂x<br />

hΓ<br />

≡ g<br />

i, Γ= 1,2<br />

(62.2)<br />

Γ<br />

∂y<br />

<strong>and</strong> the unit normal n is given by<br />

n =<br />

h1×<br />

h2<br />

h × h<br />

1 2<br />

(62.3)<br />

by<br />

The first <strong>and</strong> the second fundamental forms a ΓΔ<br />

<strong>and</strong> b ΓΔ<br />

relative <strong>to</strong> ( y Γ<br />

)<br />

are defined<br />

a<br />

∂hΓ<br />

∂n<br />

= h ⋅ h . b = n⋅ = −h ⋅<br />

(62.4)<br />

Δ<br />

Δ<br />

∂y<br />

∂y<br />

ΓΔ Γ Δ ΓΔ Γ<br />

These forms satisfy the equations <strong>of</strong> Gauss <strong>and</strong> Codazzi:<br />

R = b b − b b , b − b = 0<br />

(62.5)<br />

ΦΓΣΔ ΓΔ ΦΣ ΓΣ Φ Δ ΓΔ, Σ ΓΣ,<br />

Δ<br />

Since b ΓΔ<br />

is symmetric, at each point x ∈S there exists a positive orthonormal basis<br />

{ cΓ<br />

( x )}<br />

relative <strong>to</strong> which ( )<br />

B x can be represented by the spectral form<br />

( ) = β ( ) ( ) ⊗ ( ) + β ( ) ( ) ⊗ ( )<br />

B x x c x c x x c x c x (62.6)<br />

1 1 1 2 2 2


458 Chap. 11 • HYPERSURFACES<br />

The proper numbers β1( ) <strong>and</strong> β2( )<br />

general ( )<br />

{ Γ }<br />

x x are the principal (normal) curvatures <strong>of</strong> S at x . In<br />

c x is not the natural basis <strong>of</strong> a surface coordinate system unless<br />

[ ]<br />

c , c = 0<br />

(62.7)<br />

1 2<br />

However, locally we can always choose a coordinate system ( z Γ<br />

)<br />

basis { h Γ } is parallel <strong>to</strong> { Γ ( )}<br />

in such a way that the natural<br />

c x . Naturally, such a coordinate system is called a principal<br />

coordinate system <strong>and</strong> its coordinate curves are called the lines <strong>of</strong> curvature. Relative <strong>to</strong> a<br />

principal coordinate system the components a ΓΔ<br />

<strong>and</strong> b ΓΔ<br />

satisfy<br />

b = a = 0 <strong>and</strong> h = a c , h = a c (62.8)<br />

12 12 1 11 1 2 22 2<br />

The principal invariants <strong>of</strong> B are<br />

I<br />

B<br />

= trB<br />

= β + β = a<br />

1 2<br />

ΓΔ<br />

Γ<br />

B<br />

= =<br />

1 2<br />

= ⎣ Δ ⎦<br />

b<br />

ΓΔ<br />

II det B ββ det ⎡b<br />

⎤<br />

(62.9)<br />

In the preceding section we have defined I B<br />

<strong>to</strong> be the mean curvature. Now II B<br />

is called the<br />

Gaussian curvature. Since<br />

we have also<br />

Γ ΓΩ<br />

b = a b<br />

(62.10)<br />

Δ<br />

ΔΩ<br />

[ bΔΩ]<br />

[ a ]<br />

det det[ bΔΩ]<br />

II<br />

B<br />

= =<br />

(62.11)<br />

det a<br />

ΔΩ<br />

where the numera<strong>to</strong>r on the right-h<strong>and</strong> side is given by the equation <strong>of</strong> Gauss:<br />

[ ]<br />

2<br />

det b b b b R<br />

= − = (62.12)<br />

ΔΩ 11 22 12 1212<br />

Notice that for the two-dimensional case the tensor R is completely determined by R , since<br />

1212<br />

from (62.5) 1 , or indirectly from (59.10), R ΦΓΣΔ<br />

vanishes when Φ =Γor when Σ=Δ .<br />

We call a point x ∈S elliptic, hyperbolic, or parabolic if the Gaussian curvature threre<br />

is positive, negative, or zero, respectively. These terms are suggested by the following geometric<br />

considerations: We choose a fixed reference point x ∈ 0<br />

S <strong>and</strong> define a surface coordinate


Sec. 62 • Three-Dimensional Euclidean Manifold 459<br />

system ( y Γ ) such that the coordinates <strong>of</strong> x<br />

0<br />

are ( )<br />

i.e.,<br />

a<br />

ΓΔ<br />

( 0,0)<br />

{ Γ }<br />

0,0 ) <strong>and</strong> the basis ( )<br />

h x is orthonormal,<br />

= δ<br />

(62.13)<br />

1 2<br />

In this coordinate system a point x ∈S having coordinate ( y , y ) near ( 0,0 ) is located at a<br />

1 2<br />

1 2<br />

d y , y from the tangenet plane d y , y given approximately by<br />

normal distance ( )<br />

x o<br />

ΓΔ<br />

S with ( )<br />

1 2<br />

( , y ) = n( x0) ⋅( x−x0)<br />

d y<br />

⎛<br />

1 Dh<br />

≅n( x0) ⋅ hΓ<br />

( x0)<br />

y +<br />

⎜<br />

⎝<br />

2 Dy<br />

1<br />

Γ Δ<br />

≅ bΓΔ<br />

( x0<br />

) y y<br />

2<br />

y y<br />

Γ Γ Γ Δ<br />

Δ<br />

x0<br />

⎞<br />

⎟<br />

⎠<br />

(62.14)<br />

where (62.14) 2,3 are valid <strong>to</strong> within an error <strong>of</strong> third or higher order in y Γ . From this estimate<br />

we see that the intersection <strong>of</strong> S with a plane parallel <strong>to</strong> S<br />

x o<br />

is represented approximately by<br />

the curve<br />

b<br />

( )<br />

Γ Δ<br />

x<br />

0<br />

y y const<br />

(62.15)<br />

ΓΔ<br />

=<br />

which is an ellipse, a hyperbola, or parallel lines when x<br />

0<br />

is an elliptic, hyperbolic, or parabolic<br />

c <strong>of</strong> B x coincides with the<br />

point, respectively. Further, in each case the principal basis { } ( )<br />

principal axes <strong>of</strong> the curves in the sense <strong>of</strong> conic sections in analytical geometry. The estimate<br />

x induced by the<br />

i<br />

(62.14) means also that in the rectangular Cartesian coordinate system ( )<br />

orthonormal basis { 1( 0) ,<br />

2( 0)<br />

, },<br />

equations<br />

h x h x n the surface S can be represented locally by the<br />

1 1 2 2 3<br />

Γ Δ<br />

x y , x y , x bΓΔ<br />

0<br />

y y<br />

Γ<br />

1<br />

= = ≅ ( x )<br />

(62.16)<br />

2<br />

As usual we can define the notion <strong>of</strong> conjugate directions relative <strong>to</strong> the symmetric<br />

B x We say that the tangential vec<strong>to</strong>rs uv , ∈S x<br />

are conjugate at x<br />

0<br />

0<br />

if<br />

bilinear form <strong>of</strong> ( 0 ) .<br />

( ) ( ) ( )<br />

B u, v = u⋅ Bv = v⋅ Bu = 0<br />

(62.17)<br />

0


460 Chap. 11 • HYPERSURFACES<br />

For example, the principal basis vec<strong>to</strong>rs<br />

1( 0) <strong>and</strong><br />

2( 0)<br />

<strong>of</strong> B form a diagonal matrix relative <strong>to</strong> { c Γ }. Geometrically, conjugate directions may be<br />

explained in the following way: We choose any curve λ ( t)<br />

∈S such that<br />

c x c x are conjugate since the components<br />

( 0 ) = u∈ x o<br />

λ S (62.18)<br />

Then the conjugate direction v<strong>of</strong><br />

u corresponds <strong>to</strong> the limit <strong>of</strong> the intersection <strong>of</strong> S λ ()<br />

with S<br />

t x 0<br />

as t tends <strong>to</strong> zero. We leave the pro<strong>of</strong> <strong>of</strong> this geometric interpretation for conjugate directions as<br />

an exercise.<br />

A direction represented by a self-conjugate vec<strong>to</strong>r v is called an asymp<strong>to</strong>tic direction. In<br />

this case v satisfies the equation<br />

( ) ( )<br />

B v, v = v⋅ Bv = 0<br />

(62.19)<br />

Clearly, asymp<strong>to</strong>tic directions exit at a point x<br />

o<br />

if <strong>and</strong> only if x<br />

o<br />

is hyperbolic or parabolic. In<br />

the former case the asympo<strong>to</strong>tic directions are the same as those for the hyperbola given by<br />

(62.15), while in the latter case the asymp<strong>to</strong>tic direction is unique <strong>and</strong> coincides with the<br />

direction <strong>of</strong> the parallel lines given by (62.15).<br />

If every point <strong>of</strong> S is hyperbolic, then the asymp<strong>to</strong>tic lines form a coordinate net, <strong>and</strong><br />

we can define an asymp<strong>to</strong>tic coordinate system. Relative <strong>to</strong> such a coordinate system the<br />

components b ΓΔ<br />

satisfy the condition<br />

b<br />

= b = (62.20)<br />

11 22<br />

0<br />

<strong>and</strong> the Gaussian curvature is given by<br />

II<br />

B<br />

=−b / a<br />

(62.21)<br />

2<br />

12<br />

A minimal surface which does not reduce <strong>to</strong> a plane is necessarily hyperbolic at every point,<br />

since when β1+ β2 = 0 we must have β1β2 < 0 unless β1 = β2<br />

= 0.<br />

From (62.12), S is parabolic at every point if <strong>and</strong> only if it is developable. In this case<br />

the asymp<strong>to</strong>tic lines are straight lines. In fact, it can be shown that there are only three kinds <strong>of</strong><br />

developable surfaces, namely cylinders, cones, <strong>and</strong> tangent developables. Their asymp<strong>to</strong>tic lines<br />

are simply their genera<strong>to</strong>rs. Of course, these genera<strong>to</strong>rs are also lines <strong>of</strong> curvature, since they are<br />

in the principal direction corresponding <strong>to</strong> the zero principal curvature.


Sec. 62 • Three-Dimensional Euclidean Manifold 461<br />

Exercises<br />

62.1 Show that<br />

( )<br />

a = n⋅ h × h<br />

1 2<br />

62.2 Show that<br />

b<br />

⎛<br />

a ∂y ∂y ⎜<br />

⎝∂y ∂y<br />

2<br />

1 ∂ x ∂x ∂x<br />

ΓΔ<br />

= ⋅ ×<br />

Δ Γ 1 2<br />

⎞<br />

⎟<br />

⎠<br />

62.3 Compute the principal curvatures for the surfaces defined in Exercises 55.5 <strong>and</strong> 55.7.


________________________________________________________________<br />

Chapter 12<br />

ELEMENTS OF CLASSICAL CONTINUOUS GROUPS<br />

In this chapter we consider the structure <strong>of</strong> various classical continuous groups which are formed<br />

by linear transformations <strong>of</strong> an inner product space V . Since the space <strong>of</strong> all linear transformation<br />

<strong>of</strong> V is itself an inner product space, the structure <strong>of</strong> the continuous groups contained in it can be<br />

exploited by using ideas similar <strong>to</strong> those developed in the preceding chapter. In addition, the group<br />

structure also gives rise <strong>to</strong> a special parallelism on the groups.<br />

Section 63. The General Linear Group <strong>and</strong> Its Subgroups<br />

In section 17 we pointed out that the vec<strong>to</strong>r space L ( V ; V ) has the structure <strong>of</strong> an algebra,<br />

the product operation being the composition <strong>of</strong> linear transformations. Since V is an inner product<br />

space, the transpose operation<br />

( ) → ( )<br />

T : L V ; V L V ; V (63.1)<br />

is defined by (18.1); i.e.,<br />

T<br />

uA ⋅ v= Auv ⋅ , uv , ∈V (63.2)<br />

for any A ∈L ( V ; V ) . The inner product <strong>of</strong> any , ∈ ( ; )<br />

(cf. Exercise 19.4))<br />

AB L V V is then defined by<br />

T<br />

( )<br />

AB ⋅ =tr AB (63.3)<br />

only if<br />

From Theorem 22.2, a linear transformation ∈ ( ; )<br />

A L V V is an isomorphism <strong>of</strong> V if <strong>and</strong><br />

det A ≠0<br />

(63.4)<br />

Clearly, if A <strong>and</strong> B are isomorphisms, then AB <strong>and</strong> BA are also isomorphisms, <strong>and</strong> from Exercise<br />

22.1 we have<br />

463


464 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

( ) = ( ) = ( )( )<br />

det AB det BA det A det B (63.5)<br />

As a result, the set <strong>of</strong> all linear isomorphisms <strong>of</strong> V forms a group GL ( V ) , called the general<br />

linear group <strong>of</strong>V . This group was mentioned in Section 17. We claim that GL ( V ) is the disjoint<br />

union <strong>of</strong> two connected open sets in L ( V ; V ) . This fact is more <strong>of</strong> less obvious since GL ( V ) is<br />

the reimage <strong>of</strong> the disjoint union ( −∞,0) ∪( 0, ∞ ) under the continuous map<br />

det : L ( V ; V ) → R<br />

If we denote the primeages <strong>of</strong> ( −∞ ,0)<br />

<strong>and</strong> ( 0,∞)<br />

under the mapping det by GL ( V ) −<br />

<strong>and</strong><br />

GL ( V ) +<br />

, respectively, then they are disjoint connected open sets in ( ; )<br />

( ) = ( ) ∪ ( )<br />

− +<br />

L V V , <strong>and</strong><br />

GL V GL V GL V (63.6)<br />

Notice that from (63.5) GL ( V ) +<br />

is, but GL ( V ) −<br />

is not, closed with respect <strong>to</strong> the group<br />

operation. So GL ( V ) +<br />

is, but GL ( V ) −<br />

is not, a subgroup <strong>of</strong> GL ( V ) . We call the elements <strong>of</strong><br />

GL ( V ) +<br />

<strong>and</strong> GL ( V ) −<br />

proper <strong>and</strong> improper transformations <strong>of</strong>V , respectively. They are<br />

separated by the hypersurface S defined by<br />

( )<br />

A∈S ⇔A∈ L V ; V <strong>and</strong> det A = 0<br />

Now since GL ( V ) is an open set in L ( V ; V ) , any coordinate system in L ( V ; V )<br />

corresponds <strong>to</strong> a coordinate system in GL ( V ) when its coordinate neighborhood is restricted <strong>to</strong> a<br />

subset <strong>of</strong> GL ( V ). For example, if { e i } <strong>and</strong> { e i<br />

} are reciprocal bases forV , then { ⊗ i<br />

i }<br />

for L ( V ; V ) which gives rise <strong>to</strong> the Cartesian coordinate system{ X j<br />

i }. The restriction <strong>of</strong> { X<br />

j<br />

i }<br />

<strong>to</strong> GL ( V ) is then a coordinate system on GL ( V ) . A Euclidean geometry can then be defined on<br />

GL ( V ) as we have done in general on an arbitrary open set in a Euclidean space. The Euclidean<br />

geometry on GL ( V ) is not <strong>of</strong> much interest, however, since it is independent <strong>of</strong> the group<br />

structure on GL ( V ).<br />

The group structure on GL ( V ) is characterized by the following operations.<br />

1. Left-multiplication by any A∈GL ( V ) ,<br />

e e is a basis


Sec. 64 • The General Linear Group 465<br />

L<br />

A<br />

( ) ( )<br />

: GL V →GL V<br />

is defined by<br />

A<br />

( X) ≡AX,<br />

X∈GL ( V )<br />

2. Right-multiplication by any A∈GL ( V )<br />

L (63.7)<br />

R<br />

A<br />

( ) ( )<br />

: GL V →GL V<br />

is defined by<br />

A<br />

( X) ≡XA,<br />

X∈GL ( V )<br />

R (63.8)<br />

3. Inversion<br />

J : GL ( V ) →GL ( V )<br />

is defined by<br />

( X) = X −1 , X∈GL ( V )<br />

J (63.9)<br />

Clearly these operations are smooth mappings, so they give rise <strong>to</strong> various gradients which<br />

are fields <strong>of</strong> linear transformations <strong>of</strong> the underlying inner product space L ( V ; V ). For example,<br />

the gradient <strong>of</strong> L<br />

A<br />

is a constant field given by<br />

since for any Y ∈L ( V ; V )<br />

we have<br />

( ) , ( ; )<br />

∇ L Y A<br />

= AY Y ∈ L V V<br />

(63.10)<br />

d<br />

d<br />

( t) ( t)<br />

dt<br />

L X + Y A<br />

= + =<br />

dt<br />

AX A Y AY<br />

t= 0 t=<br />

0<br />

By the same argument ∇ R is also a constant field <strong>and</strong> is given by<br />

A


466 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

( ) , ( ; )<br />

∇ R Y A<br />

= YA Y ∈ L V V<br />

(63.11)<br />

On the other h<strong>and</strong>, ∇J is not a constant field; its value at any point X ∈GL ( V ) is given by<br />

In particular, at =<br />

−1<br />

( X) ⎤( Y) X YX, Y L ( V ; V )<br />

⎡⎣∇ J ⎦ =− ∈<br />

(63.12)<br />

X I the value <strong>of</strong> ∇ ( I)<br />

J is simply the negation operation,<br />

( I) ⎤( Y) Y, Y L ( V ; V )<br />

⎡⎣∇ J ⎦ =− ∈<br />

(63.13)<br />

We leave the pro<strong>of</strong> <strong>of</strong> (63.13) as an exercise.<br />

The gradients ∇ L A<br />

, A∈GL ( V ), can be regarded as a parallelism on ( )<br />

following way: For any two points X <strong>and</strong> Y in GL ( V ) there exists a unique ∈ ( )<br />

that Y L<br />

= A<br />

X<br />

GL V in the<br />

. The corresponding gradient ∇ L A<br />

at the point X is a linear isomorphism<br />

( X) ⎤:<br />

GL ( V ) →GL ( V )<br />

A GL V such<br />

⎡⎣∇L A ⎦<br />

(63.14)<br />

X Y<br />

where GL ( V ) X<br />

denotes the tangent space <strong>of</strong> GL ( V ) at X , a notation consistent with that<br />

introduced in the preceding chapter. Here, <strong>of</strong> course, GL ( V ) X<br />

coincides with ( ; )<br />

X<br />

is a copy <strong>of</strong> L ( V ; V ). We inserted the argument X in ∇ L A ( X)<br />

<strong>to</strong> emphasize the fact that<br />

∇ L A ( X)<br />

is a linear map from the tangent space <strong>of</strong> GL ( V ) at X <strong>to</strong> the tangent space <strong>of</strong> ( )<br />

the image point Y. This mapping is given by (63.10) via the isomorphism <strong>of</strong> GL ( V ) X<br />

<strong>and</strong><br />

GL ( V ) Y<br />

with L ( V ; V ).<br />

L V V , which<br />

GL V at<br />

From (63.10) we see that the parallelism defined by (63.14) is not the same as the Euclidean<br />

parallelism. Also, ∇L A<br />

is not the same kind <strong>of</strong> parallelism as the Levi-Civita parallelism on a<br />

hypersurface because it is independent <strong>of</strong> any path joining X <strong>and</strong> Y. In fact, if X <strong>and</strong> Y do not<br />

belong <strong>to</strong> the same connected set <strong>of</strong> GL ( V ) , then there exists no smooth curve joining them at all,<br />

X<br />

X A∈GL V the<br />

but the parallelism ∇L A ( ) is still defined. We call the parallelism ∇L A ( ) with ( )<br />

Cartan parallelism on GL ( V ), <strong>and</strong> we shall study it in detail in the next section.<br />

The choice <strong>of</strong> left multiplication rather than the right multiplication is merely a convention.<br />

The gradient RA<br />

also defined a parallelism on the group. Further, the two parallelisms ∇L A<br />

<strong>and</strong><br />

R are related by<br />

∇ A


Sec. 64 • The General Linear Group 467<br />

∇ L =∇J ∇R ∇J (63.15)<br />

A<br />

−1<br />

A<br />

Since LA<br />

<strong>and</strong><br />

RA<br />

are related by<br />

L = JR J (63.16)<br />

A<br />

−1<br />

A<br />

for all A∈GL ( V ).<br />

Before closing the section, we mention here that besides the proper general linear group<br />

GL ( V ) +<br />

, several other subgroups <strong>of</strong> ( )<br />

linear group SL ( V ) is defined by the condition<br />

GL V are important in the applications. First, the special<br />

( ) det A 1<br />

A∈SL V ⇔ =<br />

(63.17)<br />

2<br />

Clearly SL ( V ) is a hypersurface <strong>of</strong> dimension N − 1 in the inner product space L ( V ; V ).<br />

Unlike GL ( V ), SL ( V ) is a connected continuous group. The unimodular group ( )<br />

defined similiarly by<br />

( ) det A 1<br />

Thus SL ( V ) is the proper subgroup <strong>of</strong> UM ( V ) , namely<br />

UM V is<br />

A∈UM V ⇔ =<br />

(63.18)<br />

+<br />

( ) = ( )<br />

Next the orthogonal group O ( V ) is defined by the condition<br />

UM V SL V (63.19)<br />

−1 T<br />

( )<br />

A∈O V ⇔ A = A<br />

(63.20)<br />

From (18.18) or (63.2) we see that A belongs <strong>to</strong> O ( V ) if <strong>and</strong> only if it preserves the inner product<br />

<strong>of</strong> V<br />

, i.e.,<br />

Au ⋅ Av = u⋅v, u,<br />

v∈V (63.21)<br />

Further, from (63.20) if ∈ ( ),<br />

O V is a<br />

O V has a proper component <strong>and</strong> an improper component, the<br />

A O V then det A has absolute value 1. Consequently, ( )<br />

subgroup <strong>of</strong> UM ( V ). As usual, ( )


468 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

former being a subgroup <strong>of</strong> O ( V ), denoted by ( )<br />

rotational group <strong>of</strong> V . As we shall see in Section 65, O ( V ) is a hypersurface <strong>of</strong> dimension<br />

1<br />

2<br />

A∈O V , then<br />

SO V , called the special orthogonal group or the<br />

N( N − 1)<br />

. From (63.20), O ( V ) is contained in the sphere <strong>of</strong> radius N in ( ; )<br />

( )<br />

L V V since if<br />

T<br />

1<br />

A⋅ A = tr AA = tr AA − = tr I = N<br />

(63.22)<br />

As a result, O ( V ) is bounded in ( ; )<br />

L V V .<br />

Since UM ( V ), SL ( V ), O ( V ), <strong>and</strong> SO ( V ) are subgroups <strong>of</strong> ( )<br />

the group operations <strong>of</strong> GL ( V ) can be identified as the group operations <strong>of</strong> the subgroups. In<br />

particular, if XY , ∈SL ( V ) <strong>and</strong> Y L ( X)<br />

, then the restriction <strong>of</strong> the mapping ( X)<br />

by (63.14) is a mapping<br />

= A<br />

( X) ⎤:<br />

SL ( V ) → SL ( V )<br />

⎣ A ⎦ X Y<br />

GL V , the restrictions <strong>of</strong><br />

∇ A<br />

L defined<br />

⎡∇L (63.23)<br />

Thus there exists a Cartan parallelism on the subgroups as well as on GL ( V ) . The tangent spaces<br />

<strong>of</strong> the subgroups are subspaces <strong>of</strong> L ( V ; V ) <strong>of</strong> course; further , as explained in the preceding<br />

chapter, they vary from point <strong>to</strong> point. We shall characterize the tangent spaces <strong>of</strong> SL ( V ) <strong>and</strong><br />

SO ( V ) at the identity element I in Section 66. The tangent space at any other point can then be<br />

obtained by the Cartan parallelism, e.g.,<br />

for any X∈<br />

SL ( V ) .<br />

Exercise<br />

63.1 Verify (63.12) <strong>and</strong> (63.13).<br />

( )<br />

( ) = ⎡∇L<br />

( I) ⎤ ( )<br />

SL V<br />

X ⎣ X ⎦ SL V<br />

I


Sec. 64 • The Parallelism <strong>of</strong> Cartan 469<br />

Section 64. The Parallelism <strong>of</strong> Cartan<br />

The concept <strong>of</strong> Cartan parallelism on GL ( V ) <strong>and</strong> on its various subgroups was introduced<br />

in the preceding section. In this section, we develop the concept in more detail. As explained<br />

before, the Cartan parallelism is path-independent. To emphasize this fact, we now replace the<br />

notation ∇L A ( X)<br />

by the notation C ( XY , ), where YAX. =<br />

Then we put<br />

for any X∈GL ( V ). It can be verified easily that<br />

( X) ≡ ( I,<br />

X) ( I)<br />

C C ≡∇L X<br />

(64.1)<br />

−<br />

( XY , ) = ( Y) ( X) 1<br />

C C C (64.2)<br />

for all X <strong>and</strong> Y. We use the same notation C ( XY , )<br />

pairs X, Y in GL ( V ) or in any continuous subgroup <strong>of</strong> GL ( V ) , such as ( )<br />

C ( XY , ) denotes either the linear isomorphism<br />

( XY , ): GL ( V ) →GL ( V )<br />

for the Cartan parallelism from X <strong>to</strong> Y for<br />

SL V . Thus<br />

C (64.3)<br />

X<br />

Y<br />

or its various restrictions such as<br />

when X, Y belong <strong>to</strong> SL ( V ).<br />

Let V be a vec<strong>to</strong>r field on GL ( V ) , that is<br />

( XY , ): SL ( V ) → SL ( V )<br />

C (64.4)<br />

X<br />

( ) → ( )<br />

Y<br />

V : GL V L V ; V (64.5)<br />

Then we say that V is a left-invarient field if its values are parallel vec<strong>to</strong>rs relative <strong>to</strong> the Cartan<br />

parallelism, i.e.,<br />

( ) ⎤ ( )<br />

( ) ( )<br />

⎡⎣C X, Y ⎦ V X = V Y<br />

(64.6)


470 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

for any X, Y in GL ( V ). Since the Cartan parallelism is not the same as the Euclidean parallelism<br />

induced by the inner product space L ( V ; V ) , a left-invarient field is not a constant field. From<br />

(64.2) we have the following representations for a left-invarient field.<br />

Theorem 64.1. A vec<strong>to</strong>r field V is left-invariant if <strong>and</strong> only if it has the representation<br />

for all X∈GL ( V ).<br />

( ) ( ) ( )<br />

V X ⎣C X ⎦ V I<br />

(64.7)<br />

= ⎡ ⎤( )<br />

As a result, each tangent vec<strong>to</strong>r V( I ) at the identity element I has a unique extension in<strong>to</strong><br />

a left-variant field. Consequently, the set <strong>of</strong> all left-invariant fields, denoted by g l ( V ), is a copy<br />

<strong>of</strong> the tangent space GL ( V ) I<br />

which is canonically isomorphic <strong>to</strong> L ( V ; V ) . We call the<br />

restriction map<br />

I<br />

: gl ;<br />

(64.8)<br />

( V ) →GL ( V ) ≅ L ( V V )<br />

the st<strong>and</strong>ard representation <strong>of</strong> gl ( V ). Since the elements <strong>of</strong> ( V )<br />

I<br />

g l satisfy the representation<br />

(64.7), they characterize the Caratan parallelism completely by the condition (64.6) for<br />

V∈<br />

V<br />

V<br />

g l . In the next section we shall show that g l ( ) has the structure <strong>of</strong> a Lie algebra,<br />

all ( )<br />

so we call it the Lie Algebra <strong>of</strong> GL ( V ).<br />

Now using the Cartan parallelism C ( XY , )<br />

, we can define an opereation <strong>of</strong> covariant<br />

derivative by the limit <strong>of</strong> difference similar <strong>to</strong> (56.34), which defines the covariant derivative<br />

relative <strong>to</strong> the Levi-Civita parallelism. Specifically, let X ( t)<br />

be a smooth curve in GL ( V ) <strong>and</strong> let<br />

U () t be a vec<strong>to</strong>r field on X () t . Then we defince the covariant derivative <strong>of</strong> U () t along<br />

X () t relative <strong>to</strong> the Cartan parallelism by<br />

DU<br />

Dt<br />

( ) ⎤( ( ))<br />

() t ( +Δ ) − ⎡C<br />

( ),<br />

( +Δ )<br />

U t t<br />

⎣<br />

X t X t t<br />

⎦<br />

U t<br />

=<br />

Δ t<br />

(64.9)<br />

Since the Cartan parallelism is defined not just on GL ( V ) but also on the various continuous<br />

subgroups <strong>of</strong> GL ( V ), we may use the same formula (64.9) <strong>to</strong> define the covariant derivative <strong>of</strong><br />

tangent vec<strong>to</strong>r fields along a smooth curve in the subgroups also. For this reason we shall now<br />

derive a general representation for the covariant derivative relative <strong>to</strong> the Cartan parallelism<br />

without restricting the underlying continuous group <strong>to</strong> be the general linear group GL ( V ). Then


Sec. 64 • The Parallelism <strong>of</strong> Cartan 471<br />

we can apply the representation <strong>to</strong> vec<strong>to</strong>r fields on GL ( V ) as well as <strong>to</strong> vec<strong>to</strong>r fields on the<br />

various continuous subgroups <strong>of</strong> GL ( V ), such as the special linear group SL ( V ).<br />

The simplest way <strong>to</strong> represent the covariant derivative defined by (64.9) is <strong>to</strong> first express<br />

the vec<strong>to</strong>r field U () t in component form relative <strong>to</strong> a basis <strong>of</strong> the Lie algebra <strong>of</strong> the underlying<br />

continuous group. From (64.7) we know that the values ( )<br />

field <strong>of</strong> basis ( ), 1,...,M<br />

{ Γ<br />

Γ= }<br />

{ Γ<br />

, Γ= 1,...,M }<br />

E X <strong>of</strong> a left-invariant<br />

E X form a basis <strong>of</strong> the tangent space at X for all X belonging <strong>to</strong><br />

the underlying group. Here M denotes the dimension <strong>of</strong> the group; it is purposely left arbitrary so<br />

as <strong>to</strong> achieve generality in the representation. Now since U ( t)<br />

is a tangent vec<strong>to</strong>r at X () t , it can be<br />

{ Γ ( )}<br />

represented as usual by the component form relative <strong>to</strong> the basis ( t)<br />

( )<br />

( ) ˆ Γ<br />

t = U ( t) ( t)<br />

Γ<br />

E X , say<br />

U E X (64.10)<br />

where Γ is summed from 1 <strong>to</strong> M . Substituting (64.10) in<strong>to</strong> (64.9) <strong>and</strong> using the fact that the basis<br />

E is formed by parallel fields relative <strong>to</strong> the Cartan parallelism, we obtain directly<br />

{ } Γ<br />

( ) ˆ Γ<br />

( )<br />

DU<br />

t dU t<br />

=<br />

Dt dt<br />

Γ<br />

( () t )<br />

E X (64.11)<br />

This formula is comparable <strong>to</strong> the representation <strong>of</strong> the covariant derivative relative <strong>to</strong> the<br />

Euclidean parallelism when the vec<strong>to</strong>r field is expressed in component form in the terms <strong>of</strong> a<br />

Cartesian basis.<br />

As we shall see in the next section, the left-variant basis { E Γ } used in the representation<br />

(64.11) is not the natural basis <strong>of</strong> any coordinate system. Indeed, this point is the major difference<br />

between the Cartan parallelism <strong>and</strong> the Euclidean parallelism, since relative <strong>to</strong> the latter a parallel<br />

basis is a constant basis which is the natural basis <strong>of</strong> a Cartesian coordinate system. If we<br />

introduce a local coordinate system with natural basis{ H , Γ= Γ<br />

1,...,M }, then as usual we can<br />

Γ<br />

represent X () t by its coordinate functions X ( t)<br />

, <strong>and</strong> E<br />

Γ<br />

<strong>and</strong> U ( t)<br />

by their components<br />

( )<br />

( )<br />

Γ<br />

( ) ( ) ( )<br />

Δ<br />

E = E H <strong>and</strong> U t = U t H X t<br />

(64.12)<br />

Γ Γ Δ Γ<br />

By using the usual transformation rule, we then have<br />

( )<br />

( ) ( ) ( )<br />

Uˆ<br />

Γ Δ Γ<br />

t = U t F X t<br />

(64.13)<br />

Δ<br />

where ⎡<br />

⎣F Γ<br />

Δ<br />

⎤<br />

⎦ is the inverse <strong>of</strong> ⎡<br />

⎣ E Γ Δ<br />

⎤<br />

⎦ . Substituting (64.13) in<strong>to</strong> (64.11), we get


472 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

DU<br />

⎛dU Ω ∂ FΩ<br />

d X<br />

= ⎜ + U<br />

Σ<br />

Dt ⎝ dt ∂ X dt<br />

Δ Γ Σ<br />

E<br />

Δ<br />

Γ<br />

⎞<br />

⎟H Δ<br />

(64.14)<br />

⎠<br />

We can rewrite this formula as<br />

Δ<br />

DU<br />

⎛dU Ω Δ d X<br />

= ⎜ + U LΩΣ<br />

Dt ⎝ dt dt<br />

Σ<br />

⎞<br />

⎟H Δ<br />

(64.15)<br />

⎠<br />

where<br />

Γ<br />

∂FΩ<br />

L = E =−F<br />

Σ<br />

∂X<br />

Δ Δ Γ<br />

ΩΣ Γ Ω<br />

∂E<br />

∂X<br />

Δ<br />

Γ<br />

Σ<br />

(64.16)<br />

Now the formula (64.15) is comparable <strong>to</strong> (56.37) with L Δ ΩΣ<br />

playing the role <strong>of</strong> the Chris<strong>to</strong>ffel<br />

symbols, except that L Δ ΩΣ<br />

is not symmetric with respect <strong>to</strong> the indices Ω <strong>and</strong> Σ . For definiteness,<br />

we call L Δ ΩΣ<br />

the Cartan symbols. From (64.16) we can verify easily that they do not depend on the<br />

choice <strong>of</strong> the basis { E Γ }.<br />

It follows from (64.12) 1 <strong>and</strong> (64.16) that the Cartan symbols obey the same transformation<br />

X Γ is another coordinate system in which the<br />

rule as the Chris<strong>to</strong>ffel symbols. Specifically, if ( )<br />

Cartan symbols are L Δ ΩΣ , then 2<br />

L<br />

Δ Ψ Θ Δ Φ<br />

Φ ∂X ∂X ∂X ∂X ∂ X<br />

= L<br />

+<br />

∂X ∂X ∂X ∂X ∂X ∂X<br />

Δ<br />

ΩΣ ΨΘ Φ Ω Σ Φ Ω Σ<br />

(64.17)<br />

This formula is comparable <strong>to</strong> (56.15). In view <strong>of</strong> (64.15) <strong>and</strong> (64.17) we can define the covariant<br />

derivative <strong>of</strong> a vec<strong>to</strong>r field relative <strong>to</strong> the Cartan parallelism by<br />

where { Σ<br />

}<br />

Δ<br />

⎛∂U<br />

Ω Δ ⎞<br />

Σ<br />

∇ U= ⎜ + U L<br />

Σ<br />

ΩΣ ⎟HΔ<br />

⊗ H (64.18)<br />

⎝∂<br />

X ⎠<br />

H denotes the dual basis <strong>of</strong>{ H Δ }. The covariant derivative defined in this way clearly<br />

possesses the following property.<br />

Theorem 64.2. A vec<strong>to</strong>r field U is left-invariant if <strong>and</strong> only if its covariant derivative relative <strong>to</strong><br />

the Cartan parallelism vanishes.


Sec. 64 • The Parallelism <strong>of</strong> Cartan 473<br />

The pro<strong>of</strong> <strong>of</strong> this proposition if more or less obvious, since the condition<br />

∂U<br />

∂ X<br />

Δ<br />

Σ<br />

Ω Δ<br />

+ U L = 0<br />

ΩΣ<br />

(64.19)<br />

is equivalent <strong>to</strong> the condition<br />

∂Uˆ<br />

∂ X<br />

Δ<br />

Σ<br />

= 0<br />

(64.20)<br />

where U ˆ Δ denotes the components <strong>of</strong> U relative <strong>to</strong> the parallel basis{ E Γ }. The condition (64.20)<br />

means simply that U ˆ Δ are constant, or, equivalent, U is left-invariant.<br />

Comparing (64.16) with (59.16), we see that the Cartan parallelism also possesses the<br />

following property.<br />

Theorem 64.3. The curvature tensor whose components are defined by<br />

∂ ∂<br />

R ≡ L − L + L L −L L<br />

Δ<br />

Ω<br />

∂X<br />

∂X<br />

Γ Γ Γ Φ Γ Φ Γ<br />

ΣΔΩ ΣΩ ΣΔ ΣΩ ΦΔ ΣΔ ΦΩ<br />

(64.21)<br />

vanishes identically.<br />

Notice that (64.21) is comparable with (59.10) where the Chris<strong>to</strong>ffel symbols are replaced<br />

by the Cartan symbols. The vanishing <strong>of</strong> the curvature tensor<br />

R Γ = ΣΔΩ<br />

0<br />

(64.22)<br />

is simply the condition <strong>of</strong> integrability <strong>of</strong> equation (64.19) whose solutions are left-invariant fields.<br />

From the transformation rule (64.17) <strong>and</strong> the fact that the second derivative therein is<br />

symmetric with respect <strong>to</strong> the indices Ω <strong>and</strong> Σ , we obtain<br />

L − L = L −L<br />

∂X ∂X ∂X<br />

∂X ∂X ∂X<br />

Δ Ψ Θ<br />

Δ Δ Φ Φ<br />

ΩΣ ΣΩ ( ΨΘ ΘΨ ) Φ Ω Σ<br />

(64.23)<br />

which shows that the quanties defined by<br />

Δ Δ Δ<br />

T ≡L − L<br />

(64.24)<br />

ΩΣ ΩΣ ΣΩ


474 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

are the components <strong>of</strong> a third-order tensor field. We call the T the <strong>to</strong>rsion tensor <strong>of</strong> the Cartan<br />

parallelism. To see the geometric meaning <strong>of</strong> this tensor, we substitute the formula (64.16) in<strong>to</strong><br />

(64.24) <strong>and</strong> rewrite the result in the following equivalent form:<br />

Δ<br />

∂EΨ<br />

T E E = E −E<br />

Ω<br />

∂X<br />

Δ Ω Σ Ω Ω<br />

ΩΣ Φ Ψ Φ Ψ<br />

∂E<br />

∂X<br />

Δ<br />

Φ<br />

Ω<br />

(64.25)<br />

which is the component representation <strong>of</strong><br />

( )<br />

T EΦ , E ⎡<br />

Ψ<br />

= ⎣ EΦ,<br />

E ⎤<br />

Ψ ⎦<br />

(64.26)<br />

where the right-h<strong>and</strong> side is the Lie bracket <strong>of</strong><br />

EΦ<br />

<strong>and</strong> E<br />

Ψ<br />

, i.e.,<br />

Δ<br />

Δ<br />

⎛ Ω ∂EΨ<br />

Ω ∂EΦ<br />

⎞<br />

⎡<br />

⎣EΦ,<br />

E ⎤<br />

Ψ⎦ ≡ L EΨ = ⎜EΦ −E<br />

Ω Ψ Ω ⎟HΔ<br />

(64.27)<br />

EΦ<br />

⎝ ∂X<br />

∂X<br />

⎠<br />

Equation (64.27) is comparable <strong>to</strong> (49.21) <strong>and</strong> (55.31).<br />

As we shall see in the next section, the Lie bracket <strong>of</strong> any pair <strong>of</strong> left-invariant fields is<br />

itself also a left-invariant field. Indeed, this is the very reason that the set <strong>of</strong> all left-invariant fields<br />

is so endowed with the structure <strong>of</strong> a Lie algebra. As a result, the components <strong>of</strong> ⎡<br />

⎣EΦ,<br />

E ⎤<br />

Ψ⎦<br />

relative <strong>to</strong> the left-invariant basis { E Γ } are constant scalar fields, namely<br />

⎡<br />

⎣E , C Γ<br />

Φ<br />

EΨ⎤=<br />

⎦ ΦΨE Γ<br />

(64.28)<br />

We call C Γ ΦΨ<br />

the structure constants <strong>of</strong> the left-invariant basis{ E<br />

Γ}<br />

. From (64.26), C Γ ΦΨ<br />

are nothing<br />

but the components <strong>of</strong> the <strong>to</strong>rsion tensor T relative <strong>to</strong> the basis{ E Γ }, namely<br />

T = E ⊗E ⊗E<br />

(64.29)<br />

C Γ Φ Ψ<br />

ΦΨ Γ<br />

Now the covariant derivative defined by (64.9) for vec<strong>to</strong>r fields can be generalized as in the<br />

preceding chapter <strong>to</strong> covariant derivatives <strong>of</strong> arbitrary tangential tensor fields. Specifically, the<br />

general formula is<br />

DA<br />

⎧<br />

⎪dA<br />

⎛A L +⋅⋅⋅+ A L ⎞ ⎫<br />

= ⎨ + ⊗⋅⋅⋅⊗<br />

Dt ⎪ dt dt<br />

⎩ ⎝<br />

⎠ ⎭<br />

Γ1...<br />

Γ ΣΓ2... Γr Γ1<br />

Γ1... Γ<br />

r<br />

r−1Σ<br />

Γr<br />

Ω<br />

Δ1...<br />

Δ Δ<br />

s<br />

1... Δs<br />

ΣΩ Δ1...<br />

Δs<br />

ΣΩ d X ⎪<br />

Δs<br />

Γ1... Γr<br />

1...<br />

r<br />

1<br />

A L Σ<br />

Γ Γ Σ<br />

⎬HΓ<br />

H<br />

⎜−<br />

ΣΔ<br />

A L<br />

2... Δs Δ1Ω −⋅⋅⋅−<br />

⎟<br />

Δ1... Δs−1Σ ΔsΩ<br />

⎪<br />

(64.30)


Sec. 64 • The Parallelism <strong>of</strong> Cartan 475<br />

which is comparable <strong>to</strong> (48.7) <strong>and</strong> (56.23). The formula (64.30) represents the covariant derivative<br />

in terms <strong>of</strong> a coordinate system( X Γ<br />

) . If we express A in component form relative <strong>to</strong> a left-variant<br />

basis{ E Γ }, say<br />

Γ<br />

() ˆ 1... Γr<br />

() ()<br />

Δ1... Δs<br />

Γ1<br />

Δs<br />

( ) ( ())<br />

A t = A t E X t ⊗⋅⋅⋅⊗E X t<br />

(64.31)<br />

then the representation <strong>of</strong> the covariant derivative is simply<br />

DA<br />

Dt<br />

() ˆ Γ1...<br />

Γr<br />

t dA ( t)<br />

s<br />

( () t ) () t<br />

Δ1...<br />

Δs<br />

Δ<br />

Γ1<br />

= E X ⊗⋅⋅⋅⊗E X<br />

dt<br />

( )<br />

(64.32)<br />

The formulas (64.30) <strong>and</strong> (64.32) represent the covariant derivative <strong>of</strong> a tangential tensor field A<br />

along a smooth curve X () t . Now if A is a tangential tensor field defined on the continuous<br />

group, then we define the covariant derivative <strong>of</strong> A relative <strong>to</strong> the Cartan parallelism by a formula<br />

similar <strong>to</strong> (56.10) except that we replace the Chris<strong>to</strong>ffel symbols there by the Cartan symbols.<br />

Naturally we say that a tensor field A is left-invariant if ∇A vanishes. The <strong>to</strong>rsion tensor<br />

field T given by (64.29) is an example <strong>of</strong> a left-invariant third-order tensor field. Since a tensor<br />

field is left invariant if <strong>and</strong> only if its components relative <strong>to</strong> the product basis <strong>of</strong> a left-invariant<br />

basis are constants, a representation formally generalizing (64.7) can be stated for left-invariant<br />

tensor fields in general. In particular, if { E Γ }<br />

is a left-invariant field <strong>of</strong> bases, then<br />

E= E1<br />

∧⋅⋅⋅∧E (64.33)<br />

M<br />

is a left-invariant field <strong>of</strong> density tensors on the group.


476 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

Section 65. One-Parameter Groups <strong>and</strong> the Exponential Map<br />

In section 57 <strong>of</strong> the preceding chapter we have introduced the concepts <strong>of</strong> geodesics <strong>and</strong><br />

exponential map relative <strong>to</strong> the Levi-Civita parallelism on a hypersurface. In this section we<br />

consider similar concepts relative <strong>to</strong> the Cartan parallelism. As before, we define a geodesic <strong>to</strong> be a<br />

smooth curve X () t such that<br />

DX<br />

Dt = 0<br />

(65.1)<br />

where X denotes the tangent vec<strong>to</strong>r X <strong>and</strong> where the covariant derivative is taken relative <strong>to</strong> the<br />

Cartan parallelism.<br />

Since (65.1) is formally the same as (57.1), relative <strong>to</strong> a coordinate system ( X Γ<br />

) we have<br />

the following equations <strong>of</strong> geodesics<br />

2 Γ ∑ Δ<br />

d X dX dX L<br />

Γ<br />

∑Δ<br />

M<br />

+ = 0, Γ= 1,...,<br />

(65.2)<br />

2<br />

dt dt dt<br />

which are comparable <strong>to</strong> (57.2). However, the equations <strong>of</strong> geodesics here are no longer the Euler-<br />

Lagrange equations <strong>of</strong> the arc length integral, since the Cartan parallelism is not induced by a<br />

metric <strong>and</strong> the arc length integral is not defined. To interpret the geometric meaning <strong>of</strong> a geodesic<br />

relative <strong>to</strong> the Cartan parallelism, we must refer <strong>to</strong> the definition (64.9) <strong>of</strong> the covariant derivative.<br />

We notice first that if we express the tangent vec<strong>to</strong>r X ( t)<br />

<strong>of</strong> any smooth curve X ( t)<br />

component form relative <strong>to</strong> a left-invariant basis { E Γ },<br />

( )<br />

Γ<br />

( t) = G ( t) ( t)<br />

X E X<br />

(65.3)<br />

then from (64.11) a necessary <strong>and</strong> sufficient condition for X ( t)<br />

<strong>to</strong> be a geodesic is that the<br />

Γ<br />

components G () t be constant independent <strong>of</strong> t . Equivalently, this condition means that<br />

Γ<br />

( )<br />

( t) = G X( t)<br />

X (65.4)<br />

in<br />

where G is a left-invariant field having the component form<br />

G=G Γ E<br />

Γ<br />

(65.5)


Sec. 65 • One-Parameter Groups, Exponential Map 477<br />

where G Γ are constant. In other words, a curve X ( t)<br />

is a geodesic relative <strong>to</strong> the Cartan<br />

parallelism if <strong>and</strong> only if it is an integral curve <strong>of</strong> a left-invariant vec<strong>to</strong>r field.<br />

This characteristic property <strong>of</strong> a geodesic implies immediately the following result.<br />

( )<br />

Theorem 65.1. If X () t is a geodesic tangent <strong>to</strong> the left-invariant field G, then L X A () t is also a<br />

geodesic tangent <strong>to</strong> the same left-invariant field G for all A in the underlying continuous group.<br />

A corollary <strong>of</strong> this theorem is that every geodesic can be extended indefinitely from t = −∞<br />

X is a geodesic defined for an interval, say t∈ [ 0,1]<br />

, then we can extend<br />

t∈ 1, 2 by<br />

<strong>to</strong> t =+∞. Indeed, if () t<br />

X () t <strong>to</strong> the interval [ ]<br />

( ) −1<br />

X<br />

( ) ( )<br />

( )<br />

XIX 0<br />

( ) [ ]<br />

X t+ 1 ≡ L t , t∈<br />

0,1<br />

(65.6)<br />

<strong>and</strong> so forth. An important consequence <strong>of</strong> this extension is the following result.<br />

Theorem 65.2. A smooth curve X () t passing through the identity element I at t = 0 is a geodesic<br />

if <strong>and</strong> only if it forms a one-parameter group, i.e.,<br />

( ) ( ) ( )<br />

X t + t = X t X t , t , t ∈R (65.7)<br />

1 2 1 2 1 2<br />

The necessity <strong>of</strong> (65.7) is a direct consequence <strong>of</strong> (65.6), which may be generalized <strong>to</strong><br />

X<br />

( )<br />

( t t ) ( )<br />

X( t )<br />

+ = L X<br />

(65.8)<br />

1 2 t1<br />

2<br />

for all t 1<br />

<strong>and</strong> t 2<br />

. Here we have used the condition that<br />

X( 0 ) = I (65.9)<br />

Conversely, if (65.7) holds, then by differentiating with respect <strong>to</strong> t 2<br />

<strong>and</strong> evaluating the result at<br />

t<br />

2<br />

= 0 , we obtain<br />

( t ) ⎡ ( t )<br />

( ) ⎤( ( ))<br />

X<br />

1<br />

= ⎣<br />

C X<br />

1 ⎦<br />

X 0<br />

(65.10)<br />

which shows that X () t is the value <strong>of</strong> a particular left-invariant field at all () t<br />

geodesic relative <strong>to</strong> the Cartan parallelism.<br />

X , <strong>and</strong> thus X ( t)<br />

is a


478 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

Now combining the preceding propositions, we see that the class <strong>of</strong> all geodesics can be<br />

characterized in the following way. First, for each left-invariant field G there exists a unique oneparameter<br />

group X () t such that<br />

( 0 ) = G( I)<br />

X (65.11)<br />

where G( I)<br />

is the st<strong>and</strong>ard representation <strong>of</strong> G . Next, the set <strong>of</strong> all geodesics tangent <strong>to</strong> G can be<br />

represented by L X ( t)<br />

for all A belonging <strong>to</strong> the underlying group.<br />

A<br />

( )<br />

As in Section 57, the one-<strong>to</strong>-one correspondence between G( I)<br />

<strong>and</strong> X () t gives rise <strong>to</strong> the<br />

notion <strong>of</strong> the exponential map at the indentity element I . For brevity, let A be the st<strong>and</strong>ard<br />

representation <strong>of</strong> G , i.e.,<br />

( )<br />

A ≡G I (65.12)<br />

Then we define<br />

( ) ( )<br />

X 1 = exp A ≡expA (65.13)<br />

I<br />

which is comparable <strong>to</strong> (57.19). As explained in Section 57, (65.13) implies that<br />

( t) = exp( t)<br />

X A (65.14)<br />

for all t∈R . Here we have used the extension property <strong>of</strong> the geodesic. Equation (65.14) formally<br />

represents the one-parameter group whose initial tangent vec<strong>to</strong>r at the identity element I is A .<br />

We claim that the exponential map defined by (65.13) can be represented explicitly by the<br />

exponential series<br />

Clearly, this series converges for each ∈ ( ; )<br />

1 2 1 3 1 n<br />

exp( A) = I+ A+ A + A +⋅⋅⋅+ A +⋅⋅⋅<br />

(65.15)<br />

2! 3! n!<br />

A L V V . Indeed, since we have<br />

n n<br />

A ≤ A (65.16)<br />

for all positive integers n , the partial sums <strong>of</strong> (65.15) form a Cauchy sequence in the inner product<br />

L V ; V . That is,<br />

space ( )


Sec. 65 • One-Parameter Groups, Exponential Map 479<br />

1 n 1 m 1 n 1 m<br />

A +⋅⋅⋅+ A ≤ A +⋅⋅⋅+ A (65.17)<br />

n! m! n! m!<br />

<strong>and</strong> the right-h<strong>and</strong> side <strong>of</strong> (65.17) converges <strong>to</strong> zero as n <strong>and</strong> m approach infinity.<br />

Now <strong>to</strong> prove that (65.15) is the correct representation for the exponential map, we have <strong>to</strong><br />

show that the series<br />

1 2 2 1 n n<br />

exp( At) = I+ At+ A t +⋅⋅⋅+ A t +⋅⋅⋅<br />

(65.18)<br />

2! n!<br />

defines a one-parameter group. This fact is more or less obvious since the exponential series<br />

satisfies the usual power law<br />

( )<br />

( t) ( t ) = ( t + t )<br />

exp A exp A exp A (65.19)<br />

2 1 2<br />

which can be verified by direct multiplication <strong>of</strong> the power series for exp( A t 1 ) <strong>and</strong> exp( t 2 )<br />

Finally, it is easily seen that the initial tangent <strong>of</strong> the curve exp( A t)<br />

is A since<br />

A .<br />

d ⎛ 1 2 2 ⎞<br />

⎜I+ At+ A t +⋅⋅⋅ ⎟ = A (65.20)<br />

dt ⎝ 2 ⎠ t=<br />

0<br />

This completes the pro<strong>of</strong> <strong>of</strong> the representation (65.18).<br />

We summarize our results as follows.<br />

Theorem 65.3. Let G be a left-invariant field with st<strong>and</strong>ard representation A . Then the geodesics<br />

X t tangent <strong>to</strong> G can be expressed by<br />

()<br />

where the initial point X ( 0)<br />

is arbitrary.<br />

⎛<br />

2 2<br />

() t ( 0exp ) ( t) ( 0) t 1 ⎞<br />

X = X A = X ⎜I+ A + A t ⋅⋅⋅⎟<br />

(65.21)<br />

⎝ 2! ⎠<br />

In the view <strong>of</strong> this representation, we see that the flow generated by the left-invariant field<br />

G is simply the right multiplication by exp( A t)<br />

, namely<br />

ρ R<br />

(65.22)<br />

t<br />

=<br />

exp( At)


480 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

for all t . As a result, if K is another left-invariant field, then the Lie bracket <strong>of</strong> G with K is given<br />

by<br />

[ ]( )<br />

( exp( t)<br />

) − ( ) exp( t)<br />

K X A K X A<br />

GK , X = lim<br />

(65.23)<br />

t→0<br />

t<br />

This formula implies immediately that [ GK , ] is also a left-invariant field. Indeed, if the st<strong>and</strong>ard<br />

representation <strong>of</strong> K is B , then from the representation (64.7) we have<br />

K( X) = XB (65.24)<br />

<strong>and</strong> therefore<br />

( exp( t)<br />

) = exp( t)<br />

K X A X A B (65.25)<br />

Substituting (65.24) <strong>and</strong> (65.25) in<strong>to</strong> (65.23) <strong>and</strong> using the power series representation (65.18), we<br />

obtain<br />

[ GK , ]( X) = X( AB−BA )<br />

(65.26)<br />

which shows that [ GK , ] is left-invariant with the st<strong>and</strong>ard representation<br />

terms <strong>of</strong> the st<strong>and</strong>ard representation the Lie bracket on the Lie algebra is given by<br />

AB − BA . Hence in<br />

[ AB , ] = AB−<br />

BA (65.27)<br />

So far we have shown that the set <strong>of</strong> left-invariant fields is closed with respect <strong>to</strong> the operation <strong>of</strong><br />

the Lie bracket. In general a vec<strong>to</strong>r space equipped with a bilinear bracket product which obeys the<br />

Jacobi identities<br />

<strong>and</strong><br />

[ , ] = − [ , ]<br />

AB BA (65.28)<br />

[ ] [ ] [ ]<br />

⎡⎣A, B, C ⎤⎦+ ⎡⎣B, C, A ⎤⎦+ ⎡⎣C, A,<br />

B ⎤⎦=<br />

0 (65.29)<br />

is called a Lie Algebra. From (65.27) the Lie bracket <strong>of</strong> left-invariant fields clearly satisfies the<br />

identities (65.28) <strong>and</strong> (65.29). As a result, the set <strong>of</strong> all left-invariant fields has the structure <strong>of</strong> a<br />

Lie algebra with respect <strong>to</strong> the Lie bracket. This is why that set is called the Lie algebra <strong>of</strong> the<br />

underlying group, as we have remarked in the preceding section.


Sec. 65 • One-Parameter Groups, Exponential Map 481<br />

Before closing the section, we remark that the Lie algebra <strong>of</strong> a continuous group depends<br />

only on the identity component <strong>of</strong> the group. If two groups share the same identity component,<br />

then their Lie algebras are essentially the same. For example, the Lie algebra <strong>of</strong> GL ( V ) <strong>and</strong><br />

GL ( V ) +<br />

are both representable by L ( V ; V ) with the Lie bracket given by (65.27). The fact that<br />

GL ( V ) has two components, namely GL ( V ) +<br />

<strong>and</strong> GL ( V ) −<br />

cannot be reflected in any way by the<br />

Lie algebra g l ( V ). We shall consider the relation between the Lie algebra <strong>and</strong> the identity<br />

component <strong>of</strong> the underlying group in more detail in the next section.<br />

Exercises<br />

65.1. Establish the following properties <strong>of</strong> the function exp on L ( V ; V ) :<br />

(a) exp( A+ B) = ( expA)( expB)<br />

if AB=<br />

BA.<br />

−<br />

(b) ( ) ( ) 1<br />

exp<br />

− A = exp A .<br />

(c) exp 0=<br />

I<br />

−1 −1<br />

(d) B( expA) B = expBAB for regular B .<br />

Τ<br />

Τ<br />

A=<br />

A if <strong>and</strong> only if exp = ( exp )<br />

Τ<br />

(e)<br />

A A .<br />

(f) A=−A if <strong>and</strong> only if exp A is orthogonal.<br />

tr<br />

(g) det ( exp A ) = e<br />

A .<br />

65.2 If P is a projection, show that<br />

for λ∈R .<br />

( e λ<br />

)<br />

expλ P= I+ −1<br />

P


482 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

Section 66. Subgroups <strong>and</strong> Subalgebras<br />

In the preceding section we have shown that the set <strong>of</strong> left invariant fields is closed with<br />

respect <strong>to</strong> the Lie bracket. This is an important property <strong>of</strong> a continuous group. In this section we<br />

shall elaborate further on this property by proving that there exists a one-<strong>to</strong>-one correspondence<br />

V . Naturally we<br />

g l<br />

gl a subalgebra if h is closed with respect <strong>to</strong> the Lie bracket, i.e.,<br />

[ G,H]∈<br />

h whenever both G <strong>and</strong> H belong <strong>to</strong> h . We claim that for each subalgebra h <strong>of</strong><br />

gl there corresponds uniquely a connected continuous subgroup H <strong>of</strong> ( )<br />

between a connected continuous subgroup <strong>of</strong> GL ( V ) <strong>and</strong> a subalgebra <strong>of</strong> ( )<br />

call a subspace h <strong>of</strong> ( V )<br />

( V )<br />

algebra coincides with the restriction <strong>of</strong> h on H .<br />

GL V whose Lie<br />

First, suppose that H is a continuous subgroup <strong>of</strong> GL ( V ) , i.e., H is algebraically a<br />

subgroup <strong>of</strong> GL ( V ) <strong>and</strong> geometrically a smooth hyper surface in ( )<br />

GL V . For example, H may<br />

be the orthogonal group or the special linear group. Then the set <strong>of</strong> all left-invariant tangent vec<strong>to</strong>r<br />

fields on H<br />

in<strong>to</strong> a left-invariant vec<strong>to</strong>r field V on GL ( V ) . Indeed, this extension is provided by the<br />

representation (64.7). That is, we simply define<br />

forms a Lie algebra h . We claim that every element V in h can be extended uniquely<br />

( ) ( ) ( )<br />

V X ⎣C X ⎦ V I<br />

(66.1)<br />

= ⎡ ⎤( )<br />

for all X∈GL ( V ). Here we have used the fact that the Cartan parallelism on H is the restriction<br />

<strong>of</strong> that on GL ( V ) <strong>to</strong> H . Therefore, when X∈H , the representation (66.1) reduces <strong>to</strong><br />

( ) C ( ) ( )<br />

V X ⎣ X ⎦ V I<br />

= ⎡ ⎤( )<br />

since V is left-invariant on H .<br />

From (66.1), V <strong>and</strong> V share the same st<strong>and</strong>ard representation:<br />

A≡<br />

( )<br />

V I<br />

As a result, from (65.27) the Lie bracket on H is related <strong>to</strong> that on GL ( V ) by<br />

[ ]<br />

VU , = ⎡⎣<br />

VU , ⎤⎦<br />

(66.2)


Sec. 66 • Maximal Abelian Subgroups, Subalgebras 483<br />

for all U <strong>and</strong> V in h . This condition shows that the extension h <strong>of</strong> h consisting <strong>of</strong> all left-<br />

gl V<br />

invariant fields V with V in h is a Lie subalgebra <strong>of</strong> ( )<br />

. In view <strong>of</strong> (66.1) <strong>and</strong> (66.2), we<br />

simplify the notation by suppressing the overbar. If this convention is adopted, then h becomes a<br />

V<br />

gl .<br />

Lie subalgebra <strong>of</strong> ( )<br />

h gl is based on the extension (66.1), if<br />

1<br />

Since the inclusion ⊂ ( V )<br />

H <strong>and</strong> H<br />

2<br />

are<br />

continuous subgroups <strong>of</strong> GL ( V ) having the same identity components, then their Lie algebras<br />

h coincide as Lie subalgebras <strong>of</strong> gl ( V ) . In this sense we say that the Lie algebra<br />

h1<br />

<strong>and</strong><br />

2<br />

characterizes only the identity component <strong>of</strong> the underlying group, as we have remarked at the end<br />

SL V .<br />

<strong>of</strong> the preceding section. In particular, the Lie algebra <strong>of</strong> UM ( V ) coincides with that <strong>of</strong> ( )<br />

gl can be identified as the Lie algebra <strong>of</strong> a<br />

It turns out that every Lie subalgebra <strong>of</strong> ( V )<br />

unique connected continuous subgroup <strong>of</strong> gl ( V ) . To prove this, let h be an arbitrary Lie<br />

subalgebra <strong>of</strong> gl ( V )<br />

subspace <strong>of</strong> L ( V ; V ) at each point <strong>of</strong> GL ( V ) . This field <strong>of</strong> subspaces is a distribution on<br />

GL ( V ) as defined in Section 50. According <strong>to</strong> the Frobenius theorem, the distribution is<br />

integrable if <strong>and</strong> only if it is closed with respect <strong>to</strong> the Lie bracket. This condition is clearly<br />

satisfied since h is a Lie subalgebra. As a result, there exists an integral hypersurface <strong>of</strong> the<br />

distribution at each point in GL ( V ).<br />

. Then the values <strong>of</strong> the left-invariant field belonging <strong>to</strong> h form a linear<br />

We denote the maximal connected integral hypersurface <strong>of</strong> the distribution at the identity<br />

by H . Here maximality means that H is not a proper subset <strong>of</strong> any other connected integral<br />

hypersurface <strong>of</strong> the distribution. This condition implies immediately that H is also the maximal<br />

connected integral hypersurface at any point X which belongs <strong>to</strong> H . By virtue <strong>of</strong> this fact we<br />

claim that<br />

L ( H X ) = H<br />

(66.3)<br />

for all X∈H . Indeed, since the distribution is generated by left-invariant fields, its collection <strong>of</strong><br />

maximal connected integral hypersurfaces is invariant under any left multiplication. In particular,<br />

L<br />

X ( H ) is the maximal connected integral hypersurface at the point X, since H contains the<br />

identity I . As a result, (66.3) holds.<br />

1<br />

Now from (66.3) we see that X∈H implies<br />

− −1<br />

X ∈H since X is the only possible element<br />

−1<br />

L X X<br />

= I . Similiarly, if X <strong>and</strong> Y are contained in H , then XY must also be contained<br />

such that ( )<br />

in H since XY is the only possible element such that 1 1 ( )<br />

have used the fact that<br />

L L . For the last condition we<br />

XY = I<br />

Y − X −


484 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

( ) ( )<br />

L L H = L H = H<br />

−1 −1<br />

−1<br />

Y X Y<br />

−1 −1<br />

which follows from (66.3) <strong>and</strong> the fact that X <strong>and</strong> Y are both in H . Thus we have shown that<br />

GL V having h as its Lie algebra.<br />

H is a connected continuous subgroup <strong>of</strong> ( )<br />

Summarizing the results obtained so far, we can state the following theorem.<br />

Theorem 66.1. There exists a one-<strong>to</strong>-one correspondence between the set <strong>of</strong> Lie subalgebras <strong>of</strong><br />

gl ( V ) <strong>and</strong> the set <strong>of</strong> connected continuous subgroups <strong>of</strong> GL ( V ) in such a way that each Lie<br />

gl is the Lie algebra <strong>of</strong> a unique connected continuous subgroup H <strong>of</strong><br />

subalgebra h <strong>of</strong> ( V )<br />

GL ( V ).<br />

s l <strong>and</strong><br />

To illustrate this theorem, we now determine explicitly the Lie algebras ( V )<br />

( V )<br />

SL V <strong>and</strong> SO ( V ) . We claim first<br />

s o <strong>of</strong> the subgroups ( )<br />

s l (66.4)<br />

( V ) tr A 0<br />

A∈ ⇔ =<br />

where tr A denotes the trace <strong>of</strong> A . To prove this, we consider the one-parameter group exp( A t)<br />

for any A∈L ( V ; V ). In order that A∈sl ( V ) , we must have<br />

( ( t)<br />

)<br />

det exp A = 1, t∈R (66.5)<br />

Differentiating this condition with respect <strong>to</strong> t <strong>and</strong> evaluating the result at t = 0 , we obtain [cf.<br />

Exercise 65.1(g)]<br />

d<br />

0= ⎡det ( exp( t)<br />

) tr<br />

dt ⎣<br />

A ⎤<br />

⎦<br />

= A (66.6)<br />

t = 0<br />

Conversely, if tr A vanishes, then (66.5) holds because the one-parameter group property <strong>of</strong><br />

exp<br />

( At)<br />

implies<br />

while the initial condition at t = 0 ,<br />

d<br />

det ( exp( t)<br />

) det ( exp( t)<br />

) tr 0, t<br />

dt ⎡ ⎣<br />

A ⎤= ⎦<br />

A A = ∈ R


Sec. 66 • Maximal Abelian Subgroups, Subalgebras 485<br />

( )<br />

det exp0 = det I = 1<br />

is obvious. Thus we have completed the pro<strong>of</strong> <strong>of</strong> (66.4).<br />

From the representation (65.27) the reader will verify easily that the subspace <strong>of</strong><br />

L V ; V characterized by the right-h<strong>and</strong> side <strong>of</strong> (66.4) is indeed a Lie subalgebra, as it should be.<br />

( )<br />

where<br />

Next we claim that<br />

A∈so ⇔ A =−A<br />

(66.7)<br />

T<br />

( V )<br />

T<br />

A denotes the transpose <strong>of</strong> A . Again we consider the one-parameter group exp( At)<br />

any A∈L ( V ; V ). In order that A∈<br />

( V )<br />

so<br />

, we must have<br />

T<br />

T<br />

( − t) = ⎡ ( t) ⎤ = ( t)<br />

exp A ⎣exp A ⎦ exp A (66.8)<br />

for<br />

Here we have used the identities<br />

− 1<br />

T<br />

( ) = ( −<br />

T<br />

) ( ) = ( )<br />

⎡ ⎣exp A ⎤ ⎦ exp A , ⎡ ⎣exp A ⎤ ⎦ exp A (66.9)<br />

which can be verified directly from (65.15). The condition (66.7) clearly follows from the<br />

condition (66.8). From (65.27) the reader also will verify the fact that the subpace <strong>of</strong><br />

L V ; V characterized by the right-h<strong>and</strong> side is a Lie subalgebra.<br />

( )<br />

The conditions (66.4) <strong>and</strong> (66.7) characterize completely the tangent spaces <strong>of</strong> SL ( V ) <strong>and</strong><br />

SO ( V ) at the identity element I . These conditions verify the claims on the dimensions <strong>of</strong><br />

SL ( V ) <strong>and</strong> SO ( V ) made in Section 63.


486 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

Section 67. Maximal Abelian Subgroups <strong>and</strong> Subalgebras<br />

In this section we consider the problem <strong>of</strong> determining the Abelian subgroups <strong>of</strong> GL ( V ) .<br />

Since we shall use the Lie algebras <strong>to</strong> characterize the subgroups, our results are necessarily<br />

restricted <strong>to</strong> connected continuous Abelian subgroups only. We define first the tentative notion <strong>of</strong><br />

a maximal Abelian subset H <strong>of</strong> GL ( V ). The subset H is required <strong>to</strong> satisfy the following two<br />

conditions:<br />

(i) Any pair <strong>of</strong> elements X <strong>and</strong> Y belonging <strong>to</strong> H commute.<br />

(ii) H is not a proper subset <strong>of</strong> any subset <strong>of</strong> GL ( V ) satisfying condition (i).<br />

Theorem 67.1. A maximal Abelian subset is necessarily a subgroup.<br />

The pro<strong>of</strong> is more or less obvious. Clearly, the identity element I is a member <strong>of</strong> every<br />

−1<br />

maximal Abelian subset. Next, if X belongs <strong>to</strong> a certain maximal Abelian subset H , then X also<br />

belongs <strong>to</strong> H . Indeed, X∈H means that<br />

XY = YX,<br />

Y∈H (67.1)<br />

Multiplying this equation on the left <strong>and</strong> on the right by<br />

−1<br />

X , we get<br />

−1 −1<br />

YX = X Y,<br />

Y∈H (67.2)<br />

1<br />

As a result, X − ∈H since H is maximal. By the same argument we can prove also that<br />

XY∈H whenever X∈H <strong>and</strong> ∈<br />

GL V .<br />

Y H . Thus, H is a subgroup <strong>of</strong> ( )<br />

In view <strong>of</strong> this theorem <strong>and</strong> the opening remarks we shall now consider the maximal,<br />

connected, continuous, Abelian subgroups <strong>of</strong> GL ( V ) . Our first result is the following.<br />

Theorem 67.2. The one-parameter groups exp( At)<br />

<strong>and</strong> exp( Bt)<br />

initial tangents A <strong>and</strong> B commute.<br />

exp<br />

commute if <strong>and</strong> only if their<br />

Sufficiency is obvious, since when AB = BA the series representations for exp( At)<br />

( Bt)<br />

imply directly that exp( A t)<br />

exp( B t)<br />

= exp( Bt) exp( A t)<br />

. In fact, we have<br />

( ) ( ) ( )<br />

( ) ( ) ( )<br />

exp At exp Bt = exp A+ B t = exp Bt exp A t<br />

<strong>and</strong>


Sec. 66 • Maximal Abelian Subgroups, Subalgebras 487<br />

commute, then their initial tangents A <strong>and</strong> B<br />

in this case. Conversely, if exp( At)<br />

<strong>and</strong> exp( Bt)<br />

must also commute, since we can compare the power series expansions for exp( t) exp( t)<br />

exp( Bt) exp( A t)<br />

for sufficiently small t . Thus the proposition is proved.<br />

We note here a word <strong>of</strong> caution: While the assertion<br />

( ) ( ) ( ) ( )<br />

A B <strong>and</strong><br />

AB = BA ⇒ exp A exp B = exp B exp A (67.3)<br />

is true, its converse is not true in general. This is due <strong>to</strong> the fact that the exponential map is local<br />

diffeomorphism, but globally it may or may not be one-<strong>to</strong>-one. Thus there exists a nonzero<br />

solution A for the equation<br />

exp ( A)<br />

= I (67.4)<br />

For example, in the simplest case when V<br />

(65.15) that<br />

is a two-dimensional spce, we can check directly from<br />

exp ⎡0 −θ ⎤ ⎡cosθ −sinθ<br />

⎤<br />

⎢ 0 0 ⎥=<br />

⎢ sin θ cos θ<br />

⎥<br />

⎣ ⎦ ⎣ ⎦<br />

(67.5)<br />

In particular, a possible solution for (67.4) is the matrix<br />

For this solution exp( A ) clearly commutes with ( )<br />

⎡0 − 2π<br />

⎤<br />

A = ⎢<br />

2π<br />

0 ⎥<br />

(67.6)<br />

⎣ ⎦<br />

exp B for all B, even though A may or may not<br />

commute with B. Thus the converse <strong>of</strong> (67.3) does not hold in general.<br />

The main result <strong>of</strong> this section is the following theorem.<br />

Theorem 67.3. H is a maximal connected, continuous Abelian subgroup <strong>of</strong> GL ( V ) if <strong>and</strong> only if<br />

it is the subgroup corresponding <strong>to</strong> a maximal Abelian subalgebra h <strong>of</strong> gl ( V ) .<br />

conditions:<br />

Naturally, a maximal Abelian subalgebra h <strong>of</strong> gl ( V ) is defined by the following two<br />

(i) Any pair <strong>of</strong> elements A <strong>and</strong> B belonging <strong>to</strong> h commute, i.e.,[ AB , ] = 0.<br />

(ii) h is not a proper subset <strong>of</strong> any subalgebra <strong>of</strong> ( V )<br />

gl satisfying condition (i).


488 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

To prove the preceding theorem, we need the following lemma.<br />

Lemma. Let H be an arbitrary connected continuous subgroup <strong>of</strong> GL ( V ) <strong>and</strong> let N be a<br />

neighborhood <strong>of</strong> I in H . Then H is generated by N , i.e., every element X <strong>of</strong> H can be expressed<br />

as a product (not unique)<br />

X= YY<br />

1 2⋅⋅⋅Y k<br />

(67.7)<br />

where<br />

Yi<br />

or<br />

Y belongs <strong>to</strong> N . [The number <strong>of</strong> fac<strong>to</strong>rs k in the representation (67.7) is arbitrary.]<br />

−1<br />

i<br />

Note. Since H is a hypersurface in the inner product space L ( V ; V ) , we can define a<br />

neighborhood system on H simply by the intersection <strong>of</strong> the Euclidean neighborhood system on<br />

L V ; V with H . The <strong>to</strong>pology defined in this way on H is called the induced <strong>to</strong>pology.<br />

( )<br />

To prove the lemma, let H<br />

0<br />

be the subgroup generated by N . Then H<br />

0<br />

is an open set in<br />

H since from (67.7) every point X∈H 0<br />

has a neighborhood LX<br />

( N ) in H . On the other h<strong>and</strong>,<br />

H<br />

0<br />

is also a closed set in H because the complement <strong>of</strong><br />

0<br />

0<br />

H in H is the union <strong>of</strong> ( )<br />

L ,<br />

Y H 0<br />

Y∈<br />

H H which are all open sets in H . As a result H<br />

0<br />

must coincide with H since by hypothesis<br />

H<br />

has only one component.<br />

By virtue <strong>of</strong> the lemma H is Abelian if <strong>and</strong> only if N is an Abelian set. Combining this<br />

remark with Theorem 67.2, <strong>and</strong> using the fact that the exponential map is a local diffeomorphism at<br />

the identity element, we can conclude immediately that H is a maximal connected continuous<br />

Abelian subgroup <strong>of</strong> GL ( V ) if <strong>and</strong> only if h is a maximal Abelian Lie subalgebra <strong>of</strong> gl ( V ) .<br />

This completes the pro<strong>of</strong>.<br />

It should be noticed that on a connected Abelian continuous subgroup <strong>of</strong> GL ( V ) the<br />

Cartan parallelism reduced <strong>to</strong> a Euclidean parallelism. Indeed, since the Lie bracket vanishes<br />

E is also the natural basis <strong>of</strong> a coordinate system<br />

identically on h , any invariant basis { } Γ<br />

{ X Γ } on H . Further, the coordinate map defined by<br />

X≡exp( X Γ E<br />

Γ )<br />

(67.8)<br />

M<br />

is a homomorphism <strong>of</strong> the additive group R with the Abelian group H . This coordinate system<br />

plays the role <strong>of</strong> a local Cartesian coordinate system on a neighborhood <strong>of</strong> the identity element<br />

<strong>of</strong> H . The mapping defined by (67.8) may or may not be one-<strong>to</strong>-one. In the former case H is<br />

M<br />

isomorphic <strong>to</strong> R , in the latter case H is isomorphic <strong>to</strong> a cylinder or a <strong>to</strong>rus <strong>of</strong> dimension M .


Sec. 66 • Maximal Abelian Subgroups, Subalgebras 489<br />

We say that the Cartan parallelism on the Abelian group H is a Euclidean parallelism<br />

because there exists a local Cartesian coordinate system relative <strong>to</strong> which the Cartan symbols<br />

vanish identically. This Euclidean parallelsin on H should not be confused with the Eucliedean<br />

parallelism on the underlying inner product space L ( V ; V ) in which H is a hypersurface. In<br />

genereal, even if H is Abelian, the tangent spaces at different points <strong>of</strong> H are still different<br />

L V ; V . Thus the Euclidean parallelism on H is not the restriction <strong>of</strong> the<br />

subspaces <strong>of</strong> ( )<br />

Euclidean parallelism <strong>of</strong> L ( V ; V ) <strong>to</strong> H .<br />

An example <strong>of</strong> a maximal connected Abelian continuous subgroup <strong>of</strong> GL ( V ) is a<br />

dilatation group defined as follows: Let { e<br />

i<br />

, i=<br />

1,..., N}<br />

be the basis <strong>of</strong>V . Then a linear<br />

transformation X <strong>of</strong> V is a dilatation with axes { e i } if each e<br />

i<br />

is an eigenvec<strong>to</strong>r <strong>of</strong> X <strong>and</strong> the<br />

corresponding eigenvalue is positive. In other words, the component matrix <strong>of</strong> X relative <strong>to</strong> { e i } is<br />

a diagonal matrix with positive diagonal components, say<br />

⎡λ1<br />

⎤<br />

⎢<br />

⋅<br />

⎥<br />

⎢<br />

⎥<br />

i<br />

⎡<br />

⎣X j⎤= ⎦ ⎢ ⋅ ⎥, λi<br />

> 0 i=<br />

1,..., N<br />

(67.9)<br />

⎢<br />

⎥<br />

⎢<br />

⋅<br />

⎥<br />

⎢⎣<br />

λ ⎥<br />

N ⎦<br />

where λi<br />

may or may not be distinct. The dilatation group with axes { e i } is the group <strong>of</strong> all<br />

dilatations X . We leave the pro<strong>of</strong> <strong>of</strong> the fact that a dilatation group is a maximal connected<br />

Abelian continuous subgroup <strong>of</strong> GL ( V ) as an exercise.<br />

Dilatation groups are not the only class <strong>of</strong> maximal Abelian subgroup <strong>of</strong> GL ( V ), <strong>of</strong> course.<br />

For example, when V is three-dimensional we choose a basis { e1, e2,<br />

e3}<br />

forV ; then the subgroup<br />

consisting <strong>of</strong> all linear transformations having component matrix relative <strong>to</strong> { e i } <strong>of</strong> the form<br />

⎡a b 0⎤<br />

⎢<br />

0 a 0<br />

⎥<br />

⎢ ⎥<br />

⎢⎣<br />

bca⎥⎦<br />

with positive a <strong>and</strong> arbitrary b <strong>and</strong> c is also a maximal connected Abelian continuous subgroup <strong>of</strong><br />

GL ( V ). Again, we leave the pro<strong>of</strong> <strong>of</strong> this fact as an exercise.


490 Chap. 12 • CLASSICAL CONTINUOUS GROUP<br />

For inner product spaces <strong>of</strong> lower dimensions a complete classification <strong>of</strong> the Lie<br />

V is known. In such cases a corresponding classification <strong>of</strong><br />

gl<br />

subalgebars <strong>of</strong> the Lie algebra ( )<br />

connected continuous subgroups <strong>of</strong> GL ( V ) can be obtained by using the main result <strong>of</strong> the<br />

preceding section. Then the set <strong>of</strong> maximal connected Abelian continuous subgroups can be<br />

determined completely. These results are beyond the scope <strong>of</strong> this chapter, however.


________________________________________________________________<br />

Chapter 13<br />

INTEGRATION OF FIELDS ON EUCLIDEAN MANIFORDS,<br />

HYPERSURFACES, AND CONTINUOUS GROUPS<br />

In this chapter we consider the theory <strong>of</strong> integration <strong>of</strong> vec<strong>to</strong>r <strong>and</strong> tensor fields defined on<br />

various geometric entities introduced in the preceding chapters. We assume that the reader is<br />

familiar with the basic notion <strong>of</strong> the Riemann integral for functions <strong>of</strong> several real variables.<br />

Since we shall restrict our attention <strong>to</strong> the integration <strong>of</strong> continuous fields only, we do not need<br />

the more general notion <strong>of</strong> the Lebesgue integral.<br />

Section 68. Arc Length, Surface Area, <strong>and</strong> <strong>Vol</strong>ume<br />

Let E be a Euclidean maniforld <strong>and</strong> let λ be a smooth curve in E . Then the tangent <strong>of</strong> λ<br />

is a vec<strong>to</strong>r in the translation space V <strong>of</strong> E defined by<br />

As usual we denote the norm <strong>of</strong> λ by λ ,<br />

( t+Δt) −λ( t)<br />

λ<br />

λ () t = lim<br />

(68.1)<br />

Δ→ t 0 Δt<br />

() t = ⎡ () t ⋅ () t ⎤<br />

1/2<br />

λ ⎣<br />

λ λ ⎦<br />

(68.2)<br />

which is a continuous function <strong>of</strong> t , the parameter <strong>of</strong> λ . Now suppose that λ is defined for t<br />

from <strong>to</strong> .<br />

λ a <strong>and</strong> λ b by<br />

a b Then we define the arc length <strong>of</strong> λ between ( ) ( )<br />

b<br />

a<br />

()<br />

l = ∫ λ t dt<br />

(68.3)<br />

We claim that the arc length possesses the following properties which justify the definition<br />

(68.3).<br />

(i) The arc length depends only on the path <strong>of</strong> λ joining ( a)<br />

λ <strong>and</strong> λ ( b),<br />

independent <strong>of</strong> the choice <strong>of</strong> parameterization on the path.<br />

491


492 Chap. 13 • INTEGRATION OF FIELDS<br />

Indeed, if the path is parameterized by t so that<br />

with<br />

then the tangent vec<strong>to</strong>rs<br />

λ <strong>and</strong> λ are related by<br />

t<br />

( t) = ( t )<br />

λ λ (68.4)<br />

( )<br />

= t t<br />

(68.5)<br />

() t = λ ( t ) dt<br />

λ dt<br />

(68.6)<br />

As a result, we have<br />

∫<br />

b<br />

a<br />

b<br />

() = λ()<br />

λ t dt t dt<br />

(68.7)<br />

∫<br />

a<br />

which proves property (i).<br />

given by<br />

(ii) When the path joining ( a) <strong>and</strong> ( b)<br />

λ λ is a straight line segment, the arc length is<br />

( ) ( )<br />

l = λ b −λ a<br />

(68.8)<br />

This property can be verified by using the parameterization<br />

( ) ( )<br />

( ) ( ) ( )<br />

where t ranges from 0 <strong>to</strong> 1 since ( ) = ( a) ( ) = ( b)<br />

λ t = λ b − λ a t + λ a<br />

(68.9)<br />

λ 0 λ <strong>and</strong> λ 1 λ . In view <strong>of</strong> (68.7), l is given by<br />

1<br />

0<br />

( ) ( ) ( ) ( )<br />

l = ∫ λ b − λ a dt = λ b − λ a<br />

(68.10)<br />

(iii)<br />

<strong>to</strong> b<br />

The arc length integral is additive, i.e., the sum <strong>of</strong> the arc lengths from<br />

b <strong>to</strong> c<br />

λ a <strong>to</strong> λ c .<br />

λ( a) λ( ) <strong>and</strong> from λ( ) λ( ) is equal <strong>to</strong> the arc length from ( ) ( )<br />

Now using property (i), we can parameterize the path <strong>of</strong> λ by the arc length relative <strong>to</strong> a<br />

certain reference point on the path. As usual, we assume that the path is oriented. Then we<br />

assign a positive parameter s <strong>to</strong> a point on the positive side <strong>and</strong> a negative parameter s <strong>to</strong> a<br />

point on the negative side <strong>of</strong> the reference point, the absolute value s being the arc length


Sec. 68 • Arc Length, Surface Area, <strong>Vol</strong>ume 493<br />

between the point <strong>and</strong> the reference point. Hence when the parameter t is positively oriented,<br />

the arc length parameter s is related <strong>to</strong> t by<br />

1<br />

() ()<br />

where λ ( 0)<br />

is chosen as the reference point. From (68.11) we get<br />

s= s t =∫ λ t dt<br />

(68.11)<br />

ds / dt<br />

0<br />

( t)<br />

= λ (68.12)<br />

Substituting this formula in<strong>to</strong> the general transformation rule (68.6), we see that the tangent<br />

vec<strong>to</strong>r relative <strong>to</strong> s is a unit vec<strong>to</strong>r pointing in the positive direction <strong>of</strong> the path, as it should be.<br />

Having defined the concept <strong>of</strong> arc length, we consider next the concept <strong>of</strong> surface area.<br />

For simplicity we begin with the area <strong>of</strong> a two-dimensional smooth surface S in E . As usual,<br />

we can characterize S in terms <strong>of</strong> a pair <strong>of</strong> parameters ( u Γ , Γ= 1,2)<br />

which form a local<br />

coordinate system on S<br />

( u 1 , u<br />

2<br />

)<br />

x∈S ⇔ x = ζ<br />

(68.13)<br />

where ζ is a smooth mapping. We denote the tangent vec<strong>to</strong>r <strong>of</strong> the coordinate curves by<br />

h ≡∂ ζ / ∂u Γ , Γ= 1,2<br />

(68.14)<br />

Γ<br />

Then { h Γ } is a basis <strong>of</strong> the tangent plane S<br />

x<br />

<strong>of</strong> S at any x given by (68.13). We assume that<br />

u Γ is a positive coordinate system. Thus { h }<br />

S is oriented <strong>and</strong> that ( )<br />

Γ<br />

is also positive for S<br />

x<br />

.<br />

Now let U be a domain in S with piecewise smooth boundary. We consider first the<br />

simple case when U can be covered entirely by the coordinate system ( u Γ<br />

) . Then we define the<br />

surface area <strong>of</strong> U by<br />

σ<br />

= ∫∫<br />

−<br />

ζ<br />

1<br />

( U )<br />

( , )<br />

e u u<br />

du du<br />

1 2 1 2<br />

(68.15)<br />

1 2<br />

where ( , )<br />

e u u is defined by<br />

( det[ ]) ( det[ ])<br />

1/2 1/2<br />

e≡ a = a ΓΔ<br />

= h Γ<br />

⋅h Δ<br />

(68.16)


494 Chap. 13 • INTEGRATION OF FIELDS<br />

−1<br />

The double integral in (68.15) is taken over ( )<br />

for points belonging <strong>to</strong> U .<br />

ζ U , which denotes the set <strong>of</strong> coordinates ( u Γ<br />

)<br />

By essentially the same argument as before, we can prove that the surface area has the<br />

following properties which justify the definition (68.15).<br />

(iv) The surface area depends only on the domain U , independent <strong>of</strong> the choice <strong>of</strong><br />

parameterization on U .<br />

To prove this, we note that under a change <strong>of</strong> surface coordinates the integr<strong>and</strong> e <strong>of</strong><br />

(68.15) obeys the transformation rule [cf. (61.4)]<br />

Γ<br />

⎡∂u<br />

⎤<br />

e= e det ⎢ Δ<br />

∂u<br />

⎥<br />

⎣ ⎦<br />

(68.17)<br />

As a result, we have<br />

−<br />

ζ<br />

∫∫<br />

( )<br />

1 −<br />

( U)<br />

ζ 1<br />

( U)<br />

( )<br />

e u , u du du = e u , u du du<br />

∫∫<br />

1 2 1 2 1 2 1 2<br />

(68.18)<br />

which proves property (iv).<br />

(v) When S is a plane <strong>and</strong> U is a square spanned by the vec<strong>to</strong>rs h1 <strong>and</strong> h2at the point<br />

x ∈S<br />

, the surface area <strong>of</strong> U is 0<br />

σ = h1 h<br />

2<br />

(68.19)<br />

The pro<strong>of</strong> is essentially the same as before. We use the parameterization<br />

( 1 ,<br />

2<br />

)<br />

ζ u u = x + u Γ h<br />

Γ<br />

(68.20)<br />

0<br />

From (68.16), e is a constant<br />

e = h1 h<br />

2<br />

(68.21)<br />

−1<br />

<strong>and</strong> from (68.20), ζ ( U ) is the square [ 0,1] [ 0,1]<br />

× . Hence by (68.15) we have<br />

1 1<br />

1 2<br />

∫ 1 2<br />

du du<br />

1 2<br />

0 ∫ h h h h (68.22)<br />

0<br />

σ = =


Sec. 68 • Arc Length, Surface Area, <strong>Vol</strong>ume 495<br />

(vi) The surface area integral is additive in the same sence as (iii).<br />

Like the arc length parameter s on a path, a local coordinate system ( u Γ<br />

) on S is called<br />

1 2<br />

an isochoric coordinate system if the surface area density e( u , u ) is identical <strong>to</strong> 1 for all<br />

( u Γ ) . We can define an isochoric coordinate system ( u Γ<br />

) in terms <strong>of</strong> an arbitrary surface<br />

coordinate system ( u Γ<br />

) in the following way. We put<br />

<strong>and</strong><br />

( , )<br />

1 1 1 2 1<br />

u = u u u ≡ u<br />

(68.23)<br />

2<br />

u<br />

( , ) ( , )<br />

2 2 1 2 1<br />

u u u u e u t dt<br />

0<br />

= ≡∫ (68.24)<br />

where we have assumed that the origin ( 0,0 ) is a point in the domain <strong>of</strong> the coordinate system<br />

( u Γ<br />

) . From (68.23) <strong>and</strong> (68.24) we see that<br />

1 1 2<br />

∂u ∂u ∂u<br />

= 1, = 0, = eu,<br />

u<br />

1 2 2<br />

∂u ∂u ∂u<br />

1 2<br />

( )<br />

(68.25)<br />

As a result, the Jacobian <strong>of</strong> the coordinate transformation is<br />

Γ<br />

⎡∂u<br />

⎤<br />

det ⎢ = e u , u<br />

Δ<br />

∂u<br />

⎥<br />

⎣ ⎦<br />

1 2<br />

( )<br />

(68.26)<br />

which implies immediately the desired result:<br />

1 2<br />

( )<br />

e u , u = 1<br />

(68.27)<br />

by virtue <strong>of</strong> (68.17). From (68.26) the coordinate system ( u Γ ) <strong>and</strong> ( u Γ<br />

)<br />

orientation. Hence if ( u Γ ) is positively oriented, then ( u Γ<br />

)<br />

system on S .<br />

An isochoric coordinate system = ( u Γ<br />

)<br />

from a domain in<br />

are <strong>of</strong> the same<br />

is a positive isochoric coordinate<br />

x ζ in corresponds <strong>to</strong> an isochoric mapping ζ<br />

2<br />

R on<strong>to</strong> the coordinate neighborhood <strong>of</strong> ζ in S . In general, ζ is not<br />

need not be a Euclidean metric. In fact,<br />

the surface metric is Euclidean if <strong>and</strong> only if S is developable. Hence an isometric coordinate<br />

isometric, so that the surface metric a ΓΔ<br />

relative <strong>to</strong> ( u Γ<br />

)


496 Chap. 13 • INTEGRATION OF FIELDS<br />

system (i.e., a rectangular Cartesian coordinate system) generally does not exist on an arbitrary<br />

surface S . But the preceding pro<strong>of</strong> shows that isochoric coordinate systems exist on all S .<br />

So far, we have defined the surface area for any domain U which can be covered by a<br />

2<br />

single surface coordinate system. Now suppose that U is not homeomorphic <strong>to</strong> a domain in R .<br />

Then we decompose U in<strong>to</strong> a collection <strong>of</strong> subdomains, say<br />

U = U ∪U ∪ ∪U<br />

(68.28)<br />

1 2 K<br />

whee the interiors <strong>of</strong> U , , 1<br />

… UK<br />

are mutually disjoint. We assume that each U<br />

a<br />

can be covered by<br />

a surface coordinate system so that the surface area σ ( Ua<br />

) is defined. Then we define σ ( U )<br />

naturally by<br />

( U) = ( U ) + + ( U )<br />

σ σ σ<br />

(68.29)<br />

1 K<br />

While the decomposition (68.28) is not unique, <strong>of</strong> course, by the additive property (vi) <strong>of</strong> the<br />

integral we can verify easily that σ ( U ) is independent <strong>of</strong> the decomposition. Thus the surface<br />

area is well defined.<br />

Having considered the concepts <strong>of</strong> arc length <strong>and</strong> surface area in detail, we can now<br />

extend theidea <strong>to</strong> hypersurfaces in general. Specifically, let S be a hupersurface <strong>of</strong> dimension<br />

M. Then locally S can be represented by<br />

( 1 , M<br />

u u )<br />

x∈S ⇔ x=<br />

ζ …<br />

(68.30)<br />

where ζ is a smooth mapping. We define<br />

<strong>and</strong><br />

h Γ<br />

≡∂ ζ / ∂u<br />

, Γ= 1, … , M<br />

(68.31)<br />

Γ<br />

≡ h ⋅ h (68.32)<br />

a ΓΔ Γ Δ<br />

Then the { h Γ } span the tangent space S<br />

x<br />

<strong>and</strong> the a ΓΔ<br />

define the induced metric on S<br />

x<br />

. We<br />

define the surface area density e by the same formula (68.16) except that e is now a smooth<br />

1 , M<br />

u u<br />

a ΓΔ<br />

is also M × M.<br />

function <strong>of</strong> the M variables ( )<br />

… , <strong>and</strong> the matrix [ ]<br />

Now let U be a domain with piecewise smooth boundary in S , <strong>and</strong> assume that U can<br />

be covered by a single surface coordinate system. Then we define the surface area <strong>of</strong> U by


Sec. 68 • Arc Length, Surface Area, <strong>Vol</strong>ume 497<br />

∫<br />

σ = ⋅⋅⋅<br />

−<br />

ζ 1<br />

( U )<br />

∫<br />

M<br />

( , , )<br />

… (68.33)<br />

1 1 M<br />

e u u du du<br />

By the same argument as before, σ has the following two properties.<br />

(vii) The surface area is additive <strong>and</strong> independent <strong>of</strong> the choice <strong>of</strong> surface coordinate<br />

system.<br />

(viii) The surface area <strong>of</strong> an M-dimensional cube with sides h , 1<br />

… h is<br />

, M<br />

σ = h<br />

h<br />

(68.34)<br />

1 M<br />

More generally, when U cannot be covered by a single coordinate system, we decompose U by<br />

(68.28) <strong>and</strong> define σ ( U ) by (68.29).<br />

We can extend the notion <strong>of</strong> an isochoric coordinate system <strong>to</strong> a hypersurface in general.<br />

To construct an isochoric coordinate system ( u Γ<br />

), we begin with an arbitrary coordinate system<br />

( u Γ<br />

). Then we put<br />

Γ Γ 1 M Γ<br />

u = u u ,…, u = u , Γ= 1, …, M −1<br />

(68.35)<br />

( )<br />

M<br />

( 1 u<br />

, , ) ( 1 −<br />

= ≡∫ , , 1 , )<br />

u u u … u e u … u t dt<br />

(68.36)<br />

M M M M<br />

0<br />

From (68.35) <strong>and</strong> (68.36) we get<br />

Γ<br />

⎡∂u<br />

⎤ 1 M<br />

det ⎢ = e( u , u<br />

Δ<br />

)<br />

∂u<br />

⎥ … (68.37)<br />

⎣ ⎦<br />

As a result, the coordinate system ( u Γ<br />

)<br />

is isochoric since<br />

Finally, when M<br />

M<br />

M<br />

( ) ( )<br />

e u ,…, u = e u ,…, u<br />

1<br />

Γ Δ<br />

det ⎡∂u<br />

/ ∂u<br />

⎤<br />

=<br />

(68.38)<br />

⎣ ⎦<br />

1 1 1<br />

= N , S is nothing but a domain in .<br />

becomes an arbitrary local coordinate system in ,<br />

i<br />

volume relative <strong>to</strong> ( u ). The integral<br />

E In this case ( 1 , , N<br />

u … u )<br />

E <strong>and</strong> ( 1 , , N<br />

eu u )<br />

… is just the Euclidean


498 Chap. 13 • INTEGRATION OF FIELDS<br />

υ ≡<br />

∫<br />

⋅⋅⋅<br />

−<br />

ζ 1<br />

( U )<br />

∫<br />

M<br />

( , , )<br />

… (68.39)<br />

1 1 M<br />

eu u du du<br />

now defines the Euclidean volume <strong>of</strong> the domain U .


Sec. 69 • Vec<strong>to</strong>r <strong>and</strong> Tensor Fields 499<br />

Section 69. Integration <strong>of</strong> Vec<strong>to</strong>r Fields <strong>and</strong> Tensor Fields<br />

In the preceding section we have defined the concept <strong>of</strong> surface area for an arbitrary M-<br />

dimensional hyper surface imbedded in an N-dimensional Euclidean manifold E . When M = 1,<br />

the surface reduces <strong>to</strong> a path <strong>and</strong> the surface area becomes the arc length, while in the case M=N<br />

the surface corresponds <strong>to</strong> a domain in E , <strong>and</strong> the surface area becomes the volume. In this<br />

section we shall define the integrals <strong>of</strong> various fields relative <strong>to</strong> the surface area <strong>of</strong> an arbitrary<br />

hyper surface S . We begin with the integral <strong>of</strong> a continuous function f defined on S .<br />

As before, we assume that S is oriented <strong>and</strong> U is a domain in S with piecewise<br />

smoothboundary. We consider first the simple case when U can be covered by a single surface<br />

coordinate system ( 1 M<br />

x = ζ u ,… u ). We choose ( u Γ<br />

) <strong>to</strong> be positively oriented, <strong>of</strong> course. Under<br />

these assumptions, we define the integral <strong>of</strong> f on U by<br />

∫ ∫ ∫<br />

(69.1)<br />

1 M<br />

f dσ<br />

≡ fe du du<br />

U<br />

−<br />

ζ 1<br />

( U )<br />

where the function f on the right-h<strong>and</strong> side denotes the representation <strong>of</strong> f in terms <strong>of</strong> the<br />

surface coordinates ( u Γ ) :<br />

( ) ( 1 M<br />

= , , )<br />

f x f u … u<br />

(69.2)<br />

where<br />

( 1 , , M<br />

u u )<br />

x = ζ … (69.3)<br />

It is unders<strong>to</strong>od that the multiple integral in (69.1) is taken over the positive orientation on<br />

−1<br />

M<br />

ζ U in R .<br />

( )<br />

By the same argument as in the preceding section, we see that the integral possesses the<br />

following properties.<br />

(i) When f is identical <strong>to</strong> 1 the integral <strong>of</strong> f is just the surface area <strong>of</strong> U , namely<br />

( ) d<br />

σ U = ∫ σ<br />

(69.4)<br />

U<br />

(ii) The integral <strong>of</strong> f is independent on the choice <strong>of</strong> the coordinate system ( u Γ<br />

)<br />

additive with respect <strong>to</strong> its domain.<br />

(iii) The integral is a linear function <strong>of</strong> the integr<strong>and</strong> in the sense that<br />

<strong>and</strong> is


500 Chap. 13 • INTEGRATION OF FIELDS<br />

( )<br />

∫ ∫ ∫<br />

α f + α f dσ = α f dσ + α f dσ<br />

1 1 2 2 1 1 2 2<br />

U U U<br />

(69.5)<br />

for all constants α1 <strong>and</strong> α<br />

2.<br />

Also, the integral is bounded by<br />

( ) min ≤ ≤ ( )<br />

∫ U<br />

σ U f fdσ σ U max f<br />

(69.6)<br />

U<br />

U<br />

where the extrema <strong>of</strong> f are taken over the domain U .<br />

Property (iii) is a st<strong>and</strong>ard result <strong>of</strong> multiple integrals in calculus, so by virtue <strong>of</strong> the<br />

definition (69.1) the same is valid for the integral <strong>of</strong> f .<br />

As before, if U cannot be covered by a single coordinate system, then we decompose U<br />

by (68.28) <strong>and</strong> define the integral <strong>of</strong> f over U by<br />

∫ ∫ ∫<br />

f dσ = f dσ + + f dσ<br />

(69.7)<br />

U U1 UK<br />

By property (ii) we can verify easily that the integral is independent <strong>of</strong> the decomposition.<br />

Having defined the integral <strong>of</strong> a scalar field, we define next the integral <strong>of</strong> a vec<strong>to</strong>r field.<br />

Let v be a continuous vec<strong>to</strong>r field on S , i.e.,<br />

v : S →V (69.8)<br />

where V is the translation space <strong>of</strong> the underlying Euclidean manifold E . Generally the values<br />

<strong>of</strong> v may or may not be tangent <strong>to</strong> S . We choose an arbitrary Cartesian coordinate system with<br />

natural basis { e<br />

i}.<br />

Then v can be represented by<br />

1<br />

where ( )<br />

i<br />

( ) = υ ( )<br />

v x x e (69.9)<br />

υ x , i = 1, …, N,<br />

are continuous scalar fields on S . We define the integral <strong>of</strong> v by<br />

i<br />

∫<br />

U<br />

i<br />

( d )<br />

∫<br />

vdσ ≡ υ σ e<br />

U<br />

i<br />

(69.10)<br />

Clearly the integral is independent <strong>of</strong> the choice <strong>of</strong> the basis { e<br />

i}.<br />

More generally if A is a tensor field on S having the representation


Sec. 69 • Vec<strong>to</strong>r <strong>and</strong> Tensor Fields 501<br />

where { i<br />

}<br />

… r<br />

( ) A ( )<br />

A x = x e ⊗ ⊗e ⊗e ⊗ ⊗e<br />

i1 i j1<br />

js<br />

j1… js<br />

i<br />

<br />

1<br />

i<br />

(69.11)<br />

r<br />

e denotes the reciprocal basis <strong>of</strong> { e }, then we define<br />

i<br />

i1…<br />

ir<br />

( σ<br />

s<br />

)<br />

js<br />

∫ A dσ<br />

≡ ∫ A<br />

j1 j<br />

d ei<br />

⊗ ⊗e<br />

U<br />

U<br />

1<br />

…<br />

(69.12)<br />

Again the integral is independent <strong>of</strong> the choice <strong>of</strong> the basis { e<br />

i}.<br />

The integrals defined by (69.10) <strong>and</strong> (69.12) possess the same tensorial order as the<br />

integr<strong>and</strong>. The fact that a Cartesian coordinate system is used in (69.10) <strong>and</strong> (69.12) reflects<br />

clearly the crucial dependence <strong>of</strong> the integral on the Euclidean parallelism <strong>of</strong> E . Without the<br />

Euclidean parallelism it is generally impossible <strong>to</strong> add vec<strong>to</strong>rs or tensors at different points <strong>of</strong> the<br />

domain. Then an integral is also meaningless. For example, if we suppress the Euclidean<br />

parallelism on the underlying Euclidean manifold E , then the tangential vec<strong>to</strong>rs or tensors at<br />

different points <strong>of</strong> a hyper surface S generally do not belong <strong>to</strong> the same tangent space or tensor<br />

space. As a result, it is generally impossible <strong>to</strong> “sum” the values <strong>of</strong> a tangential field <strong>to</strong> obtain an<br />

integral without the use <strong>of</strong> some kind <strong>of</strong> path-independent parallelism. The Euclidean<br />

parallelism is just one example <strong>of</strong> such parallelisms. Another example is the Cartan parallelism<br />

on a continuous group defined in the preceding chapter. We shall consider integrals relative <strong>to</strong><br />

the Cartan parallelism in Section 72.<br />

In view <strong>of</strong> (69.10) <strong>and</strong> (69.12) we see that the integral <strong>of</strong> a vec<strong>to</strong>r field or a tensor field<br />

possesses the following properties.<br />

(iv) The integral is linear with respect <strong>to</strong> the integr<strong>and</strong>.<br />

(v) The integral is bounded by<br />

∫<br />

U<br />

vdσ<br />

≤<br />

∫<br />

U<br />

v<br />

dσ<br />

(69.13)<br />

<strong>and</strong> similarly<br />

∫<br />

U<br />

Adσ<br />

≤<br />

∫<br />

U<br />

A<br />

dσ<br />

(69.14)<br />

where the norm <strong>of</strong> a vec<strong>to</strong>r or a tensor is defined as usual by the inner product <strong>of</strong> V . Then it<br />

follows from (69.6) that<br />

<strong>and</strong><br />

∫ v dσ ≤ σ ( U )max v<br />

(69.15)<br />

U<br />

U


502 Chap. 13 • INTEGRATION OF FIELDS<br />

∫ A dσ ≤ σ( U )max A<br />

(69.16)<br />

U<br />

U<br />

However, it does not follow from (69.6), <strong>and</strong> in fact it is not true, that σ ( ) min<br />

bound for thenorm <strong>of</strong> the integral <strong>of</strong> v .<br />

U v is a lower<br />

U


Sec. 70 • Integration <strong>of</strong> Differential Forms 503<br />

Section 70. Integration <strong>of</strong> Differential Forms<br />

Integration with respect <strong>to</strong> the surface area density <strong>of</strong> a hyper surface is a special case <strong>of</strong><br />

a more general integration <strong>of</strong> differential forms. As remarked in Section 68, the transformation<br />

rule (68.17) <strong>of</strong> the surface area density is the basic condition which implies the important<br />

property that the surface area integral is independent <strong>of</strong> the choice <strong>of</strong> the surface coordinate<br />

system. Since the transformation rule (68.17) is essentially the same as that <strong>of</strong> the strict<br />

components <strong>of</strong> certain differential forms, we can extend the operation <strong>of</strong> integration <strong>to</strong> those<br />

forms also. This extension is the main result <strong>of</strong> this section.<br />

We begin with the simple notion <strong>of</strong> a differential N-form Z on E . By definition, Z is a<br />

completely skew-symmetric covariant tensor field <strong>of</strong> order N. Thus relative <strong>to</strong> any coordinate<br />

i<br />

system ( u ), Z has the representation<br />

Z h h h h<br />

i1<br />

iN<br />

1<br />

= Zi 1…<br />

i<br />

⊗ ⊗ = Z<br />

N<br />

1…<br />

N<br />

∧ ∧<br />

h<br />

<br />

1<br />

= z ∧ ∧<br />

h<br />

N<br />

N<br />

(70.1)<br />

where z is called the relative scalar or the density <strong>of</strong> Z . As we have shown in Section 39, the<br />

transformation rule for z is<br />

i<br />

⎡ ∂u<br />

⎤<br />

z = zdet<br />

⎢ j<br />

∂u<br />

⎥<br />

⎣ ⎦<br />

(70.2)<br />

i<br />

i<br />

This formula is comparable <strong>to</strong> (68.17). In fact if we require that ( u ) <strong>and</strong> ( )<br />

u both be positively<br />

oriented, then (70.2) can be regarded as a special case <strong>of</strong> (68.17) with M = N.<br />

As a result, we<br />

can define the integral <strong>of</strong> Z over a domain U in E by<br />

N<br />

∫ Z ≡∫ 1<br />

( , , )<br />

( )<br />

∫<br />

U<br />

1 1 N<br />

z u … u du du<br />

(70.3)<br />

ζ − U<br />

i<br />

<strong>and</strong> the integral is independent <strong>of</strong> the choice <strong>of</strong> the (positive) coordinate system x = ζ ( u ).<br />

Notice that in this definition the Euclidean metric <strong>and</strong> the Euclidean volume density e<br />

are not used at all. In fact, (68.39) can be regarded as a special case <strong>of</strong> (70.3) when Z reduces <strong>to</strong><br />

the Euclidean volume tensor<br />

= h ∧ ∧h<br />

1 N<br />

E e (70.4)


504 Chap. 13 • INTEGRATION OF FIELDS<br />

Here we have assumed that the coordinate system is positively oriented; otherwise, a negative<br />

sign should be inserted on the right h<strong>and</strong> side since the volume density e as defined by (68.16) is<br />

always positive. Hence, unlike the volume integral, the integral <strong>of</strong> a differential N-form Z is<br />

defined only if the underlying space E is oriented. Other than this aspect, the integral <strong>of</strong> Z <strong>and</strong><br />

the volume integral have essentially the same properties since they both are defined by an<br />

invariant N-tuple integral over the coordinates.<br />

Now more generally let S be an oriented hypersurface in E <strong>of</strong> dimension M , <strong>and</strong><br />

suppose Z is a tangential differential M-form on S . As before, we choose a positive surface<br />

u Γ on S <strong>and</strong> represent Z by<br />

coordinate system ( )<br />

1 M<br />

Z= z h ∧ ∧h<br />

(70.5)<br />

where z is a function <strong>of</strong> ( 1 , , M<br />

u … u ) where { h Γ<br />

} is the natural basis reciprocal <strong>to</strong> { h<br />

Γ<br />

}<br />

the transformation rule for z is<br />

. Then<br />

Γ<br />

⎡∂u<br />

⎤<br />

z = zdet<br />

⎢ Δ<br />

∂u<br />

⎥<br />

⎣ ⎦<br />

(70.6)<br />

As a result, we can define the integral <strong>of</strong> Z over a domain U in S by<br />

M<br />

∫ ≡<br />

1<br />

( , , )<br />

U ∫<br />

( )<br />

∫<br />

1 1 M<br />

Z z u … u du du<br />

(70.7)<br />

ζ − U<br />

<strong>and</strong> the integral is independen<strong>to</strong>f the choice <strong>of</strong> the positive surface coordinate system ( u Γ<br />

).<br />

the same remark as before, we can regard (70.7) as a generalization <strong>of</strong> (68.33).<br />

By<br />

The definition (70.7) is valid for any tangential M-form Z on S . In this definition the<br />

surface metric <strong>and</strong> thesurface area density are not used. The fact that Z is a tangential field on<br />

S is not essential in the definition. Indeed, if Z is an arbitrary skew-symmetric spatial<br />

covariant tensor<strong>of</strong> order M on S , then we define the density <strong>of</strong> Z on S relative <strong>to</strong> ( u Γ<br />

)<br />

by<br />

( )<br />

simply<br />

z = Z h , , 1<br />

… hM<br />

(70.8)<br />

Using this density, we define the integral <strong>of</strong> Z again by (70.7). Of course, the formula (70.8) is<br />

valid for a tangential M-form Z also, since it merely represents the strict component <strong>of</strong> the<br />

tangential projection <strong>of</strong> Z .<br />

This remark can be further generalized in the following situation: Suppose that S is a<br />

hyper surface contained in another hypersurface S<br />

0<br />

in E , <strong>and</strong> let Z be a tangential M-form on


Sec. 70 • Integration <strong>of</strong> Differential Forms 505<br />

S<br />

0<br />

. Then Z gives rise <strong>to</strong> a density on S by the same formula (70.8), <strong>and</strong> the integral <strong>of</strong> Z over<br />

any domain U in S is defined by (70.7). Algebraically, this remark is a consequence <strong>of</strong> the<br />

simple fact that a skew-symmetric tensor over the tangent space <strong>of</strong> S<br />

0<br />

gives rise <strong>to</strong> a unique<br />

skew-symmetric tensor over the tangent space <strong>of</strong> S , since the latter tangent space is a subspace<br />

<strong>of</strong> the former one.<br />

It should be noted, however, that the integral <strong>of</strong> Z is defined over S only if the order <strong>of</strong><br />

Z coincides with the dimension <strong>of</strong> S . Further, the value <strong>of</strong> theintegral is always a scalar, not a<br />

vec<strong>to</strong>r or a tensor as in the preceding section. We can regard the intetgral <strong>of</strong> a vec<strong>to</strong>r field or a<br />

tensor field as a special case <strong>of</strong> the integral <strong>of</strong> a differential form only when the fields are<br />

represented interms <strong>of</strong> their Cartesian components as shown in (69.9) <strong>and</strong> (69.11).<br />

An important special case <strong>of</strong> the integral <strong>of</strong> a differential form is the line integral in<br />

classical vec<strong>to</strong>r analysis. In this case S reduces <strong>to</strong> an oriented path λ , <strong>and</strong> Z is a 1-form w .<br />

When the Euclidean metric on E is used, w corresponds simply <strong>to</strong> a (spatial or tangential) vec<strong>to</strong>r<br />

field on λ . Now using any positive parameter t on λ , we obtain from (70.7)<br />

∫<br />

() t ()<br />

w<br />

b<br />

= ∫ w ⋅ λ t dt<br />

(70.9)<br />

λ a<br />

Here we have used the fact that for an inner product space the isomorphism <strong>of</strong> a vec<strong>to</strong>r <strong>and</strong> a<br />

covec<strong>to</strong>r is given by<br />

w, λ = w ⋅ λ (70.10)<br />

The tangent vec<strong>to</strong>r λ playes the role <strong>of</strong> the natural basis vec<strong>to</strong>r h<br />

1<br />

associated with the parameter<br />

t, <strong>and</strong> (70.10) is just the special case <strong>of</strong> (70.8) when M = 1.<br />

The reader should verify directly that the right-h<strong>and</strong> side <strong>of</strong> (70.9) is independent <strong>of</strong> the<br />

choice <strong>of</strong> the (positive) parameterization t on λ . By virtue <strong>of</strong> this remark, (70.9) is also written<br />

as<br />

in the classical theory.<br />

∫<br />

λ<br />

∫<br />

w = w⋅d<br />

λ (70.11)<br />

Similarly when N = 3 <strong>and</strong> M = 2 , a 2-form Z is also representable by a vec<strong>to</strong>r field w ,<br />

namely<br />

λ<br />

( h , h ) = w⋅ ( h × h )<br />

Z (70.12)<br />

1 2 1 2


506 Chap. 13 • INTEGRATION OF FIELDS<br />

Then the integral <strong>of</strong> Z over a domain U in a two-dimensional oriented surface S is given by<br />

−<br />

ζ 1<br />

( U )<br />

( )<br />

∫ ∫∫ ∫<br />

U<br />

Z = w⋅ h × h ≡ w⋅<br />

1 2<br />

1 2<br />

du du d<br />

U<br />

σ<br />

(70.13)<br />

where dσ is the positive area element <strong>of</strong> S defined by<br />

1 2<br />

( ) du du<br />

dσ ≡ h × h (70.14)<br />

1 2<br />

The reader will verify easily that the right-h<strong>and</strong> side <strong>of</strong> (70.14) can be rewritten as<br />

d<br />

1 2<br />

σ = en du du<br />

(70.15)<br />

where e is the surface area density on S defined by (68.16), <strong>and</strong> where n is the positive unit<br />

normal <strong>of</strong> S defined by<br />

h × h 1<br />

n h h<br />

1 2<br />

= =<br />

1×<br />

2<br />

h1×<br />

h2<br />

e<br />

(70.16)<br />

Substituting (70.15) in<strong>to</strong> (70.13), we see that the integral <strong>of</strong> Z can be represented by<br />

∫<br />

U<br />

ζ −<br />

1<br />

( U )<br />

( )<br />

Z = ∫∫ w⋅n<br />

edu du<br />

1 2<br />

(70.17)<br />

which shows clearly that the integral is independent <strong>of</strong> the choice <strong>of</strong> the (positive) surface<br />

coordinate system ( u Γ<br />

).<br />

Since the multiple<strong>of</strong> an M-form Z by a scalar field f remains an M-form, we can define<br />

the integral <strong>of</strong> f with respect <strong>to</strong> Z simply as the integral <strong>of</strong> f Z . Using a Cartesian coordinate<br />

representation, we can extend this operation <strong>to</strong> integrals <strong>of</strong> a vec<strong>to</strong>r field or a tensor field relative<br />

<strong>to</strong> a differential form. The integrals defined in the preceding section are special cases <strong>of</strong> this<br />

general operation when the differential forms are the Euclidean surface area densities induced<br />

bythe Euclidean metric on the underlying space E .


Sec. 71 • Generalized S<strong>to</strong>kes’ Theorem 507<br />

Section 71. Generalized S<strong>to</strong>kes’ Theorem<br />

In Section 51 we have defined the operation <strong>of</strong> exterior derivative on differential forms<br />

on E . Since this operation does not depend on the Euclidean metric <strong>and</strong> the Euclidean<br />

parallelism, it can be defined also for tangential differential forms on a hyper surface, as we have<br />

remarked in Section 55. Specifically, if Z is a K-form on an M-dimensional hypersurface S , we<br />

choose a surface coordinate system ( u Γ<br />

)<br />

<strong>and</strong> represent Z by<br />

Z<br />

∑ h h<br />

Γ1<br />

ΓK<br />

= ZΓ<br />

1Γ<br />

∧ ∧<br />

(71.1)<br />

K<br />

Γ< 1


508 Chap. 13 • INTEGRATION OF FIELDS<br />

where { g α , α = 1, …,P}<br />

denotes the natural basis <strong>of</strong> ( )<br />

tangential projection <strong>of</strong> W on S , i.e.,<br />

y α on S<br />

0<br />

, <strong>and</strong> suppose that Z is the<br />

α α<br />

∂y<br />

∂y<br />

Z = ∧ ∧<br />

(71.6)<br />

∑<br />

1<br />

K<br />

Γ1<br />

ΓK<br />

Wα<br />

1 αK<br />

Γ1<br />

ΓK<br />

u u<br />

1< < α ∂ ∂<br />

h h<br />

<br />

K<br />

α<br />

where { h Γ , Γ= 1, …,M<br />

} denotes thenatural basis <strong>of</strong> ( u Γ<br />

)<br />

on S . Then the exterior derivatives<br />

<strong>of</strong> Z coincides with the tangential projection <strong>of</strong> the exterior derivative <strong>of</strong> W . In other words,<br />

theoperation <strong>of</strong> exterior derivative commutes with the operation <strong>of</strong> tangentialprojection.<br />

We can prove this lemma by direct calculation <strong>of</strong> dW <strong>and</strong> dZ . From (71.5) <strong>and</strong> (71.2)<br />

dW is given by<br />

dW<br />

∂W<br />

∑ g g g<br />

β<br />

∂y<br />

α1αK<br />

β α1<br />

αK<br />

= ∧ ∧ ∧<br />

(71.7)<br />

α<br />

α<br />

1< <<br />

K<br />

Hence its tangential projection on S is<br />

β α α<br />

∂Wα<br />

∂y ∂y ∂y<br />

∑ h h h<br />

∂y ∂u ∂u ∂u<br />

1< <<br />

αK<br />

α<br />

1<br />

K<br />

1αK<br />

Δ Γ1<br />

ΓK<br />

∧ ∧ ∧<br />

(71.8)<br />

β Δ Γ1<br />

ΓK<br />

Similarly, from (71.6) <strong>and</strong> (71.7), dZ is given by<br />

α1<br />

αK<br />

∂ ⎛ ∂y<br />

∂y<br />

⎞ Δ Γ1<br />

dZ<br />

= ∑ Δ ⎜Wα1 < < α<br />

<br />

K Γ1<br />

ΓK<br />

⎟ h ∧h ∧∧h<br />

α u u u<br />

1< <<br />

α ∂<br />

K ⎝ ∂ ∂ ⎠<br />

β α1<br />

α<br />

∂W<br />

K<br />

α y y y<br />

1…<br />

α ∂ ∂ ∂<br />

K<br />

Δ Γ1<br />

ΓK<br />

= ∑<br />

h ∧h ∧∧h<br />

β Δ Γ1<br />

ΓK<br />

∂y ∂u ∂u ∂u<br />

1< <<br />

αK<br />

α<br />

ΓK<br />

(71.9)<br />

where we have used the skew symmetry <strong>of</strong> the exterior product <strong>and</strong> the symmetry <strong>of</strong> the second<br />

2 α Γ Δ<br />

derivative ∂ y / ∂u ∂ u with respect <strong>to</strong> Γ <strong>and</strong> Δ . Comparing (71.9) with (71.8), we have<br />

completed the pro<strong>of</strong> <strong>of</strong> the lemma.<br />

Now we are ready <strong>to</strong> present the main result <strong>of</strong> this section.<br />

Generalized S<strong>to</strong>kes’ Theorem. Let S <strong>and</strong> S<br />

0<br />

be hypersurfaces as defined in the preceding<br />

lemma <strong>and</strong> suppose that U is an oriented domain in S with piecewise smooth boundary ∂U .<br />

(We orient the boundary ∂U as usual by requiring the outward normal <strong>of</strong> ∂U be the positive<br />

normal.) Then for any tangential ( M −1)<br />

-form Z on S<br />

0<br />

we have


Sec. 71 • Generalized S<strong>to</strong>kes’ Theorem 509<br />

∫<br />

∂U<br />

∫<br />

Z = dZ (71.10)<br />

Before proving this theorem, we remark first that other than the condition<br />

U<br />

S ⊂ S (71.11)<br />

0<br />

The hypersurface S<br />

0<br />

is entirely arbitrary. In application we <strong>of</strong>ten take S = 0<br />

E but this choice is<br />

not necessary. Second, by virtue <strong>of</strong> the preceding lemma it suffices <strong>to</strong> prove (71.10) for<br />

tangential ( M −1)<br />

−forms Z on S only. In other words, we can choose S<br />

0<br />

<strong>to</strong> be the same as<br />

S without loss <strong>of</strong> generality. Indeed, as explained in the preceding section, the integrals in<br />

(71.10) are equal <strong>to</strong> those <strong>of</strong> tangential projection <strong>of</strong> the forms Z <strong>and</strong> dZ on S . Then by virtue<br />

<strong>of</strong> the lemma the formula (71.10) amounts <strong>to</strong> nothing but a formula for tangential forms on S .<br />

Before proving the formula (71.10) in general, we consider first the simplest special case<br />

when Z is a 1-form <strong>and</strong> S is a two-dimensional surface. This case corresponds <strong>to</strong> the S<strong>to</strong>kes<br />

formula in classical vec<strong>to</strong>r analysis. As usual, we denote a 1-form by w since it is merely a<br />

covariant vec<strong>to</strong>r field on S . Let the component form <strong>of</strong> w in ( u Γ<br />

)<br />

be<br />

w = w h Γ<br />

(71.12)<br />

Γ<br />

where Γ is summed from 1 <strong>to</strong> 2. From (71.2) the exterior derivative <strong>of</strong> w is a 2-form<br />

d<br />

∂w ⎛∂w ∂w<br />

⎞<br />

= ∧ = ⎜ − ⎟ ∧<br />

∂u ⎝ ∂u ∂u<br />

⎠<br />

Γ Δ Γ<br />

2 1 1 2<br />

w h h h h (71.13)<br />

Δ<br />

1 2<br />

Thus (71.10) reduces <strong>to</strong><br />

( )<br />

coordinate system (<br />

1 2)<br />

where 1 () t ,<br />

2<br />

() t<br />

∫<br />

−1<br />

ζ ( ∂U<br />

)<br />

⎛∂w<br />

∂w<br />

⎞<br />

w dt du du<br />

Γ<br />

2 1 1 2<br />

Γλ<br />

= ∫∫ ⎜ −<br />

1 2 ⎟<br />

(71.14)<br />

−1<br />

⎝ ∂u<br />

∂u<br />

ζ<br />

⎠<br />

( U )<br />

λ λ denotes the coordinates <strong>of</strong> the oriented boundary curve ∂U in the<br />

entirely arbitrary.<br />

u , u , the parameterization t on ∂U being positively oriented but otherwise<br />

Now since the integrals in (71.14) are independent <strong>of</strong> the choice <strong>of</strong> positive coordinate<br />

−1<br />

2<br />

0,1 × 0,1 in R .<br />

system, for simplicity we consider first thecase when ζ ( U ) is the square [ ] [ ]<br />

1 2 1 2<br />

Naturally, we use the parameters u , u ,1 − u , <strong>and</strong>1− u on the boundary segments


510 Chap. 13 • INTEGRATION OF FIELDS<br />

( 0,0) →( 1,0 ),( 1,0) →( 1,1 ),( 1,1) →( 0,1 ), <strong>and</strong> ( 0,1) →( 0,0 ),<br />

respectively. Relative <strong>to</strong> this<br />

parameterization on ∂U the left-h<strong>and</strong> side <strong>of</strong> (71.14) reduces <strong>to</strong><br />

∫<br />

1 1 2 2<br />

( ,0) + ∫ ( 1, )<br />

1 1<br />

−∫<br />

1( ,1) −∫<br />

2( 0, )<br />

1 1<br />

w u du w u du<br />

1 2<br />

0 0<br />

w u du w u du<br />

1 1 2 2<br />

0 0<br />

(71.15)<br />

1 2<br />

Similarly, relative <strong>to</strong> the surface coordinate system ( , )<br />

reduces <strong>to</strong><br />

⎛∂w ∂w ⎞ du du<br />

⎝ ∂u<br />

∂u<br />

⎠<br />

u u the right-h<strong>and</strong> side <strong>of</strong> (71.14)<br />

∫∫<br />

1 1<br />

2 1 1 2<br />

⎜ −<br />

0 0 1 2 ⎟<br />

(71.16)<br />

1 2<br />

which may be integrated by parts once with respect <strong>to</strong> one <strong>of</strong> the two variables ( , )<br />

u u <strong>and</strong> the<br />

result is precisely the sameas (71.15). Thus (71.14) is proved in this simple case.<br />

In general U may not be homeomorphic <strong>to</strong> the square [ 0,1] × [ 0,1 ],<br />

<strong>of</strong> course. Then we<br />

decompose U as before by (68.28) <strong>and</strong> we assume that each U , a = a<br />

1, …, K , can be represented<br />

by the range<br />

a<br />

a<br />

([ 0,1] [ 0,1]<br />

)<br />

U = ζ ×<br />

(71.17)<br />

By the result for the simple case we then have<br />

∫<br />

∂Ua<br />

∫<br />

w = dw, a = 1, …,<br />

K<br />

(71.18)<br />

∂Ua<br />

Now adding (71.18) with respect <strong>to</strong> a <strong>and</strong> observing the fact that all common boundaries <strong>of</strong><br />

pairs <strong>of</strong> U , 1<br />

… U are oriented oppositely as shown in Figure 10, we obtain<br />

, K<br />

∫<br />

∂U<br />

w =<br />

∫<br />

U<br />

dw<br />

(71.19)<br />

which is the special case <strong>of</strong> (71.10) when M = 2.


Sec. 71 • Generalized S<strong>to</strong>kes’ Theorem 511<br />

U<br />

3<br />

U<br />

5<br />

U<br />

7<br />

U<br />

2<br />

U<br />

1<br />

U<br />

4<br />

U<br />

6<br />

U<br />

8<br />

Figure 10<br />

The pro<strong>of</strong> <strong>of</strong> (71.10) for the general case with an arbitrary M is essentially the same as the<br />

pro<strong>of</strong> <strong>of</strong> the preceding special case. To illustrate the similarity <strong>of</strong> the pro<strong>of</strong>s we consider next the<br />

case M = 3. In this case Z is a 2-form on a 3-dimensional hypersurface S . Let<br />

( u Γ , Γ= 1, 2, 3)<br />

be a positive surface coordinate system on S as usual. Then Z can be<br />

represented by<br />

∑<br />

Z= Z h ∧h<br />

Γ Δ<br />

ΓΔ<br />

Γ


512 Chap. 13 • INTEGRATION OF FIELDS<br />

⎛ ∂ Z ∂ Z ∂ Z<br />

⎝ ∂u ∂u ∂u<br />

⎞<br />

⎠<br />

∫∫∫<br />

1 1 1<br />

12 13 23 1 2 3<br />

⎜ − + du du du<br />

0 0 0 3 2 1 ⎟<br />

(71.23)<br />

u , u , u . The<br />

1 2 3<br />

which may be integrated by parts once with respect <strong>to</strong> one <strong>of</strong> the three variables ( )<br />

result consists <strong>of</strong> six terms:<br />

1 1 1 1<br />

∫∫ Z12 ( u , u ,1) du du −<br />

12 ( )<br />

0 0 ∫∫ Z u , u ,0 du du<br />

0 0<br />

1 1 1 1<br />

− ∫∫ 13 ( ,1, ) +<br />

13 ( , 0, )<br />

0 0 ∫∫ 0 0<br />

1 1 1 1<br />

+ ∫∫ 23 ( 1, , ) −∫∫<br />

23 ( 0, , )<br />

1 2 1 2 1 2 1 2<br />

Z u u du du Z u u du du<br />

1 3 1 3 1 3 1 3<br />

Z u u du du Z u u du du<br />

2 3 2 3 2 3 2 3<br />

0 0 0 0<br />

(71.24)<br />

which are precisely the representations <strong>of</strong> the left-h<strong>and</strong> side <strong>of</strong> (71.10) on the six faces <strong>of</strong> the<br />

cube with an appropriate orientation on each face. Thus (71.10) is proved when (71.22) holds.<br />

In general, if U cannot be represented by (71.22), then we decompose U as before by<br />

(68.28), <strong>and</strong> we assume that each U<br />

a<br />

may be represented by<br />

a<br />

a<br />

([ 0,1] [ 0,1] [ 0,1]<br />

)<br />

U = ζ × ×<br />

(71.25)<br />

for an appropriate ζ<br />

a<br />

. By the preceding result we then have<br />

∫<br />

∂Ua<br />

∫<br />

Z= dZ, a = 1, …,<br />

K<br />

(71.26)<br />

Ua<br />

Thus (71.10) follows by summing (71.26) with respect <strong>to</strong> a .<br />

Following exactly the same pattern, the formula (71.10) can be proved by an arbitrary<br />

M = 4,5,6, ….<br />

Thus the theorem is proved.<br />

The formula (71.10) reduces <strong>to</strong> two important special cases in classical vec<strong>to</strong>r analysis<br />

when the underlying Euclidean manifold E is three-dimensional. First, when M = 2 <strong>and</strong> U is<br />

two-dimensional, the formula takes the form<br />

∫<br />

∂U<br />

w =<br />

∫<br />

U<br />

dw<br />

(71.27)<br />

which can be rewritten as<br />

∫<br />

∂U<br />

∫<br />

w ⋅ dλ = curl w ⋅dσ<br />

U<br />

(71.28)


Sec. 71 • Generalized S<strong>to</strong>kes’ Theorem 513<br />

where dσ is given by (70.14) <strong>and</strong> where curl w is the axial vec<strong>to</strong>r corresponding <strong>to</strong> dw , i.e.,<br />

( , ) curl ( )<br />

d w u v = w⋅ u×<br />

v (71.29)<br />

for any uv , in V . Second, when M = 3 <strong>and</strong> U is three-dimensional the formula takes the<br />

form<br />

∫∫<br />

∂U<br />

∫∫∫<br />

w ⋅ dσ = div wdυ<br />

U<br />

(71.30)<br />

where dυ is the Euclidean volume element defined by [see (68.39)]<br />

d<br />

edu du du<br />

1 2 3<br />

υ ≡ (71.31)<br />

i<br />

relative <strong>to</strong> any positive coordinate system ( u ), or simply<br />

d<br />

dx dx dx<br />

1 2 3<br />

υ = (71.32)<br />

i<br />

relative <strong>to</strong> a right-h<strong>and</strong>ed rectangular Cartesian coordinate system ( x ). In (71.30), w is the<br />

i<br />

axial vec<strong>to</strong>r field corresponding <strong>to</strong> the 2-form Z , i.e., relative <strong>to</strong> ( x )<br />

Z = we ∧e −w e ∧ e + we ∧e<br />

(71.33)<br />

1 2 1 3 2 3<br />

3 2 1<br />

which is equivalent <strong>to</strong><br />

( ) ( )<br />

Z uv , = w⋅ u× v , uv , ∈V (71.34)<br />

From (71.33), dZ is given by<br />

⎛∂w ∂w ∂w<br />

⎞<br />

d Z = + + ∧ ∧ =<br />

(71.35)<br />

( )<br />

1 2 3 1 2 3<br />

⎜<br />

div<br />

1 2 3 ⎟<br />

⎝ ∂x ∂x ∂x<br />

⎠ e e e w E<br />

As a result, we have<br />

<strong>and</strong><br />

∫<br />

∂U<br />

∫<br />

Z= w⋅dσ<br />

∂U<br />

(71.36)


514 Chap. 13 • INTEGRATION OF FIELDS<br />

∫ dZ = ∫∫∫div<br />

wdυ<br />

(71.37)<br />

U<br />

U<br />

The formulas (71.28) <strong>and</strong> (71.30) are called S<strong>to</strong>kes’ theorem <strong>and</strong> Gauss’ divergence<br />

theorem, respectively, in classical vec<strong>to</strong>r analysis.


Sec. 72 • Invariant Integrals on Continuous Groups 515<br />

Section 72. Invariant Integrals on Continuous Groups<br />

In the preceding chapter we have considered various groups which are also hypersurfaces<br />

contained in the Euclidean space L ( V ; V ). Since the integrations defined so far in this chapter<br />

can be applied <strong>to</strong> any hypersurface in a Euclidean manifold, in particular they can be applied <strong>to</strong><br />

the continuous groups. However, these integrations are generally unrelated <strong>to</strong> the group<br />

structure <strong>and</strong> thus their applications are limited. This situation is similar <strong>to</strong> that about parallelism.<br />

While the induced metric <strong>and</strong> its Levi-Civita parallelism certainly exist on the underlying<br />

hypersurface <strong>of</strong> the group, they are not significant mathematically because they do not reflect the<br />

group structure. This remark has led us <strong>to</strong> consider the Cartan parallelism which is defined by the<br />

left-invariant fields on the group.<br />

Now as far as integrationis concerned, the natural choice for a continuous group is the<br />

e is a basis for the Lie<br />

integration based on a left-invarianct volume density. Specifically, if { } Γ<br />

algebra <strong>of</strong> the group, then a volume tensor field Z is a left-invariant if <strong>and</strong> only if it can be<br />

represented by<br />

= ce<br />

∧⋅⋅⋅∧e<br />

1 M<br />

Z (72.1)<br />

where c is a constant. The integral <strong>of</strong> Z over any domain U obeys the condition<br />

∫<br />

U<br />

Z=<br />

∫L<br />

X<br />

( U)<br />

Z<br />

(72.2)<br />

for all elements X belonging <strong>to</strong> the group.<br />

A left-invariant volume tensor field Z is also right-invariant if <strong>and</strong> only if it is invariant<br />

under the inversion operation J when the dimension M <strong>of</strong> the group is even or it is mapped in<strong>to</strong><br />

−Z by J J when M is odd. This fact is more or less obvious since in general J maps any leftinvariant<br />

field in<strong>to</strong> a right-invariant field. Also, the gradient <strong>of</strong> J at the identity element<br />

Z I is invaritant<br />

coincides with the negation operation. Consequently, when M is even, ( )<br />

under J , while if M is odd, Z ( I)<br />

is mapped in<strong>to</strong> − ( I)<br />

Z by J . Naturally we call Z an invariant<br />

volume tensor field if it is both left-invariant <strong>and</strong> right-invariant. Relative <strong>to</strong> an invariant Z the<br />

integral obeys the condition (72.2) as well as the conditions<br />

∫ Z= ∫ Z,<br />

∫ Z=<br />

∫ Z<br />

( ) J( )<br />

U RX<br />

U U U<br />

(72.3)<br />

Here we have used the fact that J preserves the orientatin <strong>of</strong> the group when M is even, while<br />

J reverses the orientation <strong>of</strong> the group when M is odd.


516 Chap. 13 • INTEGRATION OF FIELDS<br />

By virtue <strong>of</strong> the representation (72.1) all left-invariant volume tensor fields differ from<br />

one another by a constant multiple only <strong>and</strong> thus they are either all right-invariant or all not<br />

SL V ,<br />

right-invariant. As we shall see, the left-invariant volume tensor fields on GL ( V ), ( )<br />

O ( V ), <strong>and</strong> all continuous subgroups <strong>of</strong> O ( V ) are right-invariant. Hence invariant integrals<br />

exist on these groups. We consider first the general linear group GL ( V ) .<br />

To prove that the left-invariant volume tensor fields on the GL ( V ) are also rightinvariant,<br />

we recall first from exterior algebra the transformation rule for a volume tensor under a<br />

linear map <strong>of</strong> the underlying vec<strong>to</strong>r space. Let W be an arbitrary vec<strong>to</strong>r space <strong>of</strong> dimension M ,<br />

<strong>and</strong> suppose that A is a linear transformation <strong>of</strong> W<br />

A : W →W (72.4)<br />

Then A maps any volume tensor E on W <strong>to</strong> ( det A)<br />

E ,<br />

A<br />

∗<br />

( ) = ( det A)<br />

E E (72.5)<br />

since by the skew symmetry <strong>of</strong> E we have<br />

for any { ,..., 1 M }<br />

( Ae ,..., Ae ) = ( det A) ( e ,..., e )<br />

E E (72.6)<br />

1 M<br />

1<br />

e e in W . From the representation <strong>of</strong> a left-invariant field on ( )<br />

GL V the value<br />

<strong>of</strong> the field at any point X is obtained from the value at the identity I by the linear map<br />

( ) → ( )<br />

M<br />

∇L<br />

: L V X<br />

; V L V ; V (72.7)<br />

which is defined by<br />

( ) , ( ; )<br />

∇ L K = XK K∈L V V (72.8)<br />

X<br />

Similiarly, a right-invariant field is obtained by the linear map<br />

( ) , ( ; )<br />

∇R<br />

X<br />

defined by<br />

∇ R K = KX K∈L V V (72.9)<br />

X<br />

Then by virtue <strong>of</strong> (72.5) a left-invariant field volume tensor field is also right-invariant if <strong>and</strong><br />

only if


Sec. 72 • Invariant Integrals on Continuous Groups 517<br />

Since the dimension <strong>of</strong> L ( V ; V )<br />

simplicity we use the product basis { ⊗ j<br />

i }<br />

be represented by<br />

det ∇ RX = det ∇L<br />

(72.10)<br />

X<br />

2<br />

is N , the matrices <strong>of</strong><br />

∇R<br />

X<br />

<strong>and</strong><br />

e e for the space L ( V ; V )<br />

∇L<br />

X<br />

are<br />

N<br />

× N . For<br />

2 2<br />

; then (72.8) <strong>and</strong> (72.9) can<br />

<strong>and</strong><br />

( ) i ik l i l i k l<br />

= = = δ<br />

XK Cjl Kk X<br />

j l<br />

Kk Xl j<br />

Kk<br />

(72.11)<br />

( ) i ik l l k i k l<br />

= = =δ<br />

KX Djl Kk K<br />

j k<br />

X<br />

j l<br />

X<br />

j<br />

Kk<br />

(72.12)<br />

2 2<br />

ik<br />

To prove (72.10), we have <strong>to</strong> show that the N × N matrices ⎡<br />

⎣C jl<br />

⎤<br />

⎦ <strong>and</strong> ik<br />

⎡<br />

⎣D jl<br />

⎤<br />

⎦ have the same<br />

ik<br />

determinant. But this fact is obvious since from (72.11) <strong>and</strong> (72.12), ⎡<br />

⎣C jl<br />

⎤<br />

⎦ is simply the<br />

ik<br />

transpose <strong>of</strong> ⎡<br />

⎣D jl<br />

⎤<br />

⎦ , i.e., ik ki<br />

D<br />

jl<br />

= C<br />

(72.13)<br />

lj<br />

Thus invariant integrals exist on GL ( V ) .<br />

The situation with the subgroup SL ( V ) or UM ( V )<br />

is somewhat more complicated,<br />

however, because the tangent planes at distinct points <strong>of</strong> the underlying group generally are not<br />

;<br />

SL V at the<br />

the same subspace <strong>of</strong> L ( V V ). We recall first that the tangent plane <strong>of</strong> ( )<br />

identity I is the hyperplane consisting <strong>of</strong> all tensors K∈L ( V ; V ) such that<br />

tr ( K ) = 0<br />

(72.14)<br />

This result means that the orthogonal complement <strong>of</strong> SL ( V ) I<br />

relative <strong>to</strong> the inner product on<br />

L ( V ; V ) is the one-dimensional subspace<br />

{ , }<br />

l = αI α∈R (72.15)<br />

Now from (72.8) <strong>and</strong> (72.9) the linear maps<br />

∇R<br />

X<br />

<strong>and</strong><br />

( α ) α L ( α )<br />

∇L<br />

X<br />

coincide on l , namely<br />

∇ R I = X=∇<br />

I (72.16)<br />

X<br />

X


518 Chap. 13 • INTEGRATION OF FIELDS<br />

By virtue <strong>of</strong> (72.16) <strong>and</strong> (72.10) we see that the restrictions <strong>of</strong> ∇R<br />

X<br />

<strong>and</strong> ∇L<br />

X<br />

on SL ( V ) I<br />

give<br />

rise <strong>to</strong> the same volume tensor at X from any volume tensor at I . As a result every leftinvariant<br />

volume tensor field on SL ( V ) or UM ( V ) is also a right-invariant, <strong>and</strong> thus invariant<br />

UM V .<br />

integrals exist on SL ( V ) <strong>and</strong> ( )<br />

Finally, we show that invariant integrals exists on O ( V ) <strong>and</strong> on all continuous subgroups<br />

<strong>of</strong> O ( V ). This result is entirely obvious because both ∇L<br />

X<br />

<strong>and</strong><br />

on L ( V ; V ) for any X∈O ( V ). Indeed, if K <strong>and</strong> H are any elements <strong>of</strong> L ( V ; V )<br />

∇R<br />

X<br />

preserve the inner product<br />

, then<br />

(<br />

T T) tr (<br />

T −1)<br />

T<br />

( KH ) K H<br />

XK ⋅ XH = tr XKH X = XKH X<br />

= tr = ⋅<br />

(72.17)<br />

<strong>and</strong> similiarly<br />

KX⋅ HX = K ⋅H (72.18)<br />

for any X∈O ( V ). As a result, the Euclidean volume tensor field E is invariant on O ( V )<br />

all continuous subgroups <strong>of</strong> O ( V ).<br />

It should be noted that O ( V ) is a bounded closed hypersurface in L ( V ; V )<br />

integral <strong>of</strong> E over the whole group is finite<br />

<strong>and</strong> on<br />

. Hence the<br />

0 < ∫ E


Sec. 72 • Invariant Integrals on Continuous Groups 519<br />

∫<br />

OV ( )<br />

f<br />

( ) ( )<br />

∫<br />

OV ( )<br />

−1<br />

( ) ( )<br />

Q E Q = f Q E Q<br />

(72.21)<br />

in addition <strong>to</strong> the st<strong>and</strong>ard properties <strong>of</strong> the integral relative <strong>to</strong> a differential form obtained in the<br />

preceding section. For the unbounded groups GL ( V ) , SL ( V ) , <strong>and</strong> UM ( V ), some<br />

continuous functions, such as functions which vanish identically outside some bounded domain,<br />

can be integrated over the whole group. If the integral <strong>of</strong> f with respect <strong>to</strong> an invariant volume<br />

tensor field exists, it also possesses properties similar <strong>to</strong> (72.20) <strong>and</strong> (72.21).<br />

Now, by using the invariant integral on O ( V ) , we can find a representation for<br />

continuous isotropic functions<br />

( ) ( )<br />

f : L V ; V ×⋅⋅⋅× L V ; V →R (72.22)<br />

which verify the condition <strong>of</strong> isotropy :<br />

f<br />

( T<br />

T<br />

1<br />

,...,<br />

P ) = f ( 1,...,<br />

P )<br />

QK Q QK Q K K (72.23)<br />

for all Q∈O ( V ), the number <strong>of</strong> variables P being arbitrary. The representation is<br />

( ) ( )<br />

( )<br />

( )<br />

∫ T<br />

T<br />

g QK1Q ,..., QK<br />

PQ E Q<br />

OV<br />

f K1,...,<br />

KP<br />

=<br />

∫ E( Q)<br />

( )<br />

OV<br />

(72.24)<br />

where g is an arbitrary continuous function<br />

( ) ( )<br />

g : L V ; V ×⋅⋅⋅× L V ; V →R (72.25)<br />

<strong>and</strong> where E is an arbitrary invariant volume tensor field on O ( V ) . By a similar argument we<br />

can also find representations for continuous functions f satisfying the condition<br />

f<br />

( ,..., ) = f ( ,..., )<br />

QK QK K K (72.26)<br />

1 P<br />

1<br />

P<br />

for all Q∈O ( V ). The representation is<br />

1<br />

( )<br />

f ( 1,...,<br />

P )<br />

∫ OV<br />

K K =<br />

∫OV<br />

g<br />

( QK ,..., QK ) E( Q)<br />

( )<br />

E<br />

( Q)<br />

P<br />

(72.27)


520 Chap. 13 • INTEGRATION OF FIELDS<br />

Similiarly, a representation for functions f satisfying the condition<br />

f<br />

( ,..., ) = f ( ,..., )<br />

KQ K Q K K (72.28)<br />

1 P<br />

1<br />

P<br />

for all Q∈O ( V ) is<br />

( K ,..., K )<br />

∫ OV<br />

f<br />

1 P<br />

=<br />

∫<br />

( )<br />

g ( KQ<br />

1<br />

,..., KPQ) E( Q)<br />

( )<br />

E( Q<br />

OV<br />

)<br />

(72.29)<br />

We leave the pro<strong>of</strong> <strong>of</strong> these representations as exercises.<br />

If the condition (72.23), (72.26), or (72.28) is required <strong>to</strong> hold for all Q belonging <strong>to</strong> a<br />

continuous subgroup G <strong>of</strong> ( ) O V , the representation (72.24), (72.27), or (72.29), respectively,<br />

remains valid except that the integrals in the representations are taken over the group G .


_______________________________________________________________________________<br />

INDEX<br />

The page numbers in this index refer <strong>to</strong> the versions <strong>of</strong> <strong>Vol</strong>umes I <strong>and</strong> II <strong>of</strong> the online versions. In addition <strong>to</strong> the<br />

information below, the search routine in Adobe Acrobat will be useful <strong>to</strong> the reader. Pages 1-294 will be found in the<br />

online editions <strong>of</strong> <strong>Vol</strong>ume 1, pages 295-520 in <strong>Vol</strong>ume 2<br />

Abelian group, 24, 33, 36, 37, 41<br />

Absolute density, 276<br />

Absolute volume, 291<br />

Addition<br />

for linear transformations, 93<br />

for rings <strong>and</strong> fields, 33<br />

for subspaces, 55<br />

for tensors, 228<br />

for vec<strong>to</strong>rs, 41<br />

Additive group, 41<br />

Adjoint linear transformation, 105-117<br />

determinant <strong>of</strong>, 138<br />

matrix <strong>of</strong>, 138<br />

Adjoint matrix, 8, 154, 156,279<br />

Affine space, 297<br />

Algebra<br />

associative,100<br />

tensor, 203-245<br />

Algebraic equations, 9, 142-145<br />

Algebraic multiplicity, 152, 159, 160, 161,<br />

164, 189, 191, 192<br />

Algebraically closed field, 152, 188<br />

Angle between vec<strong>to</strong>rs, 65<br />

Anholonomic basis, 332<br />

Anholonomic components, 332-338<br />

Associated homogeneous system <strong>of</strong><br />

equations, 147<br />

Associative algebra, 100<br />

Associative binary operation, 23, 24<br />

Associative law for scalar multiplication, 41<br />

Asymp<strong>to</strong>tic coordinate system, 460<br />

Asymp<strong>to</strong>tic direction, 460<br />

Atlas, 308<br />

Augmented matrix, 143<br />

Au<strong>to</strong>morphism<br />

<strong>of</strong> groups, 29<br />

<strong>of</strong> linear transformations, 100, 109,<br />

136, 168,<br />

Axial tensor, 227<br />

Axial vec<strong>to</strong>r densities, 268<br />

Basis, 47<br />

Anholonomic, 332<br />

change <strong>of</strong>, 52, 76, 77, 118, 210, 225,<br />

266, 269, 275, 277, 287<br />

composite product, 336, 338<br />

cyclic, 164, 193, 196<br />

dual, 204, 205, 215<br />

helical, 337<br />

natural, 316<br />

orthogonal, 70<br />

orthonormal,70-75<br />

product, 221, 263<br />

reciprocal, 76<br />

x


INDEX<br />

xi<br />

st<strong>and</strong>ard, 50, 116<br />

Beltrami field, 404<br />

Bessel’s inequality, 73<br />

Bijective function, 19<br />

Bilinear mapping, 204<br />

Binary operation, 23<br />

Binormal, 391<br />

Cancellation axiom, 35<br />

Canonical isomorphism, 213-217<br />

Canonical projection, 92, 96, 195, 197<br />

Cartan parallelism, 466, 469<br />

Cartan symbols, 472<br />

Cartesian product, 16, 43, 229<br />

Cayley-Hamil<strong>to</strong>n theorem, 152, 176, 191<br />

Characteristic polynomial, 149, 151-157<br />

Characteristic roots, 148<br />

Characteristic subspace, 148, 161, 172, 189,<br />

191<br />

Characteristic vec<strong>to</strong>r, 148<br />

Chris<strong>to</strong>ffel symbols, 339, 343-345, 416<br />

Closed binary operation, 23<br />

Closed set, 299<br />

Closure, 302<br />

Codazzi, equations <strong>of</strong>, 451<br />

C<strong>of</strong>ac<strong>to</strong>r, 8, 132-142, 267<br />

Column matrix, 3, 139<br />

Column rank, 138<br />

Commutation rules, 395<br />

Commutative binary operation, 23<br />

Commutative group, 24<br />

Commutative ring, 33<br />

Complement <strong>of</strong> sets, 14<br />

Complete orthonormal set, 70<br />

Completely skew-symmetric tensor, 248<br />

Completely symmetric tensor, 248<br />

Complex inner product space, 63<br />

Complex numbers, 3, 13, 34, 43, 50, 51<br />

Complex vec<strong>to</strong>r space, 42, 63<br />

Complex-lamellar fields, 382, 393, 400<br />

Components<br />

anholonomic, 332-338<br />

composite, 336<br />

composite physical, 337<br />

holonomic, 332<br />

<strong>of</strong> a linear transformation 118-121<br />

<strong>of</strong> a matrix, 3<br />

physical, 334<br />

<strong>of</strong> a tensor, 221-222<br />

<strong>of</strong> a vec<strong>to</strong>r, 51, 80<br />

Composition <strong>of</strong> functions, 19<br />

Conformable matrices, 4,5<br />

Conjugate subsets <strong>of</strong> LV ( ; V ), 200<br />

Continuous groups, 463<br />

Contractions, 229-234, 243-244<br />

Contravariant components <strong>of</strong> a vec<strong>to</strong>r, 80,<br />

205<br />

Contravariant tensor, 218, 227, 235, 268<br />

Coordinate chart, 254<br />

Coordinate curves, 254<br />

Coordinate functions, 306<br />

Coordinate map, 306<br />

Coordinate neighborhood, 306<br />

Coordinate representations, 341<br />

Coordinate surfaces, 308<br />

Coordinate system, 308


xii<br />

INDEX<br />

asymp<strong>to</strong>tic, 460<br />

bispherical, 321<br />

canonical, 430<br />

Cartesian, 309<br />

curvilinear, 312<br />

cylindrical, 312<br />

elliptical cylindrical, 322<br />

geodesic, 431<br />

isochoric, 495<br />

normal, 432<br />

orthogonal, 334<br />

paraboloidal, 320<br />

positive, 315<br />

prolate spheroidal, 321<br />

rectangular Cartesian, 309<br />

Riemannian, 432<br />

spherical, 320<br />

surface, 408<br />

<strong>to</strong>roidal, 323<br />

Coordinate transformation, 306<br />

Correspondence, one <strong>to</strong> one, 19, 97<br />

Cosine, 66, 69<br />

Covariant components <strong>of</strong> a vec<strong>to</strong>r, 80, 205<br />

Covariant derivative, 339, 346-347<br />

along a curve, 353, 419<br />

spatial, 339<br />

surface, 416-417<br />

<strong>to</strong>tal, 439<br />

Covariant tensor, 218, 227, 235<br />

Covec<strong>to</strong>r, 203, 269<br />

Cramer’s rule, 143<br />

Cross product, 268, 280-284<br />

Cyclic basis, 164, 193, 196<br />

Curl, 349-350<br />

Curvature, 309<br />

Gaussian, 458<br />

geodesic, 438<br />

mean, 456<br />

normal, 438, 456, 458<br />

principal, 456<br />

Decomposition, polar, 168, 173<br />

Definite linear transformations, 167<br />

Density<br />

scalar, 276<br />

tensor, 267,271<br />

Dependence, linear, 46<br />

Derivation, 331<br />

Derivative<br />

Covariant, 339, 346-347, 353, 411,<br />

439<br />

exterior, 374, 379<br />

Lie, 359, 361, 365, 412<br />

Determinant<br />

<strong>of</strong> linear transformations, 137, 271-<br />

279<br />

<strong>of</strong> matrices, 7, 130-173<br />

Developable surface, 448<br />

Diagonal elements <strong>of</strong> a matrix, 4<br />

Diagonalization <strong>of</strong> a Hermitian matrix, 158-<br />

175<br />

Differomorphism, 303<br />

Differential, 387<br />

exact, 388<br />

integrable, 388<br />

Differential forms, 373


INDEX<br />

xiii<br />

closed, 383<br />

exact, 383<br />

Dilitation, 145, 489<br />

Dilatation group, 489<br />

Dimension definition, 49<br />

<strong>of</strong> direct sum, 58<br />

<strong>of</strong> dual space, 203<br />

<strong>of</strong> Euclidean space, 297<br />

<strong>of</strong> fac<strong>to</strong>r space, 62<br />

<strong>of</strong> hypersurface, 407<br />

<strong>of</strong> orthogonal group, 467<br />

<strong>of</strong> space <strong>of</strong> skew-symmetric tensors,<br />

264<br />

<strong>of</strong> special linear group, 467<br />

<strong>of</strong> tensor spaces, 221<br />

<strong>of</strong> vec<strong>to</strong>r spaces, 46-54<br />

Direct complement, 57<br />

Direct sum<br />

<strong>of</strong> endomorphisms, 145<br />

<strong>of</strong> vec<strong>to</strong>r spaces, 57<br />

Disjoint sets, 14<br />

Distance function, 67<br />

Distribution, 368<br />

Distributive axioms, 33<br />

Distributive laws for scalar <strong>and</strong> vec<strong>to</strong>r<br />

addition, 41<br />

Divergence, 348<br />

Domain <strong>of</strong> a function, 18<br />

Dual basis, 204, 234<br />

Dual isomorphism, 213<br />

Dual linear transformation, 208<br />

Dual space, 203-212<br />

Duality opera<strong>to</strong>r, 280<br />

Dummy indices, 129<br />

Eigenspace, 148<br />

Eigenva1ues, 148<br />

<strong>of</strong> Hermitian endomorphism, 158<br />

multiplicity, 148, 152, 161<br />

Eigenvec<strong>to</strong>r, 148<br />

Element<br />

identity,24<br />

inverse, 24<br />

<strong>of</strong> a matrix, 3<br />

<strong>of</strong> a set, 13<br />

unit,24<br />

Empty set, 13<br />

Endomorphism<br />

<strong>of</strong> groups, 29<br />

<strong>of</strong> vec<strong>to</strong>r spaces, 99<br />

Equivalence relation, 16, 60<br />

Equivalence sets, 16, 50<br />

Euclidean manifold, 297<br />

Euclidean point space, 297<br />

Euler-Lagange equations, 425-427, 454<br />

Euler’s representation for a solenoidal field,<br />

400<br />

Even permutation, 126<br />

Exchange theorem, 59<br />

Exponential <strong>of</strong> an endomorphism, 169<br />

Exponential maps<br />

on a group, 478<br />

on a hypersurface, 428<br />

Exterior algebra, 247-293<br />

Exterior derivative, 374, 379, 413<br />

Exterior product, 256


xiv<br />

INDEX<br />

ε -symbol definition, 127<br />

transformation rule, 226<br />

Fac<strong>to</strong>r class, 61<br />

Fac<strong>to</strong>r space, 60-62, 195-197, 254<br />

Fac<strong>to</strong>rization <strong>of</strong> characteristic polynomial,<br />

152<br />

Field, 35<br />

Fields<br />

Beltrami, 404<br />

complex-lamellar, 382, 393, 400<br />

scalar, 304<br />

screw, 404<br />

solenoidal, 399<br />

tensor, 304<br />

Trkalian, 405<br />

vec<strong>to</strong>r, 304<br />

Finite dimensional, 47<br />

Finite sequence, 20<br />

Flow, 359<br />

Free indices, 129<br />

Function<br />

constant, 324<br />

continuous, 302<br />

coordinate, 306<br />

definitions, 18-21<br />

differentiable, 302<br />

Fundamental forms<br />

first, 434<br />

second, 434<br />

Fundamental invariants, 151, 156, 169, 274,<br />

278<br />

Gauss<br />

equations <strong>of</strong>, 433<br />

formulas <strong>of</strong>, 436, 441<br />

Gaussian curvature, 458<br />

General linear group, 101, 113<br />

Generalized Kronecker delta, 125<br />

Generalized S<strong>to</strong>kes’ theorem, 507-513<br />

Generalized transpose operation, 228, 234,<br />

244, 247<br />

Generating set <strong>of</strong> vec<strong>to</strong>rs, 52<br />

Geodesic curvature, 438<br />

Geodesics, equations <strong>of</strong>, 425, 476<br />

Geometric multiplicity, 148<br />

Gradient, 303, 340<br />

Gramian, 278<br />

Gram-Schmidt orthogonalization process, 71,<br />

74<br />

Greatest common divisor, 179, 184<br />

Group<br />

axioms <strong>of</strong>, 23-27<br />

continuous, 463<br />

general linear, 101, 464<br />

homomorphisms,29-32<br />

orthogonal, 467<br />

properties <strong>of</strong>, 26-28<br />

special linear, 457<br />

subgroup, 29<br />

unitary, 114<br />

Hermitian endomorphism, 110, 139, 138-170<br />

Homogeneous operation, 85<br />

Homogeneous system <strong>of</strong> equations, 142<br />

Homomorphism


INDEX<br />

xv<br />

<strong>of</strong> group,29<br />

<strong>of</strong> vec<strong>to</strong>r space, 85<br />

Ideal, 176<br />

improper, 176<br />

principal, 176<br />

trivial, 176<br />

Identity element, 25<br />

Identity function, 20<br />

Identity linear transformation, 98<br />

Identity matrix, 6<br />

Image, 18<br />

Improper ideal, 178<br />

Independence, linear, 46<br />

Inequalities<br />

Schwarz, 64<br />

triangle, 65<br />

Infinite dimension, 47<br />

Infinite sequence, 20<br />

Injective function, 19<br />

Inner product space, 63-69, 122, 235-245<br />

Integer, 13<br />

Integral domain, 34<br />

Intersection<br />

<strong>of</strong> sets, 14<br />

<strong>of</strong> subspaces, 56<br />

Invariant subspace, 145<br />

Invariants <strong>of</strong> a linear transformation, 151<br />

Inverse<br />

<strong>of</strong> a function, 19<br />

<strong>of</strong> a linear transformation, 98-100<br />

<strong>of</strong> a matrix, 6<br />

right <strong>and</strong> left, 104<br />

Invertible linear transformation, 99<br />

Involution, 164<br />

Irreduciable polynomials, 188<br />

Isometry, 288<br />

Isomorphic vec<strong>to</strong>r spaces, 99<br />

Isomorphism<br />

<strong>of</strong> groups, 29<br />

<strong>of</strong> vec<strong>to</strong>r spaces, 97<br />

Jacobi’s identity, 331<br />

Jordon normal form, 199<br />

Kelvin’s condition, 382-383<br />

Kernel<br />

<strong>of</strong> homomorphisms, 30<br />

<strong>of</strong> linear transformations, 86<br />

Kronecker delta, 70, 76<br />

generalized, 126<br />

Lamellar field, 383, 399<br />

Laplace expansion formula, 133<br />

Laplacian, 348<br />

Latent root, 148<br />

Latent vec<strong>to</strong>r, 148<br />

Law <strong>of</strong> Cosines, 69<br />

Least common multiple, 180<br />

Left inverse, 104<br />

Left-h<strong>and</strong>ed vec<strong>to</strong>r space, 275<br />

Left-invariant field, 469, 475<br />

Length, 64<br />

Levi-Civita parallelism, 423, 443, 448<br />

Lie algebra, 471, 474, 480<br />

Lie bracket, 331, 359


xvi<br />

INDEX<br />

Lie derivative, 331, 361, 365, 412<br />

Limit point, 332<br />

Line integral, 505<br />

Linear dependence <strong>of</strong> vec<strong>to</strong>rs, 46<br />

Linear equations, 9, 142-143<br />

Linear functions, 203-212<br />

Linear independence <strong>of</strong> vec<strong>to</strong>rs, 46, 47<br />

Linear transformations, 85-123<br />

Logarithm <strong>of</strong> an endomorphism, 170<br />

Lower triangular matrix, 6<br />

Maps<br />

Continuous, 302<br />

Differentiable, 302<br />

Matrix<br />

adjoint, 8, 154, 156, 279<br />

block form, 146<br />

column rank, 138<br />

diagonal elements, 4, 158<br />

identity, 6<br />

inverse <strong>of</strong>, 6<br />

nonsingular, 6<br />

<strong>of</strong> a linear transformation, 136<br />

product, 5<br />

row rank, 138<br />

skew-symmetric, 7<br />

square, 4<br />

trace <strong>of</strong>, 4<br />

transpose <strong>of</strong>, 7<br />

triangular, 6<br />

zero, 4<br />

Maximal Abelian subalgebra, 487<br />

Maximal Abelian subgroup, 486<br />

Maximal linear independent set, 47<br />

Member <strong>of</strong> a set, 13<br />

Metric space, 65<br />

Metric tensor, 244<br />

Meunier’s equation, 438<br />

Minimal generating set, 56<br />

Minimal polynomial, 176-181<br />

Minimal surface, 454-460<br />

Minor <strong>of</strong> a matrix, 8, 133, 266<br />

Mixed tensor, 218, 276<br />

Module over a ring, 45<br />

Monge’s representation for a vec<strong>to</strong>r field, 401<br />

Multilinear functions, 218-228<br />

Multiplicity<br />

algebraic, 152, 159, 160, 161, 164,<br />

189, 191, 192<br />

geometric, 148<br />

Negative definite, 167<br />

Negative element, 34, 42<br />

Negative semidefinite, 167<br />

Negatively oriented vec<strong>to</strong>r space, 275, 291<br />

Neighborhood, 300<br />

Nilcyclic linear transformation, 156, 193<br />

Nilpotent linear transformation, 192<br />

Nonsingular linear transformation, 99, 108<br />

Nonsingular matrix, 6, 9, 29, 82<br />

Norm function, 66<br />

Normal<br />

<strong>of</strong> a curve, 390-391<br />

<strong>of</strong> a hypersurface, 407<br />

Normal linear transformation, 116<br />

Normalized vec<strong>to</strong>r, 70<br />

Normed space, 66


INDEX<br />

xvii<br />

N-tuple, 16, 42, 55<br />

Null set, 13<br />

Null space, 87, 145, 182<br />

Nullity, 87<br />

Odd permutation, 126<br />

One-parameter group, 476<br />

One <strong>to</strong> one correspondence, 19<br />

On<strong>to</strong> function, 19<br />

On<strong>to</strong> linear transformation, 90, 97<br />

Open set, 300<br />

Order<br />

<strong>of</strong> a matrix, 3<br />

preserving function, 20<br />

<strong>of</strong> a tensor, 218<br />

Ordered N-tuple, 16,42, 55<br />

Oriented vec<strong>to</strong>r space, 275<br />

Orthogonal complement, 70, 72, 111, 115,<br />

209, 216<br />

Orthogonal linear transformation, 112, 201<br />

Orthogonal set, 70<br />

Orthogonal subspaces, 72, 115<br />

Orthogonal vec<strong>to</strong>rs, 66, 71, 74<br />

Orthogonalization process, 71, 74<br />

Orthonormal basis, 71<br />

Osculating plane, 397<br />

Parallel transport, 420<br />

Parallelism<br />

<strong>of</strong> Cartan, 466, 469<br />

generated by a flow, 361<br />

<strong>of</strong> Levi-Civita, 423<br />

<strong>of</strong> Riemann, 423<br />

Parallelogram law, 69<br />

Parity<br />

<strong>of</strong> a permutation, 126, 248<br />

<strong>of</strong> a relative tensor, 226, 275<br />

Partial ordering, 17<br />

Permutation, 126<br />

Perpendicular projection, 115, 165, 173<br />

Poincaré lemma, 384<br />

Point difference, 297<br />

Polar decomposition theorem, 168, 175<br />

Polar identity, 69, 112, 116, 280<br />

Polar tensor, 226<br />

Polynomials<br />

characteristic, 149, 151-157<br />

<strong>of</strong> an endomorphism, 153<br />

greatest common divisor, 179, 184<br />

irreducible, 188<br />

least common multiplier, 180, 185<br />

minimal, 180<br />

Position vec<strong>to</strong>r, 309<br />

Positive definite, 167<br />

Positive semidefinite, 167<br />

Positively oriented vec<strong>to</strong>r space, 275<br />

Preimage, 18<br />

Principal curvature, 456<br />

Principal direction, 456<br />

Principal ideal, 176<br />

Principal normal, 391<br />

Product<br />

basis, 221, 266<br />

<strong>of</strong> linear transformations, 95<br />

<strong>of</strong> matrices, 5<br />

scalar, 41<br />

tensor, 220, 224


xviii<br />

INDEX<br />

wedge, 256-262<br />

Projection, 93, 101-104, 114-116<br />

Proper divisor, 185<br />

Proper subgroup, 27<br />

Proper subset, 13<br />

Proper subspace, 55<br />

Proper transformation, 275<br />

Proper value, 148<br />

Proper vec<strong>to</strong>r, 148<br />

Pure contravariant representation, 237<br />

Pure covariant representation, 237<br />

Pythagorean theorem, 73<br />

Quotient class, 61<br />

Quotient theorem, 227<br />

Radius <strong>of</strong> curvature, 391<br />

Range <strong>of</strong> a function, 18<br />

Rank<br />

<strong>of</strong> a linear transformation, 88<br />

<strong>of</strong> a matrix, 138<br />

Rational number, 28, 34<br />

Real inner product space, 64<br />

Real number, 13<br />

Real valued function, 18<br />

Real vec<strong>to</strong>r space, 42<br />

Reciprocal basis, 76<br />

Reduced linear transformation, 147<br />

Regular linear transformation, 87, 113<br />

Relation, 16<br />

equivalence, 16<br />

reflective, 16<br />

Relative tensor, 226<br />

Relatively prime, 180, 185, 189<br />

Restriction, 91<br />

Riemann-Chris<strong>to</strong>ffel tensor, 443, 445, 448-<br />

449<br />

Riemannian coordinate system, 432<br />

Riemannian parallelism, 423<br />

Right inverse, 104<br />

Right-h<strong>and</strong>ed vec<strong>to</strong>r space, 275<br />

Right-invariant field, 515<br />

Ring, 33<br />

commutative, 333<br />

with unity, 33<br />

Roots <strong>of</strong> characteristic polynomial, 148<br />

Row matrix, 3<br />

Row rank, 138<br />

r-form, 248<br />

r-vec<strong>to</strong>r, 248<br />

Scalar addition, 41<br />

Scalar multiplication<br />

for a linear transformation, 93<br />

for a matrix, 4<br />

for a tensor, 220<br />

for a vec<strong>to</strong>r space, 41<br />

Scalar product<br />

for tensors, 232<br />

for vec<strong>to</strong>rs, 204<br />

Schwarz inequality, 64, 279<br />

Screw field, 404<br />

Second dual space, 213-217<br />

Sequence<br />

finite, 20<br />

infinite, 20


INDEX<br />

xix<br />

Serret-Frenet formulas, 392<br />

Sets<br />

bounded, 300<br />

closed, 300<br />

compact, 300)<br />

complement <strong>of</strong>, 14<br />

disjoint, 14<br />

empty or null, 13<br />

intersection <strong>of</strong>, 14<br />

open, 300<br />

retractable, 383<br />

simply connected, 383<br />

single<strong>to</strong>n, 13<br />

star-shaped, 383<br />

subset, 13<br />

union, 14<br />

Similar endomorphisms, 200<br />

Simple skew-symmetric tensor, 258<br />

Simple tensor, 223<br />

Single<strong>to</strong>n, 13<br />

Skew-Hermitian endomorphism, 110<br />

Skew-symmetric endomorphism, 110<br />

Skew-symmetric matrix, 7<br />

Skew-symmetric opera<strong>to</strong>r, 250<br />

Skew-symmetric tensor, 247<br />

Solenoidal field, 399-401<br />

Spanning set <strong>of</strong> vec<strong>to</strong>rs, 52<br />

Spectral decompositions, 145-201<br />

Spectral theorem<br />

for arbitrary endomorphisms, 192<br />

for Hermitian endomorphisms, 165<br />

Spectrum, 148<br />

Square matrix, 4<br />

St<strong>and</strong>ard basis, 50, 51<br />

St<strong>and</strong>ard representation <strong>of</strong> Lie algebra, 470<br />

S<strong>to</strong>kes’ representation for a vec<strong>to</strong>r field, 402<br />

S<strong>to</strong>kes theorem, 508<br />

Strict components, 263<br />

Structural constants, 474<br />

Subgroup, 27<br />

proper,27<br />

Subsequence, 20<br />

Subset, 13<br />

Subspace, 55<br />

characteristic, 161<br />

direct sum <strong>of</strong>, 54<br />

invariant, 145<br />

sum <strong>of</strong>, 56<br />

Summation convection, 129<br />

Surface<br />

area, 454, 493<br />

Chris<strong>to</strong>ffel symbols, 416<br />

coordinate systems, 408<br />

covariant derivative, 416-417<br />

exterior derivative, 413<br />

geodesics, 425<br />

Lie derivative, 412<br />

metric, 409<br />

Surjective function, 19<br />

Sylvester’s theorem, 173<br />

Symmetric endomorphism, 110<br />

Symmetric matrix, 7<br />

Symmetric opera<strong>to</strong>r, 255<br />

Symmetric relation, 16<br />

Symmetric tensor, 247<br />

σ -transpose, 247


xx<br />

INDEX<br />

Tangent plane, 406<br />

Tangent vec<strong>to</strong>r, 339<br />

Tangential projection, 409, 449-450<br />

Tensor algebra, 203-245<br />

Tensor product<br />

<strong>of</strong> tensors, 224<br />

universal fac<strong>to</strong>rization property <strong>of</strong>,<br />

229<br />

<strong>of</strong> vec<strong>to</strong>rs, 270-271<br />

<strong>Tensors</strong>, 218<br />

axial, 227<br />

contraction <strong>of</strong>, 229-234, 243-244<br />

contravariant, 218<br />

covariant, 218<br />

on inner product spaces, 235-245<br />

mixed, 218<br />

polar, 226<br />

relative, 226<br />

simple, 223<br />

skew-symmetric, 248<br />

symmetric, 248<br />

Terms <strong>of</strong> a sequence, 20<br />

Torsion, 392<br />

Total covariant derivative, 433, 439<br />

Trace<br />

<strong>of</strong> a linear transformation, 119, 274,<br />

278<br />

<strong>of</strong> a matrix, 4<br />

Transformation rules<br />

for basis vec<strong>to</strong>rs, 82-83, 210<br />

for Chris<strong>to</strong>ffel symbols, 343-345<br />

for components <strong>of</strong> linear<br />

transformations, 118-122, 136-138<br />

for components <strong>of</strong> tensors, 225, 226,<br />

239, 266, 287, 327<br />

for components <strong>of</strong> vec<strong>to</strong>rs, 83-84,<br />

210-211, 325<br />

for product basis, 225, 269<br />

for strict components, 266<br />

Translation space, 297<br />

Transpose<br />

<strong>of</strong> a linear transformation, 105<br />

<strong>of</strong> a matrix, 7<br />

Triangle inequality, 65<br />

Triangular matrices, 6, 10<br />

Trivial ideal, 176<br />

Trivial subspace, 55<br />

Trkalian field, 405<br />

Two-point tensor, 361<br />

Unimodular basis, 276<br />

Union <strong>of</strong> sets, 14<br />

Unit vec<strong>to</strong>r, 70<br />

Unitary group, 114<br />

Unitary linear transformation, 111, 200<br />

Unimodular basis, 276<br />

Universal fac<strong>to</strong>rization property, 229<br />

Upper triangular matrix, 6<br />

Value <strong>of</strong> a function, 18<br />

Vec<strong>to</strong>r line, 390<br />

Vec<strong>to</strong>r product, 268, 280<br />

Vec<strong>to</strong>r space, 41<br />

basis for, 50<br />

dimension <strong>of</strong>, 50<br />

direct sum <strong>of</strong>, 57


INDEX<br />

xxi<br />

dual, 203-217<br />

fac<strong>to</strong>r space <strong>of</strong>, 60-62<br />

with inner product, 63-84<br />

isomorphic, 99<br />

normed, 66<br />

Vec<strong>to</strong>rs, 41<br />

angle between, 66, 74<br />

component <strong>of</strong>, 51<br />

difference,42<br />

length <strong>of</strong>, 64<br />

normalized, 70<br />

sum <strong>of</strong>, 41<br />

unit, 76<br />

<strong>Vol</strong>ume, 329, 497<br />

Wedge product, 256-262<br />

Weight, 226<br />

Weingarten’s formula, 434<br />

Zero element, 24<br />

Zero linear transformation, 93<br />

Zero matrix,4<br />

Zero N-tuple, 42<br />

Zero vec<strong>to</strong>r, 42

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!