Sokolnikoff, Ivan Stephen - Tensor Analysis
Sokolnikoff, Ivan Stephen - Tensor Analysis
Analysis
Theory and Applications to Geometry and
Mechanics of Continua
I. S. Sokolnikoff
RICHARD C. FREY
1003 SUNSET
CINCINNAT! 5, OHIO
° betI a
F A i
APPLIED MATHEMATICS SERIES
Edited by
I. S. SOKOLNIKOFF
TENSOR ANALYSIS
THEORY AND APPLICATIONS TO GEOMETRY
AND MECHANICS OF CONTINUA
APPLIED MATHEMATICS SERIES
Second Edition
I. S. SOKOLNIKOFF
PROFESSOR OF MATHEMATICS
UNIVERSITY OF CALIFORNIA
LOS ANGELES
. Coordinate Systems
The Geometric Concept of a Vector
Linear Vector Spaces. Dimensionality of Space
. N-Dimensional Spaces
Linear Vector Spaces of n Dimensions
. Complex Linear Vector Spaces
. Summation Convention. Review of Determinants
. Linear Transformations and Matrices
. Linear Transformations in Euclidean 3-space
_ . Orthogonal Transformation in E,
_ Linear Transformations in n-Dimensional Euclidean Spaces
—VN. Reduction of Matrices to the Diagonal Form
133 Real Symmetric Matrices and Quadratic Forms
. Illustrations of Reduction of Quadratic Forms
. Classification and Properties of Quadratic Forms
. Simultaneous Reduction of Two Quadratic Forms to a Sum of Squares
. Unitary Transformations and Hermitean Matrices
2 TENSOR THEORY
1. Coordinate Systems
Q A P xe
———_ OO OCS
Fig. 1.
between the set of
1 Although the idea of one-to-one reciprocal correspondence
and the totality of real numbers had it roots in the Eudoxus
points composing a line
B.c., the invention of
theory of incommensurables, dating back to the fourth century
nth century. It should
coordinate systems did not come until the first part of the seventee
linear sets of points and
be also noted that a rigorous analysis of the relation between
last century, chiefly through
real numbers was made only during the closing years of the
of rigor depends entirely on conven-
the efforts of Dedekind and Cantor. The concept sophistication
ng tastes indicati ve of the degree of mathema tical
tions dictated by prevaili made rigorous
concepts are usually
in a given chronological period. Fruitful intuitive of definable
agreeme nts as to which ideas fall into a category
by (a) making explicit
into mathematical theories new modes
concepts and which do not, and (6) introducing
hopes) are free of contradi ction.
of reasoning which (one
1
M LINEAR VECTOR SPACES. MATRICES [CHaP. 1
Fig. 2.
designate one of these as the positive and the other as the negative half-ray.
On the positive half-ray we choose a point A and call the length of the
line segment OA the unit length. We next coordinate points on X with a
set of real numbers in the following way: If P is any point on the positive
half-ray, we define a number x associated with P by the formula
where OP and OA are lengths of the line segments OP and OA. The
number 2 is the coordinate of P. The coordinate x of the point Q on
the negative half-ray is defined by the ratio
We also assume that each real number x corresponds to one and only
one point on X. This association of the set of points on X with the set
of real numbers constitutes a coordinate system of the one-dimensional
space consisting of points on X.
The coordination of the set of points lying in the plane with sets of
real numbers is accomplished by taking two straight lines X, and X,
intersecting at a single point O-(Fig. 2). On each line a coordinate system
is constructed as above, but the units on each line need not be equal.
A pair of such lines with unit points 4 and B marked on them form the
coordinate axes X,, X,. With each point P in the plane of coordinate
axes we associate an ordered pair of real numbers (x,, x.) determined as
follows. The line through P drawn parallel to the X-axis intersects the
X-axis in a point M, with coordinate a,, and the line through P parallel
to the X,-axis cuts X, in a point M, with coordinate x. The ordered
pair of numbers (x,, x.) are the coordinates of P in the plane, and the
SEG is THE, GEOMETRIC CONCEPT (OF Ax VECTOR 3
one-to-one correspondence of ordered pairs of numbers with the set of
points in the plane X,X, is the coordinate system of the two-dimensional
space consisting of points in the plane.
The extension of this representation to points in a three-dimensional
space is obvious. We take three noncoplanar lines Xj, X, X3 intersecting
at the common point O. On each of these lines we establish coordinate
systems, and we associate with each point P an ordered triplet of numbers
(x1, Xa, %3) determined by the intersection with the axes of three planes
drawn through P parallel to the coordinate planes X,X>, XX, and X,X3.
The coordinate systems just described are called oblique cartesian
systems. Their construction makes use of the notions of length and
parallelism of ordinary Euclidean geometry, and the essential feature of
it is the concept of one-to-one correspondence of points with ordered
sets of numbers. In the event the coordinate axes Xj, X2, X, intersect
at right angles, the coordinate system is said to be orthogonal cartesian,
or rectangular cartesian. In applications, orthogonal coordinate systems
are generally used because the expression for the length d of the line
segment AB joining a pair of points with coordinates A(q,, a2, a3) and
B(b,, by, bs) has the simple form
where the g,,’s are constants that depend on the coefficients in the above-
mentioned linear transformation of coordinates. We will be concerned
in the sequel with a detailed study of quadratic forms appearing under
of
the radical in formula 1.2 and with their bearing on metric properties
space.
B b C
Fig. 3.
(3.2) = ma + nb.
Formula 3.2 follows at once from the rule for addition of vectors and
from the definition of multiplication of vectors by scalars. Equation 3.2
can be rewritten in symmetric form to read
aa + fb + ye = 0,
which is the condition for linear dependence of the set of three vectors,
since not all constants in this formula vanish. The formula ma + nb,
where a and b are two linearly independent vectors and m and n are
arbitrary real numbers, defines a two-dimensional real linear vector space.
We see that in a two-dimensional linear vector space a set of three vectors
is always linearly dependent.
O a b
Fig. 4.
SEC. 3] DIMENSIONALITY OF SPACE f|
Fig. 5. Fig. 6.
(3.3) d = ma + nb +
pe,
from which it follows that among four vectors a,b,c,d there always
exists a nontrivial relation of the form
aa + Pb + ye + dd = 0.
where (2,, 2, #3) are called the physical components of x, and the terminal
points of the base vectors a,, (i = 1, 2, 3), have the coordinates
sare (15050):
asin 0.4180):
as (0, 07 1%,
We conclude this section by noting the rules for the addition and
multiplication of vectors when the latter are referred to an orthonormal
system of base vectors a,, (i = 1, 2,3). If we have two vectors x and y
whose components are (a, 2, %3) and (Yj, Ys, Ys), respectively, then the
vector x + y has the components (7, + 41, 22 + Yo, 73 + Ys). If « is a
real number, the components of the vector «x are (a, «22, x3). From
the distributive law of scalar multiplication of vectors it follows at once
that the product of
X = Xa, + Xa, + Xa
and
y Yay + Yoas + Y3a3
is
X°Y = XWYy + TeYo + Xz,
since a,-a; = 0,,, where 6,,; = 1 if i=/, and 6,,;= 0 when i #/j. This
follows from the assumed orthonormal nature of the base vectors a;.
The foregoing formula leads at once to the familiar expression for the
length |x| of the vector x referred to an orthogonal cartesian reference
system. Thus
Xe
x = v7 ey?
4 aw?
= [xis
so that
[x] = Via? + 2? + a5”.
Clearly |x|-> 0, unless z°=="a5 =", = 0:
4. N-Dimensional Spaces
in which the functions x; are single-valued and are such that, in the region
under consideration, they yield N single-valued solutions
Ys = Yl%1, Te, --- > xy),
follows
can be satisfied only by choosing ¢y = ¢, =*"* = Cm = 0. It
that x, ¥ 0, for, if it were zero, the numbers
=¢, =90
¢=1@=G=-°
y dependent,
would satisfy (5.2) and hence the vectors would be linearl
produc t of x, by
which is contrary to our hypothesis. Denote by a, the
12 LINEAR VECTOR SPACES. MATRICES [CuaP.1
Xp * a, — (X2) + a,)a, + a, = 0.
is linearly independent, and we can define the vector aj’ by the formula
(5.3) 930 Ao ae
O50: 0: a
The constants a, %,...,«, in (5.4) are called the components of the
vector x. Multiplying (5.4) scalarly by aj, a,...,a n in turn, and
remembering that? a, +a; = 0;;, we obtain
a,°X = a, a, +X = %, oe a
Thus the vector x can be represented in the form
(5,5) X = (a,-x)a, + (a,-x)a. +°+* + (a, + x)a,.
If we introduce the notation a; +x = 2,, equation 5.5 assumes the form
eS 1 Se tes Sp OO Sea
y: (Y, a
by the rule
PG (2, + Yy, Fe + Yo... + 5.%y + ¥,),
We note that
(6.3) xe y = y-xX,
since the conjugate of the sum is the sum of the conjugates and the
conjugate of the product is equal to the product of conjugates.
Formula 6.1 is adopted for the calculation of the scalar product in
order to ensure that
2 OC >)LYis
i=1
then
(xp yy Ze Xe ZiT y 2,
x-(y +z) =x-y + x-Z,
(kx) -y =K(x-y),
x + (ky) = k(x-y).
2. Prove that, ifa™, a, ..., a‘ is a set of n linearly independent vectors in
l to each
a complex n-dimensional vector space, then the only vector x orthogona
of the vectors a‘) is the zero vector.
3. Prove that a set of mutually orthogonal nonzero vectors is always linearly
independent. — :
4. Let the set of vectors a® in E,: (a, af,... , ae) im 1,25. 5 A, de
linearly dependent, and suppose that r of them, a, a®),.. .,a', r <n, are
linearly independent. Show that every vector x that is orthogon al to this set of r
linearly independent vectors forming the subset of E,, is also orthogonal to the
remaining n — r vectors in the given set.
16 LINEAR VECTOR SPACES. MATRICES [CuaP.1
i=1
twice; we will omit the summation symbol & and write a,;7; to mean
Ay%, + AX. + ast, + a,x, Of course, the range of admissible values of
the index, | to 4 in this case, must be specified. If the symbol 7 has the
range of values | to 3 andj ranges from | to 4, the expression
(7A) Aj ;X 5, (i a L 2. 3),
o™
5,
I 1, 2, 3, 4),
represents three linear forms
In expression 7.1 the index i is the identifying index. It denotes one of the
forms in (7.2), depending on the chosen value of i. The index j, however,
since it occurs twice, is the summation index. The summation (or dummy)
index can be changed at will. Thus (7.1) can be written in the form a,,2,
if k has the same range of values as 7. The summation index is analogous
to a variable of integration in a definite integral, which also can be
changed at will.
Unless a statement to the contrary is made, we will assume that the
summation and the identifying indices have the ranges of values from
1 ton. Thus a,x, will represent a linear form
The quadratic form > > a,;,2, will be written a,,v,7;. An expression
t=1j=1
a,;«y; represents a bilinear form containing n? terms, whereas 4;,a,,
represents n® sums of the type
since each of the identifying, or free, indices i and k can have values
from | to n. We will not trouble to enclose the indices in parentheses
when the context makes it clear (as in the above expression) that such
indices have fixed values. If, however, we wish to discuss a particular
term in this sum we will write @,,;)@,,),.
Frequently, it is convenient to identify the different symbols by using
superscripts rather than subscripts. For instance, we may write the
sequence of terms x’, x, ... , #", where the superscripts are not the powers
of the variable but the identifying indices. The typical term in this
sequence is x’, (i= 1,2,...,m). A linear form in the x’, with the co-
efficients a;, will be written as a,x‘. A bilinear form, with the coefficients
a’), in the variables x, and y; will be written as a’’a,y;.
A determinant
4, U2 Qn
Gx, Ag9 Aen
For the multiplication of two determinants |a;*| and |b,*| we have the
familiar rule: by
la,'|-[5;"| = lel,
where c,’ = a,‘b;*. If we deal with determinants |a,,| and |b,;|, then the
and
element c,, in the ith row and the jth column of the product of |a,;|
|b;;| isc, = AinDy 3:
18 LINEAR VECTOR SPACES. MATRICES [CuaP. 1
éi=1,
I
ifi=y,
=0, ifi¥),
then for the expansion of |a,‘| in terms of cofactors we have the following
formulas:
(7.3) GA oo
(7.4) a;'A* — ao;*,
We can derive Cramer’s rule for the solution of the system of 7 linear
equations
Thus
tak
ai;A = a,
A;;)A™) = a,
Sec. 8] LINEAR TRANSFORMATIONS AND MATRICES 19
To gain familiarity with this notation, the reader is advised to derive
Cramer’s rule when the system of linear equations is written in the form
a,;x? =-b;. He will also prove that, if a;‘b,? = 6,', then |a,'| = 1/[b,'|.
We will return to the subject of determinants in Sec. 41, where a
different notation permits us to eliminate references to rows and columns
of the determinant and enables us to write it in terms of its elements,
without reference to cofactors.
Problems
1. Write out in full the following expressions:
a ,
of;
(e) meee (f)0;*. (g) a = yi b!, (h) aijayv'y™.
é ay* dy*
© 26 —3253,)° (j) aiqjye. (k) 5;;0™.
The symbols 6;', 6;;, and 6” all denote the Kronecker deltas.
2. Verify that (7.6) is the solution of (7.5).
(b) The two sets of numbers (a1, %,..., ,) and (2,’, Bike Sse ean
be regarded as components of the same vector x when x is referred to two
different sets of cartesian reference frames determined by the base vectors
41, Mg, 4.58, and a,’ a, 400.9185 in thisvevent equations 61 give a
transformation of coordinate axes.
1 1 1
44, 12 Ain ay a, a,
2
(a,;) = Ao, Ag9 Gan or (aj)= a, 4a: Ane
m™m mm ™m
ami Ame Amn ay as a,
We shall also write the symbol A for the matrix (a,;). We shall say that the
matrix A = (a;,;) is equal to the matrix B = (6,,) if, and only if, a,; = 5,;
for each i and j. That is, if 4 = B the elements in the corresponding
rows and columns of the matrices must be equal.
By the sum 4 + B of two matrices A = (a,;) and B = (b,,;) of the same
type, that is, containing the same number of rows and columns, we mean
the matrix
A+ B= (a, + b,).
If we have an m X n matrix A and ann X p matrix B, we can define
the product of matrices A and B, written AB, by the formula
(8.3) AB = (4,;b;,).
Such matrices are called diagonal matrices. The diagonal matrices will
be found to be of considerable importance in what follows.
A particular diagonal matrix
L4,0 0
ase | 0
[=
OO 1
AI =1A=A.
when
We also observe that the product of two matrices may vanish
neither of the matrices is a zero matrix.
le bad 0.00 00 0
z, 0 0 0
a n.
(8.5) x = A-ly’,
where
An An Ana
|A| |A| |A|
Aj. Ags Ans
and the 4,,’s are the cofactors of the elements a,,; in the determinant |A].
The matrix A-! is called the inverse of the matrix A, and it is defined
for any nonsingular matrix A. From definition 8.6 it follows that the
matrices A and A~! are related by the formulas ‘
AAt=1, AtA=],
the transpose of A.
and multi-
Using the definition of transpose and the laws of addition
plication of matrices it is easy to show that
y verified by multi-
have unique solutions X¥ = A™, as can be immediatel
plying them by 4~? on both sides and noting that
A1A = AAT =I.
24 LINEAR VECTOR SPACES. MATRICES [CHaP. 1
Since the product BA, in general, is not equal to AB, we see that the order
in which the transformations are performed is not immaterial.
It should be observed that the matrix A in the equation 2’ = Az can
be interpreted as an operator which converts a vector x into another
vector x’. Because of the properties
A(kx) = kAx
and
A(x + y) = Ax + Ay,
where k is any scalar, A is frequently called a linear vector operator or
linear vector function. It can be viewed as an apparatus for the manu-
facture of a new vector from a given vector. We shall expound these
points in greater detail by considering a number of examples of the uses
of matrices in several situations familiar from analytic geometry and
elementary vector analysis.
Sec. 9] TRANSFORMATIONS IN EUCLIDEAN 3-SPACE 25
(9.3) x, = Ax,
(9.4) =)
al ba?
denote
We shall suppose that the matrix (6,,) = B is nonsingular and
of x relative to the new system by (&,, &, &3), so that
the components
(9.5) x = Gal”.
a‘) in
If we insert in (9.5) the expressions 9.4 for the base vectors
terms of a‘), we obtain
(9.6) x = &,b,a”.
tion between
A comparison of this equation with (9.1) yields the connec
the components &; and 2,, namely,
a, = bi56;.
(9.7)
of base vectors
We note that the matrix B in the transformation 9.4
9.7 of components
a) differs from the matrix B’ in the transformation
26 LINEAR VECTOR SPACES. MATRICES [CuapP.|
of the vector x in that the rows and columns in these matrices are inter-
changed. Thus the matrix B’ is the transpose of the matrix B. We write
(9.7) in the form
(9.8) ena bk
The solution of (9.8) for & is given by
(9.9) EB = (B’)"x.
To simplify writing we denote (B’)! by C, so that (9.9) becomes
(9.10) E = Cx,
where
(9.11) C= (By.
Formula 9.10 permits us to calculate the components of the vector x
when it is referred to a new system of base vectors a"), determined by
(9.4). Consequently the components &,’, 4’, 5’ of x’, relative to the
reference frame with base vectors a‘), are given by
(9.12) E’ = Cx’,
and the question of the expression (in the new frame) for the deformation
of space characterized by (9.3) amounts to finding the relation connecting
the components &,, &, &; with §&,’, &’, €;’. The substitution from (9.3)
in (9.12) gives
E = CAx:
and, since by (9.10)
x= CF,
we get the desired relation
(9.13) &’ = CACHE,
The transformation determined by the matrix S = CAC~? is called similar
to the transformation produced by A because formulas 9.13 and 9.3
characterize the same deformation of space relative to two different
reference frames.
If we recall the definition (9.11), we can write (9.13) in the form
(9.14) B= (BY) 1ABE,
which brings into explicit evidence the matrices A and B characterizing,
respectively, the deformation of space and the transformation of base
vectors. We note that the determinants of all similar transformations
are equal. An important special case of the transformation 9.2, corre-
sponding to the rotation of the vector x to a new position, is discussed
in the next section.
Sec. 10] ORTHOGONAL TRANSFORMATION IN £; os|
Let us suppose that the base vectors a, a), a) in Sec. 9 are orthogonal
unit vectors, so that the measure numbers «; in (9.1) are the, physical
components of x. Then the square of the length of the vector x is given
by the formula
|x|? = a ,, (i =e 1275):
or
since
Oj4XjXy = UX, = UX;
equations
Equating the coefficients of like products in (10.2), we obtain six
Ay? + Gy;” + ag = 1,
Ayo? + doe” + Ago” = 1,
Ays? + eg” + 433” = |,
y9Ay3 + A223 1 432433 = 0,
43411 + 23421 + 433431 = 0,
AyAya + Ao1429 + 431432 = 0,
or
(10.4) |a,;4i| = 1.
unchanged when its rows
Since the value of the determinant |a,;| is
see from the rule for multiplication of
and columns are interchanged, we
determinants (Sec. 7) that
\a;;4%| = |a,;| ° |a;;| = |A|?.
28 LINEAR VECTOR SPACES. MATRICES [CHAP.1
Thus (10.4) yields the result that the square of the determinant |a,,| in
(9.2) has the value 1 whenever the length of the vector is unchanged by
the transformation. We conclude that |A| = +1. The case when
|A| = +1 corresponds to the transformation of rotation of space relative
to fixed axes. The circumstance when |A| = —1 corresponds to the
transformation of reflection (say, 7,’ = —2,, %2' = —%,%3) = —3) or a
reflection followed by a rotation.
A linear transformation
(10.5) Ly = Azz;
in which a;,,a,;, = 6,, is called an orthogonal transformation. It is called
the transformation of rotation when |a;;| = +1. If we denote by A’
the transpose of A in (10.5), we can write the orthogonality conditions
(10.3) in the form
AA =I.
Multiplying this equation on the right by 4, we get
(10.6) A’ = A,
so that in an orthogonal transformation the inverse matrix A~ is equal to
the transpose A’ of A when the base vectors are orthonormal.
It follows that, if we write equations 10.5 in the form
XestAy:
then
xX = Aix.
and by virtue of (10.6)
x = A’x’
or
(10.7) XL, = a;;X;'.
ag di, 0, 0, Vo 0),
a): (0,1,0,...,0),
Sec. 11] LINEAR TRANSFORMATIONS Dose)
and represent any vector x: (#, 2% ,...,2,) in the form (cf. equation 9.1)
(11.1) x = za, (i at ot rt)
A linear transformation of components, corresponding to equation
9.2, is
(11:3) Ss be
where A = (4;;).
We suppose that |A| ¥ 0, and denote the solution of (11.3) by
x = Aly’,
where
An = (A;,) ;
|A|
The A,,’s denote the cofactors of the elements a,; in |A|.
Just as was done in the three-dimensional case, we can show that the
product of transformations x’ = Ax and x” = Bx’ is x” = BAx. We
can still use the suggestive language of geometry and speak of the set of
equations 11.3 as representing the deformation of space E,, and consider
that the transformation of the form
(11.4) x’ = CACx
by the
represents the same deformation of space as that characterized
similar.
matrix A in (11.3). The matrices A and CA C- are still termed
rmation
By analogy with the three-dimensional case, a real linear transfo
, x,) invaria nt is
that leaves the length of every real vector x: (%,...
10 it is obvious that the
called orthogonal. From computations of Sec.
(11.2) satisfy the relation s
coefficients a,, in an orthogonal transformation
(11:5) jin = Ojn9
transformation is related
and that the matrix A = (a,,) of an orthogonal
condition (11.5) is both
to its inverse by the formula A’ = A, The
be orthogonal. Since
necessary and sufficient for a transformation to
ormation is equal to
the transpose of the matrix of an orthogonal transf
its inverse, we deduce that 4;,4,; = 9jx-
ions (11.5) is called
Any matrix satisfying the orthogonality condit
such a matrix has the
orthogonal. The square of the determinant of
value 1.
30 LINEAR VECTOR SPACES. MATRICES [CuaP.1
then C = (8).
If the vectors a are orthonormal and the matrix B orthogonal, the
new set of vectors a” will obviously be orthonormal. Whenever |5;;| = 1,
we shall speak of (11.6) as representing a rotation of base vectors in E,,.
We now raise the question: Is it possible to find a matrix C such that
the matrix CAC has the diagonal form
0 0
an Bidecd be 0 |2
0 0 os
This means that relative to a suitable reference frame the deformatioa
of space, characterized by (11.2), assumes the form
the &,”s being the components of x’ and the €,’s of x in the new coordinate
system.
In the language of transformations in £3, equations 11.7 state that for
a suitably chosen reference frame the linear deformation of space is
equivalent to simple extensions or contractions along the coordinate axes.
Clearly the possibility of such reduction depends on the nature of
coefficients a,,; in (11.2).
A detailed discussion of the problem of reduction of matrices to various
canonical forms is involved. In the following sections we treat only
those cases that occur most frequently in applications, referring the
reader for an exhaustive treatment to standard treatises on higher algebra.
XL, = a,x;
assumes the form [see (11.7)]
or
(12.2) AS = SA,
where A = (a,;) and
A, 0 0
ares ‘ay, 5 0
0 O A
where
Siy Siz Sin
obtain a system of n
If in (12.3) we set i = 1,2,...,m, and fix k, we
appearing in the kth
equations containing the elements (Sips Sons ++ > Sng)
can be viewed as com-
column of S. The elements (5), 52x; - - . 5 Snp)
ination of the matrix S is
ponents of the vector s‘), so that the determ
(k = 1,...,7”), whose com-
equivalent to finding a set of n vectors s®,
ponents satisfy equations 12.3. Accordingly we write equation 12.3 in
the form
ee 6,54 = 0,
ay,—A ay ey Ay
This nth order algebraic equation in A has n roots, which are known as
the characteristic values* of the matrix A. If these n roots are distinct,
we can readily show that the system of equations 12.4 yields a set of n
linearly independent vectors s“), and hence a nonsingular matrix S, as
required by (12.1), exists. If the roots are not distinct, it may not be
possible to determine the desired matrix S.
‘We consider the case when the roots are distinct, and denote them
by A,, 4,,...,4,. If we set A, for A, in (12.5) we obtain a system of n
homogeneous equations. This system will have a nontrivial solution
$1, So1,+++55S_,- Setting A, = A, in (12.5) we get the system yielding a
solution S19, 5g, ..., Sng. This gives the second column of S. Proceeding
in this fashion we can determine the remaining columns and hence the
matrix S, which satisfies the equation 12.2. To show that the trans-
formation 12.1 is possible, we must demonstrate that the vectors s\) so
calculated are linearly independent, so that S possesses an inverse S~1.
We shall prove this by supposing that the matrix S is singular and reaching
a contradiction.
If |S| = 0, the vectors s“) appearing in the columns of S are linearly
dependent, and hence there exists a set of constants c,, not all zero, such
that
cs) + cs@ + -+-+c,5™ = 0.
* They are also called eigenvalues and the corresponding vectors s‘*) are termed
characteristic vectors or eigenvectors.
Sec. 12] REDUCTION OF MATRICES 33
From (12.4) we deduce the relations
3
+ cys@A5-b= + +++ + ¢8MA-* = 0.
1
csMAL-}
. —1
A= tei te Ca
a re aoe 3 He?
reduction of the matrix A to the diagonal form, even when the charac-
teristic equation |A — A/| = 0 has multiple roots, can be achieved. We
turn to the consideration of these cases in the following sections.
is real and symmetric, so that a;; = a;; (or A’ = A) for all values of 7
and j. We will show that the matrix A can be reduced to the diagonal
form by the transformation S-14S. Moreover, S can be an orthogonal
matrix.
Linear transformations with real symmetric matrices occur commonly
in the study of deformations taking place in elastic media. Real symmetric
matrices also enter prominently in the study of real quadratic forms
(13.2) OCs, Sane 3) = wea (Gj = A. 25.7)
which arise in many problems in dynamics and geometry. We can assume
without loss of generality that the coefficients a,; in (13.2) are symmetric,
since (13.2) can always be written
OQ = 4,i(S.F)(552F,)
= S755
1F461
SEc. 13] REAL SYMMETRIC MATRICES 35
We denote the coefficients of &,&, by c;,,, so that
0 = cubs
RISkS
where
Thus
(13.6), Cer = Sibir
can be regarded as the element in the Ath row and the /th column of the
matrix S’B, and
(i237) C= S’AS.
We have established a
TueorEM. If the variables x, in the quadratic form Q = 4,;x,x;, with
a matrix A, are subjected to a linear transformation x, = 8;;§;, with a
matrix S, the resulting quadratic form has the matrix S'AS.
We note, as a corollary of this theorem, that the determinant of the
resulting quadratic form has the value |A| |S|?.
If the transformation 13.4 is orthogonal, then S’ = S~! and we can
write (13.7) as
C = S-1AS.
It follows from this result that the determination of an orthogonal trans-
formation which reduces the form 13.2 to the sum of the squares 13.3
reduces to the solution of the matrix equation
(13.8) SOAS = A.
This is precisely the problem we considered in Sec. 12. It follows from
the discussion of that section that the system of homogeneous equations
(13.9) AijSin = Sirdw (no sum on k),
obtained from
AS = SA
VECTOR SPACES. MATRICES [CHaP. |
36 LINEAR
solution for the vectors slay
(see equations 12.3) will have a nontrivial
4’s in (13.9) satisfy the equation
(Sit Sea ree) IG and only if, the
la,; — ;;4| = 9, or
(13.10) |A — Al| = 0.
ion 13.10, in general,
If the matrix A is arbitrary, the characteristic equat
discussed
has complex roots; and if these roots are distinct, the methods
ly independent vectors
in Sec. 12 permit us to calculate a set of n linear
er, the matrix S
s composing the matrix S. In the present case, howev
roots of the charac-
has to be orthogonal and hence real. Now, if the
from equation
teristic equation 13.10 are real, then it follows at once
be taken to be real
13.9 that the solutions s™: (Sy, Sex + - > Snx) Can
since the a,,’s are real. We prove a
of the
TuHeoreM. Jf the matrix A is real and symmetric, then the roots
characteristic equation |A — AI| = 0 are all real.
The system of equations 13.9 can be written compactly as
is real if a,, = a,;,. To prove this, note that the conjugate of 4,;5;.5;, is
equal to the original expression,
A SxSix = GSuSin = ViSaSin-
Since the left-hand member of (13.12) is real, and |s“)|? is real, it follows
that J, is real. This completes the proof of the theorem.
We prove next that, if 4; and 4; are two distinct roots of (13.10), then
the vectors s‘ and s“), corresponding to these roots, are orthogonal.
Since s‘ and s‘ satisfy (13.11), we have the identities
and the left-hand member vanishes since s‘” - As‘) = As‘ - s) because
of symmetry of 4. This establishes the orthogonality of s‘) and s"),
whenever the roots 4; and A; are unequal. Since equation 13.11 is homo-
geneous, we can multiply it by a suitable constant making the length of
s) equal to 1. We shall suppose that this has been done.
We recall that a set of orthogonal vectors is necessarily linearly inde-
pendent. Hence, if all roots of |A — AJ| = 0 are distinct, the vectors s®)
will be orthonormal, and, accordingly, the matrix S, accomplishing the
transformation S-1AS = A, will be orthogonal.
It remains to consider the case of reduction of real quadratic forms
13.2 to the diagonal form 13.3 when the equation
(13.10) |A — Al| =0
has multiple roots. The demonstration that the reduction is possible in
this case hinges on one important property of all similar matrices, namely:
the characteristic roots of all similar matrices are equal. The proof of this
is easy. We replace A in the left-hand member of (13.10) by some similar
matrix S-!AS and obtain the polynomial in 4,
ISAS — Al| = |S(A — ADS!
== |S-4|-|A — Al |S|
See
and
It follows that the characteristic equations associated with S'AS
A are identical, and hence their roots are equal.
A, be
Now let us suppose that (13.10) has multiple roots. Let A =
of (13.11) s™:
some root of (13.10), and let us determine the solution
= 1.
(eRe SD corresponding to A = Ay, which is such that s™ -s@)
We can adjoin
This can be done whether A, is a multiple root or not.
a complete
to the vector s a set of n — | orthonormal vectors forming
vectors can be
system of vectors in our n-dimensional manifold. These
of orthono rmal
used as a basis for our space instead of the original set
and we can pass from the reference frame
base vectors a™,...,a'”),
transformation.
determined by the a‘”s to the new frame by an orthogonal
referred to the new
Hence the matrix of the quadratic form 13.2, when
the form A, = S,AS,, where S, is orthogonal.
frame, will assume
38 LINEAR VECTOR SPACES. MATRICES [CHaP.1
Moreover,
(13.14) As = sh
for A = A, has the solutions s™: (1,0,0,...,0), since we chose it to
be a unit vector, and s® is one of the base vectors of the new reference
frame. If we insert this solution in (13.14), we get an identity
1 hy
0 0
A, 7 = °
0 0
from which it follows that the matrix A, has the following elements:
(315) ay AP = ae Sa = 0.
The original matrix A is symmetric, and, since orthogonal transforma-
tions do not destroy the symmetry, the matrix A, is also symmetric.”
Thus
A,’ = A,,
Ay faire
0 af) +++ ale
gem.
(08 14: 2
Ci ee ee
0 ane ann
Thus the quadratic form 13.2, when referred to our new frame, has the
structure
O71, 8 PAREEL ey = Dae n):
We succeeded in separating one square and reduced the problem to
a consideration of the form a{},é; in n — 1 variables. We can apply
"For Ay’ = (S,1AS,)’ = Sy'A(S,>)’ = S,1AS,, since S-1 = S’ for orthogonal
matrices.
REAL SYMMETRIC MATRICES 39
Sec. 13]
consider
similar reasoning to the (n — 1) x (m — 1) matrix 4, = (a\?) and
onal subspace
the form aWPE.é, Gj = 2,3,.-. _n), in the n — | dimensi
s™. In E,_4, we
E,_; of E,, determined by the base vectors other than
can calculate a unit vector s®) satisfying the equation
A,s = SA, -
by an orthog-
corresponding to A = ,, and construct a new base system
will yield a matrix
onal transformation in which s® is a base vector. This
(2) (2)
As — 0 Ass agp
(2) (2)
0 ang ann
0 he Sees 0
SHAS = A= he ; ids
0) @) Aviles Ae
For the determination of the new base vectors s‘, we have the system
of equations 13.9,
QijSin = Sihys (no sum on k),
or
(a;; — 9:;A,)Sjn = 9.
Writing these out, we have
(2 — Ay)Sixz + 482% — 953, = 0,
(14.3) 454, + (2 — Ay)Sox — 653, = 9;
65}, a OSs, aa (15 4+ Ax) Sax = 0.
—
$41 = ¢, oa =e Sy, = 0,
S.
These determine the first column of the matrix
(14.3) leads to three homogeneous
The substitution of A, = —18 in
equations
+ 4555 ae 6532 — 0,
208i
20559 Se 653 = 0,
4545 +
be
the solution of which is readily found to
res wey =
iC, Soo = 4¢, Sg. = C.
Sig =
42 LINEAR VECTOR SPACES. MATRICES [CHAP.1
Su + 53, = 0
Sec. 15] REAL QUADRATIC FORMS 43
for the determination of s™, so that the normalized solution can be taken
as
Sy = 0, tegen
(14.4) Syo + Sg = 0,
(14.5) ia “ are = 0.
Sig — S33 = 0,
ig eS1 | pS1
oe cule
the variables x; and &;
from which the equations of connection between
can be written down at once.
atic Forms
15. Classification and Properties of Real Quadr
rties of real quadratic forms
In this section we summarize several prope
O = 4,4 ;%;, (G,j=1,..-; n),
(15.1)
applications.
which are of considerable importance in
44 LINEAR VECTOR SPACES. MATRICES [CHAP.|
and we shall say that the rank of (15.1) is r. The number of positive 2’s
appearing in (15.6) is called the index of Q. If we have a form (15.6)
with p positive and r — p negative /’s, we can introduce a real trans-
formation &; = (UNDE for terms with positive ’s and €; = (1/V —A,)é;’
for terms with negative /’s so that it assumes the form
3. If the index p is equal to the rank r and r <n, then the form is
said to be positive. On the other hand, if the index is zero and the rank
r <n, the form Q is negative.
4. The forms whose canonical representation 15.3 contains both
positive and negative /’s are called indefinite.
We observe that positive and negative definite forms never vanish for
real nonzero values of the variables x; They vanish if, and only if, all
a,s vanish. In contradistinction, the positive and negative forms may
vanish for nonzero values of the arguments 2,. To see this, note that,
if r <n, then
QO = AE? + Ak? + °° + Ag,
We can make (15.1) vanish by choosing the 2; in (15.2) so that
fos, = os, = 0.
of r
The nonvanishing values of x, will surely exist, since the system
homogeneous equations,
$,;%; = 9, fp ot Se ee dP
er r < 7.
in n unknowns 2,, has nontrivial solutions whenev
that in a positive
It follows at once from (15.4), and from the fact
inant |a,,| of
definite form the A,’s in A are all positive, that the determ
is necessa rily positiv e. The convers e of this,
the positive definite form
by noting that |A| = |Al,
clearly, is not true. This can be readily seen
as well as definite forms.
and the positive value of |A| admits indefinite
PACS = Ayn.
|S| of the transformation from the initial variables x; to the final variables
yn; Since this determinant does not vanish, and since it contains no
parameter A, the roots of polynomials 16.8 and 16.9 are identical. Taking
account of the structure of expression 16.8, we conclude thatthe coeffi-
cients A, in (16.4) are the roots of the determinantal equation
(17.3) A’A = 1,
48 LINEAR VECTOR SPACES. MATRICES [CHAP.1
fact that the characteristic roots 2, must necessarily be real follows from
the observation that U-1AU is a Hermitean matrix whenever A is Her-
mitean and Uis unitary.” Thus A in (17.6) is Hermitean, and consequently
the elements along its diagonal are real. :
Problems
1. Reduce the matrix
(1 =)
A = (a;;) =
—1 1
feeet 4 0 0
C= ,Ci= , and S =
cal I role
nie
Ome
the
Discuss the meaning of A when it is viewed as an operator characterizing
deformation of space.
2. Diagonalize the matrices:
te ae Ot | =1 1 2
= | 1 -1], DO ge 1
—{-~ —1 1 Ora OS —3
Y. Since A is Hermitean,
7 For (U-!AUY = U’A(U-1)’ and (U“-AU) = U’A(U-
(U-!AU)’ =
is unitary, U’ = U-1and (U~)’ = U. Thus we have
A’ = A, and since U
WatAUs
yy
TENSOR THEORY
Again, a pair of points (P,, P.) determines a vector P,P,. This vector, in
a particular reference frame, is uniquely determined by a set of components
A;. A transformation of coordinates does nothing to the vector P, Ps, but
. . . & arta
ia
in the new reference frame P,P, is characterized by a different set of com-
ponents B,. A set of points, such as those forming a curve or surface, is
also invariant. The curve may be described in a given coordinate system
by an equation which usually changes its form when the coordinates are
changed, but the curve itself remains unaltered. We shall say, in general,
that an object, whatever its nature, is an invariant, provided that it is not
altered by a transformation of coordinates.
where a,’ is the value of 0y'/0z/ evaluated at some point P’ of the region R.
The point P’ depends, of course, on the choice of values (2', x*,..., x”).
Thus the transformation 19.1, with stated properties, is /ocally linear. The
nonvanishing of the Jacobian guarantees that this system of linear equa-
tions has a unique solution. Throughout the rest of this book we shall
suppose that all encountered transformations of coordinates are of the
form 19.1, in which the functions y‘(z) are at least of class C! in some
oy’ ’
region R, and that = ~ Oat any point of R. For brevity we shall refer
x
to a class of coordinate transformations with these properties as admissible
transformations.
As an example of an admissible transformation consider a system of
equations specifying the relation between the spherical polar coordinates
a‘ and the rectangular cartesian coordinates y’,
y' = x! sin x? cos 2%,
T: (y? =z! sin 2? sin 23,
if ==5 COS ue
If we suppose that 1 > 0,0 < 2? < z, and 0 < 2° < 2z, thenJ ¥ O and
the inverse transformation is given by
8 I
Problem
Discuss the transformations in which the coordinates y‘ are rectangular
cartesian:
2
yt = —=
/
a! + V6
— a? 4 V6
— 23,
1 1
(a) y? = —= a! — —~ 2g? 4 —_ 23
V2 V3 IS
if 1
y> = — gl i
Vy v2
(6)
SEc. 20] ADMISSIBLE TRANSFORMATIONS 53
The transformation
Tied =a ye eee ae oe tte) |
T; = 727}. If the Jacobian
is called the product of 7; and T;, and we write
of T; is denoted by J, it follows that
‘CRC as le Ozt ar
3 | ayt ax? dy’ || Ox?
ET SI a
T, and Tj, respectively.
where J, and J, are the Jacobians of
54 TENSOR THEORY [CHAP. 2
Ty*: x= xy),
te - \ y(2),
3
with T, = T,7T,, so that
Ty: « = x[y@)],
and a scalar F(P) whose component in the X-frame is f(x), we can compute
the transforms of f(z). Indeed, the component g(y) of F(P) in the Y-
frame is determined by the law
Gy»: gy) =fleM],
given by
whereas the component h(z) of F(P) in the Z-frame is
G,°: h(z) = gly@).
mation T, = 727;, we get
On the other hand, using the product transfor
GP: he =ficly@]},
from which it is clear that G,° = G,°G,°.
and the corre-
We can represent these transformations of coordinates
diagra mmatic ally as in
sponding transformation of components of F(P)
a group T of admissible
Fig. 7. Thus, as coordinates are subjected to n trans-
under go a certai
transformations, the components of a scalar T and
sive transf ormati ons
formation G®. The relation between the succes to the
ons 7,7, corres ponds
G® is such that the product of two transformati
g(y)
y
A Ts Gy, Ga’
z f(x) h(z)
x
56 TENSOR THEORY [CHAP. 2
(22.2) Jie)
ras | ae) of ’ or {fai},
Ox’ Ou" Ox”
the question arises: What does the set {f,:} become when the coordinates
a‘ are subjected to a transformation 22.1? This question is quite without
meaning unless one specifies precisely what is to be done with the set 22.2.
These fractions do not automatically “become” anything until one states
what law he is to use in calculating the “‘corresponding functions” in the
Y-frame. In other words, it is necessary to agree on what the term
“corresponding function” is to mean in a given situation.
For example, we might calculate the corresponding functions by the
transformation of invariance G° of Sec. 21; that is, we can insert in each
function f,.(z', ..., x”), the-values of the 2’s from (22.1). This will yield
a set of n functions
(22.3) gaily’, Sener CARE gly’, ae ey y"), a ee SAY. <u 2.5 y").
On the other hand, if one has in mind the notion of a gradient of f(P), it
is necessary to say that the set of functions corresponding to (22.2) is not
(22.3), but the set of n partial derivatives,
(22.4) eat
dy!’ ay?’ *”
wh
ay”?
SEC. 22] TRANSFORMATION BY COVARIANCE 57
computed by the rule for differentiation of composite functions, namely,
transformation
If we have a function f(«!,...,«#") and a
Tee aes ng 28,
Note that the law G?, for the determination of the quantities 22.8, is
different from G!. If we have a set of quantities 4,(x), A,(x),..., 4,(2),
then the law G?, determining the corresponding quantities B,(y), B.(y),...,
B,{y), is
(22.9) B, =
The law G? is the contravariant law, and we call the sets of quantities
transforming in accordance with it the components of a contravariant vector.
The laws G®, G!, and G? play a fundamental role in the development of
tensor analysis.
Problems
1. Show that if the transformation T: y* = a;‘x’ is orthogonal, then the
distinction between the covariant and contravariant laws disappears.
2. Prove the theorem: If f(x',2?,...,2”) is a homogeneous function of
of :
degree m, then Fees ee mf.
3. Given f(z, x?,...,%2") and a set of equations of transformation 2* =
xi(y}, y?,...,y"), where.each y* = y(t). If the transform of f by invariance is
gy’, y*,...,y"), show that df/dt = dg/dt. Hint: (0f]@x*)(dx*/dt) = df/dt and
dx*|dt = (dx%/ dy?)(dy?/dt).
4. Write out the laws of transformation of components of covariant and
contravariant vectors when TJ is the transformation from rectangular cartesian
to spherical polar coordinates given in Sec. 19.
(b) If T,, T:, Ts are three transformations of the type T, and Gy, G2, G3
are the corresponding induced transformations G, and if T; = T2T,, then
G,; = G,G,. In other words, the sets of transformations T and G are
isomorphic. If the given set of functions {/;} satisfies conditions (a) and
(b), we shall say that the set {f,} represents the components f; of a tensor f
in the X-coordinate system, the tensor f itself being the totality of sets of
functions {f,(x)}, {g;(y)}. ete.
It should be remarked that the term tensor was used by A. Einstein*
only in connection with the sets of quantities transforming in accordance
with the contravariant and covariant laws. The formulation of contra-
variant and covariant laws, as well as an outline of the essential features of
the algebra and calculus of contravariant and covariant tensors, is due to
G. Ricci.® The much broader characterization of tensors by the iso-
morphism of transformations of coordinates and induced transformations
is essentially due to H. Weyl and O. Veblen.® Because of the usefulness
and commonness of covariant and contravariant laws of transformation
in applications of analysis to geometry and physics, the term tensor is
generally used in the sense contemplated by Einstein. This usage is
followed in the sequel. However, the isomorphism between the laws of
transformation of coordinates and the induced transformations is so
tensor
fundamental to the idea of a tensor and to the invariant nature of
calculus that it justifies the degree of emphasis placed on it in the fore-
going.
and mixed
We now turn to a consideration of covariant, contravariant,
ns
tensors. It will be convenient to introduce (with Ricci) different notatio
at a glance.
for each type of such tensors so that they can be recognized
Let us consider first a set of functions of the variables (z!,..., 2”),
where {A(i; x)} represents the tensor in the X-coordinate system and {B(i; y)}
in the Y-coordinate system.
We denote components of contravariant tensors by superscripts. Thus
. oy '
Bie = A* — (contravariant law).
Es
’ The only exception to this convention is in the use of superscripts to identify the
variables x‘, y', etc. These quantities do not transform according to a covariant or
contravariant law unless the transformation T is affine.
SEC. 23] THE TENSOR CONCEPT 61
The definitions of contravariant and covariant tensors of rank one are
identical with those of contravariant and covariant vectors given in Sec.
22.
We speak of scalars, defined in Sec. 21, as tensors of rank zero.
We can generalize the definitions of tensors of rank one to include
tensors of any rank as follows.
DEFINITION 3. A set of n" quantities A; ;,...;,(&), associated with the
X-coordinate system, represents the components of a covariant tensor of
rank r if the corresponding set of n’ quantities B;;,...:,Y), associated
with the Y-coordinate system, is given by
att a Op
pauate a ae er eimaee
oy dy? OY uae
The tensor itself is the totality of sets of such quantities as {Aj 5, ---4,}-
DEFINITION 4. A set ofn"quantities A'\'>’**'"(x) represents the components
of a contravariant tensor of rank r in the X-coordinate system whenever the
corresponding set B's'?’’'"*(y) of n" quantities in the Y-system is given by
s A} of the
We note that this law for the transformation of component
o J
example of a mixed
mixed tensor gives Buy) = = = A®(a). Asa simple
y' Ox
62 TENSOR THEORY [CHAP. 2
tensor that already has occurred in our discussion, we cite the Kronecker
OxuOy hoy é
delta 64. Thus — —, 62 = 6/. The verification of the fact that
Oy’ ax? *dy?
the definition of covariant, contravariant, and mixed tensors satisfies
properties (a) and (b), stated in the beginning of this section, is given in
Sec. 24.
To distinguish tensors defined over a region of space from tensors whose
domain of definition is a single. point, one occasionally speaks of the
former as constituting a tensor field.
Now, if T = J, then
ea = yt, a y*2, .
and hence
Oat
oy"!
a en
1
ala
Oy" tr
so that
SEc. 24] TENSOR CHARACTER OF LAWS 63
Inserting these values of partial derivatives in (24.1) gives
Beis
On Omi: ean. ex)
ae Ae); ’
Hence G = Jif T= I.
Suppose now that, under a transformation 7,, the variables x’ transform
into y’, and the variables y’ transform into z‘ by the transformation 7).
The corresponding induced transformations G, and G, yield
i gataee ox --ox™ oy" oy’®
(
Dee
) 1
ee
r (y)
ee dy"?
aoy" Ox? —Z—
Oxks Abr:
ay "Bs
a2),
and
Ee aah Apistinamaay)
Performing the summation on «’s and f’s yields
rast is
Gag Co. ui'(2) m= i CL Ra ae RTE
dz = zt Ax = Oa’
The resulting law G; is precisely the law of transformation of the com-
ponents of a mixed tensor when the variables x’ are transformed into the
z' by the transformation T;. Thus the law of transformation G is transitive,
and this completes the proof.
The results for covariant and contravariant tensors appear as special
cases obtained by suppressing the superscripts or subscripts.
The only types of tensors with which we will deal in this book are scalars,
covariant, contravariant, mixed, and relative tensors. The last are defined
in Sec. 28.
of
We establish next a useful property of the law of the transformation
tensors, which is frequently used in the sequel.
system be
Let the components of a mixed tensor in the X-coordinate
denoted by 4j1:::4s(z) and its components in the Y-system by Bis: ///:(y).
64 TENSOR THEORY [CHAP. 2
On the other hand, if we are given the components Bix" dy), the com-
ponents Af:'''/s(a) of the same tensor in the X-reference frame are
determined by the formula
Sidyt de oc
(24.5) AGL ax) — Ox™ dy aye
“i(Y).
We note that we can obtain (24.5) from (24.4) formally by treating the
partial derivatives and sums in (24.4) as though they were fractions and
products appearing in simple algebraic expressions.
From the structure of formulas (24.4) and (24.5) we deduce an important
THEOREM. If all components of a tensor vanish in one coordinate system,
then they necessarily vanish in all other admissible coordinate systems.
This particular theorem is of profound significance in the formulation
of physical laws. It states, in effect, that, if a certain law is implied by the
vanishing of components of a tensor in one particular coordinate system,
then the rules for transformation of the tensor components guarantee that
they will vanish in all admissible coordinate systems. A physicist has
little interest in the formulation of a law that might be valid only in some
special reference frame. Indeed the notion of invariance and the universality
of physical laws is the cornerstone about which mathematical physics is
built.
ay)
= ine a Tee
Oat Dye
ays aa(®)s
Aan:
Then
Be a
Beta), (ets 28 Oy (Abe
er(otalorae Oe
Ate inaBe).
Oy — dy'")— \dxF ~— ahs
It follows from this that A + 4 is a tensor, and we write
If we equate the indices 7 and k and sum, we obtain the set of n? quantities
dy’ Ox? dx’ Ox? |,
where
(x) = =y 1°).
a a
Inserting this expression for &* in the right-hand member of the above
formula and transposing all terms on one side of the equation yields
However, 7*(y) is an arbitrary vector; hence the bracket must vanish, and
we obtain
Bia, j, k) =
This is precisely the law of transformation of the tensor of the type 4%.
Clearly, we can state an analogous theorem in which the vector § is a
covariant vector. For example, if A(i, /, k, «)é, is known to be a tensor of
the type 4j,, for an arbitrary vector é,, then A(i, j, k, «) = Ay. On the
hand, if A(i,j,k,«)é&, = A’, then A(i, j, k, ©) = A”, These
other
to
expressions suggest that an algorithm of division can be employed
determine the tensor character. Thus let A(i, j, k, a)&. = Aix, and write
symbolically.
A(i,j,k, «) = Axe ;
é
TENSOR THEORY [CHapP. 2
68
the
Now, if we should regard the covariant quantities appearing below
division line as contravariant when written above the line, we have
A(i, j,k, %) = Ai,é* :
where & is the symbolic reciprocal of &,. From the product Ai,é* we see
that A(i, j,k, «) = Ajz. Similarly, if AG j,k, #)5, = A®™*, then
tik
— AiskEa giska
A(i,j,k, «) =
-
On the other hand, if A(a,j,K)E* = Ai,
(that is, the inner product is a scalar), then the set offunctions A(iy, ... , i,)
represents a contravariant tensor of rank r in the X-coordinate system.
Proof. Since the €, are the components of a covariant vector,
Therefore
However, mbes ..., § are arbitrary; hence the term in the bracket must
vanish. Therefore
f] dy?
Bb... ., 8) = oo = (x, ep te),
Ox! Ox
which shows that
Alogi. 1G ajpmderos
This particular form of the quotient law is taken by some authors as
the definition of the contravariant tensor of rank r. Thus, if the multilinear
form A(o,...,¢,)6) +++ € is an invariant, then A(o%,,..., %,) =m
A%"’*°r, provided that the €, are the components of arbitrary vectors.
|
SEC. 28] RELATIVE TENSORS 69
On the other hand, if A(a,,..., aE Cd) tee a is an invariant, for an
arbitrary choice of &’s, then
Aly, .- +5 %) = Ag, ...0,°
r >
It is obvious from proofs of Theorems I and II that many other quotient
laws can be stated. For example, if the inner product A(i, «)&,; of the set
of n? functions A(i,j) with an arbitrary tensor is a covariant tensor of
rank two, then A(i,j)represents a mixed tensor of the type 4/. The reader
can prove this fact by following the pattern used in proving Theorem I.
The tensor properties of the set A(i,7)may be surmised from the division
algorithm. Thus, if A(i, «)&,; = ~V%;;, then AQ, a) = i Now if we
i. ag :
write the symbolic reciprocal of &,; as &”, we have A(i,a)=—=- =
6° — Az, Sad
= g(y)| =
where we have made use of Theorem II of Sec. 20. Thus the formula 28.1
determines a class of invariant functions known as relative scalars of
weight W.
A relative scalar of weight zero is the scalar defined in Sec. 21. Some-
times a scalar of weight zero is called an absolute scalar.
A relative scalar of weight 1 is called scalar density. The reason for this
terminology may be seen from the expression for the total mass of a
distribution of matter of density p(z!, 2, 23), the coordinates x’ being
rectangular cartesian. The mass contained in a volume 7 is given by the
integral M -{f{ p(x", 2, 2°) dx! dx? dx®. If the coordinates z* are
changed with the aid of the equations oftransformation x’ = x'(y', y?, y),
(i = 1, 2, 3), the mass M is given by theBee
The sets of quantities Aj! F(x) obeying this law of transformation are
called the components of a relative tensor of weight W.
From the discussion in Sec. 24, and from the transitive property of
Jacobians, namely, »
dx'| | dx ||dy"
’
az ay! dz?
(a) Relative tensors of the same type and weight may be added, and the
sum is a relative tensor of the same type and weight.
being
(b) Relative tensors may be multiplied, the weight of the product
the sum of the weights of tensors entering in the product.
yields a relative
(c) The operation of contraction on a relative tensor
tensor of the same weight as the original tensor.
sections, from
To distinguish mixed tensors, considered in the preceding
used to designate
relative tensors, the term absolute tensor is frequently
in applica tions of
the former. We shall encounter several relative tensors
tensor theory.
Problems
B’* is an arbitrary symmetric
1. Given the relation A(i,j,)B”* = C', where if
Hence deduce that,
tensor. Prove that A(i, j,k) + A(i, k,j) is a tensor.
A(i,j,k) is a tensor.
A(i,j,k) is symmetric in j and k, then
the relation A(i,j,k)B* = C?, where B’* is an arbitrary skew-
2. Given Hence, if
tensor.
symmetric tensor. Prove that A(i, j,k) — A, k,j) is a
then A(i,j,k) is a tensor .
A(i,j,k) is skew-symmetric inj and k, and a(i,/) is
arbitrary vector dx‘,
3. If ali,j)dx’ dx is an invariant for an
4;;.
symmetric, show that a(i,/)is a tensor
cofactor of a,; in |a;j| divided by
4. If a;; is a tensor, show that A¥, the
|a;;| # 0, is a tensor.
{224/0x* @x7} is a tensor with respect
5. If d(x}, ..., 2”) is a scalar, show that
coordinates.
to a set of Jinear transformations of 0
6. If |a;; — 4b,;| =O for’ = 4, in one set of variables, then |a,;’ — 4b;;,|=
In other words, the roots of the polynomial
for A = 4,, in the new set of variables.
|a,; — Ab;;| are invariants.
ip) TENSOR THEORY [CHAP. 2
recy
29.3
(29.3) = ay ay?dy* dy?,
SEC. 29] THE METRIC TENSOR %
since dx’ = (dx'/dy*) dy*.. We can thus write the formula for the square of
the element of arc in the Y-reference frame as a quadratic form
(29.5) i
Sap(¥)
ant Bat
= eae:
ay® oy? 4
These coefficients are functions of the variables (y’), and they are obviously
symmetric with respect to the indices « and f.
Since the square of the element of arc ds is an invariant, we conclude
(see Problem 3) that the set of functions g,,(y) represents a symmetric
tensor. This tensor is called the metric tensor, because, as will be shown
in Chapter 3, all essential metric properties of Euclidean space are com-
pletely determined by this tensor.
We have obtained the formula 29.4 by starting with expression 29.1,
which is characteristic of the Euclidean space. A transformation of co-
ordinates 29.2 clearly does not alter its metric properties, and formula
when
29.4 simply enables us to calculate distances in the Euclidean space
form 29.1 and
it is covered by a coordinate system Y. By starting with the
functions 29.2
the transformation 29.2, we have shown that the set of n
in which
satisfies a system of 4n(n + 1) partial differential equations 29.5,
if the functions
the g,,(y) are known functions of the variables y. Now,
+ 1) partial differential
8a, are specified arbitrarily, the system of 4n(n
in general, will have no
equations 29.5 for 7 unknown functions x‘(y),
29.5 has a solution,
solution. In the event the g,,’s are such that the system
which reduces the quad-
the existence of a transformation of coordinates
assured. In that event the
ratic form 29.4 to the sum of squares 29.1 is
. If, on the other hand,
metric tensor g,, defines an Euclidean manifold
29.5 has no solution, then no
the functions g,,(y) are such that the system
exists which reduces the ex-
admissible transformation of coordinates
to the Pythago rean form
pression 29.4 for the square of the arc element set of
is non-Eucl idean. A
29.1. We shall say then that the manifold s Sale
integrabi lity of equation
necessary and sufficient conditions for the
will be deduced in Sec. 39.
that our tensors are defined
We suppose in the remainder of this chapter atic
ds is given by the quadr
in metric manifolds and that the element of arc
gj; are funct ions belonging to the
form ds? = g,,(x) dx‘ dx’, where the is such that
r g,,(x)
class C!. We also assume that the symmetric tenso
discussion, but do not assume
lg,;| % 0 at any point of the region under
dean.
that our manifold is necessarily Eucli
74 TENSOR THEORY [Cuap. 2
Problems
Let g,,(x) represent a symmetric tensor such that the g;,,(x) belong to
class C* and g = |g;;| # 0 at any point of the region. We construct, with
the aid of the set of functions g;,(x), a new set of functions g(x), repre-
senting a contravariant tensor, which is such that g’g,; = 6). The tensors
g,;(x) and g(a) will play an essential role in all our subsequent consider-
ations, and for that reason they will be called the fundamental tensors.
Let us form a set of n? functions
Aeeeliis
(30.1) g(i,j) = < ,
and call them the Christoffel 3-index symbols of the first kind. The set of
functions
(31.5) Bis‘4ew
Formulas 31.2 and 31.5 are easy to remember if it is noted that the op-
eration of inner multiplication of [ij, «] with g** raises the index and
replaces the square brackets by the braces. The multiplication of “|by
Y
8x On the other hand, lowers the index and replaces the braces by the
square brackets. Formally, these operations of multiplication by g** and
8x are analogous to raising and lowering the indices on tensors, but we
will see that the Christoffel symbols, in general, are not tensors.
SEC. 31] CHRISTOFFEL’S SYMBOLS at
From (31.1) we readily deduce an expression for the partial derivative
of the fundamental tensor g,; in terms of the symbols of the first kind.
It is
(31.6) 0g;; ake ea) =
ago [ik,j] + Lk, i],
08;; {a | {|
(31:7) t= g,, + g.;
=v rai ye sees
if we note (31.5). An analogous formula for the partial derivatives of the
contravariant tensor g” can be obtained by differentiating the identity
£42" = 6! with respect to x". We get
O8ia ni £3 dg”!
aur ©
—— e
BH Gat
ta = 0,
or
g dg” a) OSin
* Oak Oxk™ -
where we made use of the formula 31.6. Noting the definition 31.2, we
have finally
she ee ta
Arena ile Pak)”
which is the same as
31.8
PY) ed
TR,
Ox* e
aL
ak)»
—
ied
g%”
al
O3
;
where G“ is the cofactor of the element g;;. Since the g,,’s are functions of
av, ..., a”, the G’®’s are also functions of the same variables. From (31.9)
we deduce that
dg (ZinG")
08; 08%
28 _ G49} = Gi
But
Og _ 98 O8sg _ cap 8x
O* Ogi, Ox! Ox*
and, if we recall that g*? = G**/g, the foregoing formula becomes
og 0g,
—,
Ox
= ggtt Eel.
Ox
= 29 ("|:
xi)
We close this section with some remarks about different notations used
for the Christoffel symbols by various authors. The notation [ij, k] for the
symbol of the first kind is fairly universal, but there are several different
k
notations for the symbol til Thus, many writers use the symbol {ij, kt.
SEC. 32] TRANSFORMATION OF SYMBOLS 79
ton school generally use the symbol If, for the symbol ("|adopted in this
y
book. Although the notation It has some advantages, it suggests that
the symbol of the second kind is a tensor. This, however, is not always
true, as will be seen from the developments of Sec. 32.
Problems
05; Bix ae 2
1. Show that onk ~i Oat = Ljk, i) = lij, ae
2. Show that, if g;; =0 for i #/, then (|= 0 whenever i, j, and k are
distinct. 4
3. Show that, if g;; = 0 fori #/, then
a (Bp Fk B , f
raw 4 ear ae a 4 Candee age
5. If y’ = a;‘«’ is a transformation from a set of orthogonal cartesian variables
y* to a set of oblique cartesian coordinates x* covering E3, what are the metric
coefficients g,; in ds* = g;; dx‘ dx’?
(32.2)
BS Je a oy' eh
Differentiating (32.1) we get
Tp
aye
A
(atearg AE Ny aM
“*\Oy* dy! dy? — Oy" dy? dy'
dy’ dy’ dy" Ox?
Since g,5 = Zs., we can interchange the dummy indices « and f in the
second term within parentheses and obtain
|7 = i Li, uh
where
he = oy" Oy" 60
On Ox"
If we multiply (32.3) (with k replaced by «) on the left by h*“ and on the
right by its equal from the formula written just above, and simplify, we get
k dy" Ox* Ox? gf? 4 oy 07.2% ob
lol Ox? dy’ ay?5 BalP y ay dy’ dy’ aL
Thus
(32.4) peers
Ke B a
The system of equations 32.4 can be solved for 0?x%/dy' dy’ as follows.
Multiply (32.4) by dx™/dy*, sum with respect to the common value K= 1;
and obtain
f o
vlij a
) t o n u t ( | ee B o a Oy! Oy? dy” dar
The important formulas 32.5 and 32.6 were first deduced in an entirely
of
different way by E. B. Christoffel in a memoir concerned with a study
of these
equivalence of quadratic differential forms.® We will make use
formulas to define the operations of tensorial differentiation.
ipdy?a
dyi dy’ \dx" dy"
By) = ue Ay,
oy’
and
OB, ex’ Ox CA, C7 a"
te aif ey dit a Onan
33.1 SS YH _——
so that the derivatives ofa vector do not form a tensor unless the coordinate
2 %
(33.3) A, = oa — |,
Ox? lij
— ie|e ie (?km
i Se He Agee
Ay A Site a ia oy ae
as [PlagesFe teAve. ee) [eilate ce :
al] l all
A verification of the fact that the set of functions Aj!’!:/s (~) forms a
tensor of the type indicated by the indices presents no difficulty.
If A is a tensor of rank zero, we define its covariant derivative to be the
ordinary derivative. Thus 4, = 0A/dx'. This definition is consistent
with the formula 33.5. We We note that, if the g,,’s are constants, the
Christoffel symbols vanish identically, and hence the covariant derivatives
reduce to the ordinary derivatives. This will surely be true if the gj 8 are
the metric coefficients of an Euclidean space covered by acartesian reference
system.
We remark in conclusion that the covariant x’ derivatives of relative
tensors are defined s follows. If f(x) is a relative scalar of weight W,
at W
so that g(y) =f@ |= , then
Problems
ag aAv i Ax J Aix
(a) ai aa! Bs al ie al :
(b)
ee
A; 1 Ox!
ey,
jl a + iin
:
0A,, (xil Aa; [x
(c) Ai) = aa! a = (jl A,
(d) 0Ai 5, el oe J (
7 = ae _ en <7 "hte = Tr
Liba 2
n + fi
Ava
k k
2. Prove that " = i are components of a tensor of rank three, where
ui
‘iand ‘|are the Christoffel symbols formed from the symmetric tensors
y by
a;;(x) and b;,(x). a | ay ay* ax® | ayt
3. Use the the form
formu lanG fe = ———~-,
at ax —_}
By |—
Ba and the law of transfor =
It is easy to deduce from the structure of formula 33.5 that the rules for
covariant differentiation of sums and products of tensors are identical
with those used in the ordinary differentiation. Indeed, if Ay: ‘4s(a) and
Ai 1 fe(a) are two tensors, then the formula
of the outer and inner products are given by the familiar rules,
Tee aie ae ++
andyp—10,19
we need only insert for A in formula 33.5 the product 4.7. We illustrate
the procedure by considering the product Aid, , = Up. We have
4 2 ats - f ‘ats
al al
== Atta (Sus
= |. \ li
Ox! i,U) i he|
ial
no (ies
a Ox! (1 al
Problems
to
raction of indices A%, is equivalent
1. Note that the operation of cont cont ract ion can
show that the operation of
multiplying 47; by 6/. Using this, .
r eithe r befor e or after covariant differentiation
be performed on a tenso of indic es can be perf ormed
ng or lowe ring
2. Show that the operation of raisi
t differentiation.
either before or after covarian
86 TENSOR THEORY [CHAP. 2
We will show in this section that the fundamental tensors g,; and g”
behave in covariant differentiation as though they were constants. This
follows from
Ricci’s THEOREM. The covariant derivative of either of the fundamental
tensors is zero.
Proof. Consider first the tensor g;; and form
Sigs aes
ag., mya). (o 1
we efra\ oe
Ox! lil jl
The right-hand member of this expression vanishes identically by virtue
of (31.7), so that g;;, = 0.
We can perform a similar calculation for the tensor g”, but it may prove
more instructive to differentiate the inner product g‘*g,; = 6. Thus
81805
ied
+ 8"ix 8.5, —= 3Sin
since 6;, = 0 and g,;; = 0, we have
. q baie —
SiS. = 0.
However, since |g,;| 4 0, the only solution of this system of homogeneous
equations is gi = 0.
As an immediate corollary of Ricci’s theorem we note that the funda-
mental tensors may be taken outside the sign of covariant differentiation,
and hence the operations of lowering and raising indices are permutable
with covariant differentiation. Thus
(Lai At.) = 24:1 Ajit
(36.2) A,
=»Aes " Ay,4— Mel
a, _ 7
9 oO
a) OA, a) 0A,
(36.4) axl =
Aji. = ax Ba} aa" ieee LalOak = fe xi
OL a\ 0A; a) fy
4 (NE < lsBatt 3 ae
OL
8A, a il ec) Svea 24s
+f) —olse-
CO a Ox’ Ox Qa! Ae — ff ax? (stox"
a afik
a
a) (p ow i
ah ri s A, - Ata Ay,
Asn —Avs le ae Ox* ~* ‘ife ash Ox?
88 TENSOR THEORY [CHAP. 2
and an interchange of « and f in the first terms of each preceding line gives
3 w al;
i i (P| (*\ (A [ + .
(36.6) Aj ix a Ain; =
avi ssax® ~——«CNik) lpi) Lad pad
<i: =- =< f = a
AiG)
ind 2s
ax? Gels
Ssax®* | Niki lai las) Waa
Furthermore, if the left-hand member of (36.6) is to vanish, that is, if the
order of covariant differentiation is to be immaterial, then
hea
since A, is arbitrary. In general, however, Ri, # 0, so that the order of
covariant differentiation is not immaterial. It is clear from (36.6) that a
necessary and sufficient condition for the validity of inversion of the order of
covariant differentiation is that the tensor R¥, vanishes identically.
22) 1 1 pam
The tensor \
scalvaueealtateet
is called the mixed Riemann-Christoffel tensor or the Riemann-Christoffel
tensor of the second kind.
The associated tensor
0 0 a a
(36.9) Rim =| O2* at |+ a 4
TG i) GLa] |tik e) file]
which will be found useful in listing properties of this tensor in Sec. 37.
SEC. 37] RIEMANN-CHRISTOFFEL TENSORS 89
2. Show that
i
ee) dad
Oz Nomex 6 ide ex
+ g(Lik, BILil, x] — (jl, Blk, «]),
from which it is obvious that
by
We define the Ricci tensor R;; by the formula R;; = Rij,» which,
virtue of (36.7), can be written as
Ree
a al
On Or i
fe) fal
(Bi) (Bal
fy fi]
ij) \ial|
|, (0Laced
| Ais)
a -
In Sec. 31 we have shown that ani log J/g = fa so that
due to Bianchi.
tensor g;; vanishes,
Since the covariant derivative of the fundamental
the Bianchi identity can be written in the form
(38.2) Rijxtm at Rijun,k ct Rinks = 0.
make use of the skew-symmetric
If we multiply equation 38.2 by g''g’* and
We get
properties of the Riemann tensor Rijn
o Karen, = £ hm,bin g Rim = O-
10 See Problem 2.
92 TENSOR THEORY [Cuap. 2
(38.3) CA m m
I =
Problems
1. Show that R% 7, = 0.
2. If Rj; = aie p = R/n, where R = g”R;;. (The equation R;; = pgzis
is known as the Einstein gravitational equation at points where matter is present.
It corresponds to the Poisson equation V?V =p in the Newtonian theory of
gravitation.)
3. If = 2, show that Ry/911 = Reo/ $22 = Rio/ S12 = — Riailg-
4. If n = 3, the tensor R;;,, has six distinct components, and there are six
equations Rj; = g"R,;,,;. Prove that the solutions of these equations for Rj5x;
are given by
where R = g¥R;;.
5. Verify Bianchi’s identity 38.2.
with y'(x) of class C* in R,,, in which the tensor g;;(%) has constant components
h,; throughout R,,?
Sec. 39] RIEMANNIAN AND EUCLIDEAN SPACES 93
This is one of the basic problems of differential geometry, which occurs
also under a different guise in dynamics, elasticity, relativity, and other
branches of applied mathematics. .
We note first that the components of g,,(x), when referred -to the Y-
frame, are given by
at Ab
(39.2) yeoy’ es
oy’
k Reg nd
If h,;'s are constants, then the Christoffel symbols ‘ivanish identically.
y
k 1 eee Oh;
Conversely, if the iP vanish identically, h;;; = one > and, since h,;;; = 0
y\L]
by Ricci’s theorem ,we have 0h;;/0y' = 0 in R,,. Consequently, the h,; are
constants throughout R,,. This permits us to state a
TueoreM I. A necessary and sufficient condition that the metric coeffi-
cients g,,(~) reduce to constants h,, in some reference frame Y is that the
k
Christoffel symbols Eevanish identically.
WY
From this theorem we can deduce at once a system of differential
equations that must be satisfied by functions y'(z", ... x”), if there is to be
a coordinate system Y in which the h,,;’s are constants. The law of trans-
formation 32.6 demands that
_ aout am (a
vlaB) @at dx? Oxi dx’ — aij) Ox" :
m ;
and, since |4 = 0, we have the system of equations
y\a
ou _ {1 "=o,
2,,m m
it
Ox*
5
(39.4) '
gus face (vies nD’ seein)
Ox’ ij
we now turn to the
This system, in general, will be incompatible, and
ent condit ions for the existence of
determination of the necessary and suffici
solution of the system 39.4.
94 TENSOR THEORY [CHAP. 2
ey ee ayPLES Me OAT re
Axlirw Ofte ta) HORS DFRd GaN Test waevein.
We see that if the system 39.5 has a solution, then either (39.7) are identities
in f* and 2* or else there are certain functional relations existing between
the f’s and a’s. If (39.7) are identities, the system of equations 39.5 is said
to be completely integrable. Itis then possible to prove that the integrability
conditions (39.7) are not only necessary but also sufficient to guarantee the
existence of solutions of the system 39.5.
SEC. 39] RIEMANNIAN AND EUCLIDEAN SPACES 95
There are several proofs of the existence of solution of complete systems
of partial differential equations; perhaps the simplest of these was given
by T. Y. Thomas in 1934 in a paper entitled “Systems of Total Differential
Equations Defined over Simply Connected Domains,” Annals of Mathe-
matics, 35, 730-734 (1934). An earlier proof, assuming the analyticity of
functions F;*, was given by Bouquet! in 1872, and there are other proofs
by G. Darboux and E. Cartan. We shall not go into a discussion of the
sufficiency of conditions 39.7, but will merely state an
EXISTENCE THEOREM. Let R be an open n-dimensional simply connected
region referred to the X-system of coordinates, and R’ the region composed
of R and the ranges —« < f' < o. If the functions F,'(x,f) are of class
C1 in R’ and have bounded derivatives OF,,'/0f’ in R’, and iffurthermore the
integrabiltiy conditions 39.7 are satisfied identically, then the system 39.5
has one and only one set of solutions
fin pe igiet ioe (a=
I... 6a5.07));
which for an arbitrary set of values (xp,..., % ") take on the arbitrarily
prescribed values C* = f*(xq', . . . , Xo").
We will now apply these results to the special case of the system 39.4 by
identifying it with (39.5).
The dependent variables in (39.4) are y, 4,...,U,, Whereas in (39.5)
they are
fs. [7.2 2.5f », Laus, we set
fi=y, fP = wy, ...,f7 = Ug
and
of é | y |! (tee 28s lee ite 1),
at, F; uy, F
Ox" a—l i (iso atly pret ya):
The substitution of the expressions for F;* in the integrability conditions
39.7 gives
fa ay 12h
(39.8) ij) ji
Rj, iu, = 0.
The first of these sets of equations is satisfied identically because of the
symmetry of Christoffel symbols. The second set states that the set of
Le¢gons sur
1 J, C. Bouquet, Bull. Sci. Math. et Astron., 3, (1872) p. 265, G. Darboux,
des espaces de
les systémes othogonaux, (1910) pp. 326-335, E. Cartan, Géometrié
Riemann, (1928) pp. 54-57. The proof by T. Y. Thomas is quite close in spirit to that
given by Cartan.
96 TENSOR THEORY [CHaAP. 2
Problems
1. Verify the substitutions in the integrability conditions 39.7 leading to
equations 39.8.
2. Referring to the system 39.5, show that it is completely equivalent to the
system of total differential equations
df* = F% dx’,
3. What are the integrability conditions for the equation
P(x, y, z) dx + Q(x, y, z) dy + R(x, y, 2) dz =0?
Consider also the system
OF OF )
te e oe ~
4. Prove a theorem: If Pdx + OQ dy + Rdz =Ois integrable, then
AP dx + 4Qdy + AR dz =0
is also integrable for any A(x, y, z) of class C}.
5S. Deduce the integrability conditions for the equation
Pia, ..., 2") de® = 0, (ES Ke
SEc. 40] THE e-SYSTEMS 97
1 = 63 = 63 = 63 =-"°,
-l= op Os eet
We prove in Sec. 41 that the generalized Kronecker deltas are tensors.
From definition 3, it follows that the direct product e'"2"""'e; 5...
of the two systems e’"’' and e; ...;, is the generalized Kronecker
delta. For example, e*’’e,;, has the following values:
(a) Zero, if two or more subscripts or superscripts are alike.
(b) +1, if the difference in the number of transpositions of “Py and ijk
from 123 is an even number.
(c) —1, if the difference in the number of transpositions of “By and ijk
from 123 is an odd number.
A little reflection will show that another way of phrasing statements (6)
and (c) is the following:
6a = 464+ 02 + 053)-
es
and
n!
n(n — 1)(n — De 2 (1 = tots 1) = SET :
(40.3) Tae a =
CS 3: Ape o car = 0.
Problems
5. If A;;, is completely symmetric and the indices run from 1 to , show that
the number of distinct terms in the set {A4;;,.} is
n(n — 1)(n — 2)
N=n+n(n —1) + 7
Sec. 41] APPLICATION OF THE e-SYSTEMS 101
Hint: Consider the cases where the subscripts ijk are all alike, when only two
are distinct, and when all are distinct,
6. Show that the number of distinct, nonvanishing A;;,’s in Problem 5 is
n(n — 1)(n — 2) “
when 4,;,;, is completely skew-symmetric.
3!
We recall that the determinant la’ of nth order, with elements a’, con-
sists of the sum of products of the elements where each term in the sum
contains one and only one element from each row and each column of
the determinant. The sign of each term in the sum is determined by the
character of permutation of the indices. Thus, if the superscripts in the
product aja; --- aj, are arranged in the normal order 12---7n, then
the product will carry the plus sign if the number of transpositions neces-
sary to arrange the subscripts in the normal order is even. The sign is
minus if the required number of transpositions is odd. Since ef" "** =
Side vin and OL 3°21", = i,i,---4,» the determinant
aah at
(41.1) jaja | 2 "lea
aj ay -"* ay
can be written compactly as
Pioieog,. & 2
om oft giay + * aj,
pe i
(41.2)
SS Cs ig +s sigh C2 an
As an example consider
a, a2 43
a,
ye a2 ene43|=
| 4.
jaj) =|
3 3
ay ag ag
a= x + aiaiat,
If this determinant is expanded by columns we get the
sign is assigned to
where ijk is a permutation of 123. The plus or minus
is even or odd.
term aiajat according to whether this permutation
determinant can be written |ai| = €,;,a@, a}as. On the other
Hence this
|a| = ei ava;
hand, if it is expanded by rows, we can write
Consider next the sum
Nie . 3).
C jp VAs (i,ve k, a, B, Ve
102 TENSOR THEORY [CHAP. 2
Ifkand iare interchanged in e,,;;, this e-symbol will change sign, and hence
This shows that an interchange of « and y changes the sign, so that the
system under consideration is skew-symmetric in « and y. Similar results
obviously hold for other indices. A special case of this system is the
determinant |a'| = e;,;,a,a3a5, and it follows from the foregoing that
Siete Ls
C 5540434, = |45| Cx,
We use formula 41.4 to establish the formula for the product of two
determinants. The power and compactness of this notation are strikingly
demonstrated in this derivation.
Since [b'| = e,;...,b{b3 - - b*, we can write
ee tr ng | ee aa eg eae
we see that the Kronecker delta Opie is obtained by multiplying
together two e-symbols, one of which is a relative tensor of weight +1
and the other of weight —1, and contracting with respect to a number of
indices. The result is a tensor of weight zero, that is, an ordinary tensor.
Thus we have proved that the generalized Kronecker deltas are absolute
tensors. Ji wea De é
Since 6/1/''40,
1 Ch
reduces to ee
‘8
= (0 when the coordinate system X
is cartesian, we conclude that the covariant derivatives of generalized
Kronecker deltas vanish identically. Thus the Kronecker deltas behave as
constants in a covariant differentiation.
Problems
1. Verify that 5¥/,a%? = a) — a’.
2. Verify that O7F ae = qgiik — giki 4 qgiki — gitk 4 gkii — ghii,
3. If a,; satisfies the equation
ba;; 5 ca; = 0,
ba;; + ca;; = 0.
where ¢ is a real parameter varying continuously in the interval t; < t < fy.
The one-dimensional manifold C is called an arc of a curve. In this book
we deal only with those curves for which 2‘(t) and #‘(t) = dx'/dt are
continuous functions in t; < t < fy. The definition of the arc of a curve
given here is a direct generalization of the parametric representation of
curves of elementary analytic geometry.
Let F(2t,..., 2", %,..., 2%"), viewed.as a function of t bea pre-
scribed continuous function in the interval 4; <<t<t,. We suppose
Sec. 43] LENGTH OF ARC 107
that! F(a, @) > 0, unless every #' = 0, and that for every positive number k
A Be “A : :
le eo ahem hae) SOR ee sag Re ey pa ay he)
«
The integral
“bs
(43.2) S -| F(x, x) dt
t
3 ag da oite B Beaters
(43.3 ) s = {, “Sap(X) ————
dt dt dt,
t (a n),
: wee RSH
and if this is to equal
S92
s -| F[&(s), &(s)] 4s,
81
= kF(x, #). Conversely, if this relation
we must have the relation F(é, §&’) = F(#, kaw)
C and each k > 0, then the equality of integrals is
is true for every line element of $(51)
= d(s), 6'(s) > 0, 51 <s <S, with 4 =
assured for every choice of parameter t
and ty = (52).
108 GEOMETRY [CHAP. 3
We recall from Sec. 39 that, if there exists an admissible transformation
of coordinates T: y* = y'(x!,..., 2”), such that the square of the element
of are ds,
(43.4) ds" = Pp pdrede
can be reduced to the form
(43.5) ds* = dy'‘dy’,
then the Riemannian manifold R,, is said to reduce to an n-dimensional
Euclidean manifold E,,. The reference frame Y in which the element of
arc of C in E,, is given by (43.5) is called an orthogonal cartesian reference
frame. Obviously, £,, is a generalization of the so-called Euclidean plane
determined by the totality of pairs of real values (y', y). If these values
(y', y*) are associated with the points of the plane referred to a pair of
orthogonal cartesian axes, then the square of the element of arc ds assumes
the familiar form ds? = (dy')? + (dy?)?.
In what follows we find it convenient to represent pairs of real values
(y', y*) as points in a cartesian plane even when the metric of the y’-
manifold is not Euclidean. To illustrate what is meant, consider a sphere
S of radius a, immersed in a three-dimensional Euclidean manifold Es,
with center at the origin (0, 0, 0) of the set of orthogonal cartesian
axes O-X1X?X%, Let T be a plane tangent to S at (0,0, —a), and let
the points of this plane be referred to a set of orthogonal cartesian
axes O’-¥1Y* as shown in Fig. 8. If we draw from O(0, 0, 0) a radial
x3
Fig. 8
SEC. 43] LENGTH OF ARC 109
line OP, intersecting the sphere S at P(a!, x, v*) and the plane T at
Oy’, y®, —a), then the points P on the lower half of the sphere S are
in one-to-one correspondence with points (y', y*) of the tangent plane T.
To obtain an explicit analytic form for this correspondence, we note
that, if P(x’, x, x) is any point on the radial line OP, then the symmetric
equations of this line furnish us with the ratios
a—O 2-0 8) =
he wipe) basa
Od we
or
Or
2U(y)?
+(2)? +a] =a.
Solving for A and substituting in (43.6), we get
1 2
oe? = ele see ae ee .
+ oF +a
5
VPP
A
VU'y + YP + a?
(43.7) es eee
a =a
Vy? + (y+ a?
These are the desired equations giving the analytical one-to-one corre-
spondence of the points Q on T and points P on the portion of S under
consideration.
Let P,(a}, x?, 2°) and P,(x! + dxt, a? + dx®, a + dx®) be two nearby
points on some curve C lying on S. The Euclidean distance P,P2, along
C, is given by the formula
(43.8) ds? = dx‘dx', (f= 19233),
dx = oe dy’, («=1,2).
Thus (43.8) yields a formula
ws dx" Oa! diy" dy?
S
dy" ay?
= Sap(Y) dy* dy’, (x, B — Lp 2);
110 GEOMETRY [CHAP. 3
where the g,,(y) are functions of y’ computed from (43.7) with the aid
Ox’ Ox'
of the definition g,; =
dy* dy?”
If the image K of C on 7 is given by the equations
(y= yO),
ly? = y(t); igs ES ts,
(oe
S =| Vo V7y dt.
ty
ie = (yy? + v7]
ty
Thus, for large values of a, metric properties of the sphere S are indis-
tinguishable from those of the Euclidean plane. The sum of the angles
of a curvilinear triangle drawn on S will be nearly equal to 180°, since
the sum of the angles of the corresponding triangle on T is 180° by
Euclidean geometry. Because of the limitations of measuring devices it
SEc. 43] LENGTH OF ARC 1a
may be impossible to decide a priori whether Euclidean formula 43.10
or the more involved Riemannian formula 43.9 should be adopted as a
basis for physical measurements.
The chief point of this illustration is to indicate that the geometry of a
sphere, imbedded in a Euclidean 3-space with the element of arc in the
form 43.8, is indistinguishable from the Riemannian geometry of a two-
dimensional manifold R, with metric 43.9. The latter manifold, although
referred to a cartesian frame Y, is not
Euclidean since (43.9) cannot be reduced Q
by an admissible transformation to (43.10).
Similarly, the geometry of Lobachevski
can be visualized on a surface of a “pseu-
dosphere,” a surface of constant negative
curvature generated by revolving a tractrix,
¥ = asint, C
Fig. 10 Fig. 11
* For details on hyperbolic geometry we refer the reader to specialized treatises on the
subject, especially to F. Klein’s Nicht-Euklidische Geometrie, 1, pp- 161-232.
Sec. 44] CURVILINEAR COORDINATES IN E, 113
orthogonal cartesian axes Y (Fig. 12). Consider a general functional
transformation
Te ate. yf), (i= 1, 2,3), “
Ox?
such that the 2’ are of class C1, and J = |: ~ 0 in some region R of
E,. The inverse transformation, oy
. . j
Fig. 12
114 GEOMETRY [CHAP. 3
along this line the variable «* is the only one that is changing. As an
example, consider a coordinate system defined by the transformation
Davyyt yy +O,
a? = tan
avyy +?
2
y
a = tan? g .
y
-if 21 > 0,0 < 2 < 7,0 < 2 < 27. This is the familiar spherical coor-
dinate system.
As another illustration, the transformation
y= x,
defines a cylindrical coordinate system (Fig. 14).
Fig. 13
Sec. 44] CURVILINEAR COORDINATES IN £, 115
Fig. 14
Let P(y}, y?, y?) and O(y! + dy’, y? + dy*, y® + dy®) be two neighboring
points in R. The Euclidean distance between a pair of such points is
determined by the quadratic form
(ds)? = (dy*)? +(dy?)? + (dy*)?
= dy' dy’,
; ay
and, since dy* = ee dx*, we have
x
(44.3) te = Bae Lt
where
Oy" Oy"
Peace oh = j, 2, 3).
ae Ox' Ox’ ie )
Obviously, g,; is symmetric. Moreover, it is a tensor, since (ds)? is an
invariant and the vector dz‘ is arbitrary. Denote by g the determinant
|g,;; this is positive in R since g,; dx‘ dx’ is a positive definite form.
Hence we can introduce the conjugate symmetric tensor g"’, defined in
Sec. 30 by the formula g’ = G"’/g, where G’” is the cofactor of the element
8, in g.
Consider now a contravariant vector A‘(x), and form the invariant
(44.4) A = (g,,A1A))*.
Since in the orthogonal cartesian frame the invariant 44.4 assumes the
of the
form [(A1)? + (42)? + (A%)?]’4, we see that A represents the length
by
vector A‘. Similarly, the length of the covariant vector 4, is defined
the formula
(44.5) As AAs)
In orthogonal cartesian coordinates gi = 6", and we get A = (A,A,)%.
116 GEOMETRY [CHAP. 3
OR? = PQ?
+ PR? — 2PO PRcos6,
Fig. 15
Sec. 44] CURVILINEAR COORDINATES IN £3; 117
Fig. 16
It follows from (44.6) and (44.7) that the invariant gijA*w? is equal to
cos 6, and we can write
Fig. 17
For, if A',): (de"/dsq, 0, 0) and ui): (0, dx?/ds(2), 0) are two unit vectors
directed along the X1- and X*-coordinate lines, respectively, then
= ie S12 dz" dt « §12
COS O49 = B54 M2) =
ds(1) dS(2) aj 211832
Since 211, 222,833 never vanish (see equation 44.9), we deduce from
(44.10) a
THEOREM. A necessary and sufficient condition that a given curvilinear
coordinate system X be orthogonal is that g;;=0, for iA j, at every
point of the region R.
From the definition of the element of volume dV in curvilinear
coordinates,
rs) B
dV=+
E
oy’
x?
dx) dx dx’,
where + |
5 is the absolute value of the Jacobian J of the transformation
x
y? x?
Fig. 18
120 GEOMETRY [Cuap. 3
Let a set of equations of transformation
and
Ore Ora,
ds 2 = dr- dr = an
r-dr —+-—
Da! dr‘ dz’
(45.7) dr = a, dx’
and
Sig = A; ° 4;-
We observe that the base vectors a; are no longer independent of the
coordinates (a1, 2?,:2°).
The use of covariant notation for the base vectors a; and b; can be
justified by observing from (45.2) and (45.7) that
a; dz? = b, dy‘
= bode,
Ox?
We see that the base vectors a; transform according to the law for the
transformation of components of covariant vectors,
Oy’
ba Hantnt
since the dx’’s are arbitrary.
The components of base vectors a;, when referred to the X-coordinate
Sec. 45] RECIPROCAL BASE SYSTEMS 121
system, are
a: (a,, 0, 0), a,: (0, a, 0), az: (0, 0, as),
and we note that they are not necessarily unit vectors, since, in general
(see equation 45.5),
8u = a+ a, ~ 1 S22 = a+ a, # I, 833 = Ag* as # I.
If the curvilinear coordinate system X is orthogonal, then
21; = a,° a; = |a,| |a,| cos 0,; = 0, if i ¥j.
This is the result stated in the theorem of Sec. 44.
We note that any vector A can be written in the form A = k dr, where
k is a suitable scalar. Since dr = (dr/0x*) dx’, we have
xe
Fig. 19
This formula states that the differential change in A arises from two
sources:
(a) Change in the components A’ as the values (x1, 2, 2°) are changed.
(b) Change in the base vectors a; as the position of the point (2, 2°, x°)
is altered.
(46.4) sah pp
Ox? ij)
O8 ixstseset
da; da
ome s 43
Oxt Ox' Met a;
Sec. 46] MEANING OF COVARIANT DERIVATIVES 125
If we assume that T is of class C?, then
da, ae da;
Ox? = Or" .
warn i+)
We form
pd c z
and obtain
Ca;
(46.5) =o “a, = [ij, ki).
It follows from (46.5) that
0a; >
ope om [ij, k]a*.
Hence
da; a 2 * k a
ae = {ij,kla°-a
— [ij, k]g"*
[+f Ox
0A
au! =
fe
A* Ay:
(46.6)
ae a,=—a'- aS
Ae oa
= —a'-a,
at jk
by (46.4). Since a‘- a, = 64, the foregoing result is epuivalent to
dat ,= ("|
Ox” 4 jk)”
Hence
(46.9) vs | at
pe jk
The differentiation of (46.7) with respect to 2* and the substitution from
(46.9) lead at once to (46.8).
We observe that, if the Christoffel symbols vanish identically in R,
the reference frame associated with these symbols is cartesian (see Theorem
I, Sec. 39), and, in this case, the base vectors a; are independent of the
0A OA?
coordinates x'. The formula 46.3 then states that —— = — a,, and
OAt Ox? Ox?
hence A’. =
Ox) ~
dA _OA dx!
dai ae dix,
SEC. 47] INTRINSIC DIFFERENTIATION 127.
By virtue of (46.6) this can be written
= [2+ bee
dite
The vector 6A*/6t, defined by the formula
a
hijlowdt
and
dA aie (« = 1, 2, 3),
(47.1)
vel z 5 x
6t dt ij) dt
is called the absolute or intrinsic derivative of A* with respect to the
parameter f.
Following McConnell’ we will make free use of intrinsic differentiation
in the treatment of geometry of curves and surfaces.
If the vector field A* is defined in the neighborhood of C, as well as
on C, we can write
a
d ie, OA! j «,
1. Prove that 4 SAA) = 2¢;;A° ape
aA, A;
2. Show that A, ,— 4;; = weep
d os a i hae OB!
3. Show that i SAB) are Bi + 9,;Aé ae
a a agi
5. Show that =,(g,,4‘B’) = A,B + A'B,,,.
6. Prove that if A is the magnitude of A‘, then A , = A, ,A‘/A.
7. If y’ are rectangular cartesian coordinates, show that in E3
i ay? ay? y a ay? Oa?
Fig. 20
Sec. 48] PARALLEL VECTOR FIELDS 129
We can show, conversely, that every solution of the system 48.1 yields
a parallel vector field along C. Indeed, from the theory of differential
equations it is known that this system of three first-order differential
equations has a unique solution when the values of the components At
are specified at a given point of C. But it was shown previously that the
vector field formed by constructing a family of vectors of fixed lengths,
parallel to a given vector, satisfies the system. Hence every solution of
equation 48.1 satisfying the initial conditions must form a parallel field
along C.
Let A‘(t) and B‘(t) be any two solutions of the system 48.1. We verify
that the lengths of vectors A‘ and B* indeed do not change as we move
along the curve. Moreover, the angle 6 between the vectors A’ and
B‘ remains fixed as the parameter ¢ is allowed to change. To prove
this we note that (Sec. 44) A- B = ABcos 6 = g,;A‘B’, and, if g,,A'B’
d = Sa
is to remain constant along C, then i (¢,,A'B’) = 0. But g,,A‘B’ is an
invariant, and, since the g,; behave like constants in the process of co-
variant differentiation, we can write
d Aig ep
ne
—(g..A'B’)) = —pe tsALB" )
g,OA
== 813 Bi
St Bor 834
; OB"
Ot e
0 and
Since, by hypothesis, the fields A’ and B’ satisfy (48.1), 64*/6t =
follows
6B‘/dt = 0, and we conclude that g,,;A‘B’ is constant along C. It
that, if A‘ = B’, then g,,A'A’ = A? is constan t
directly from this result
along C, and this implies that 6 = constan t.
extended to
The notion of a parallel vector field along a curve can be
over three- dimens ional Euclid ean manifolds.
define parallel vector fields
P(x) and a vector A localiz ed at P. If we construct
Thus consider any point
magnitude and
at every point of the manifold a vector equal to A in
vector field in the
parallel to it in direction, there will result a parallel
passing through P,
space of three dimensions. If a curve C is drawn
l field along C,
the vectors 4‘ of the field lying on C will form a paralle
A‘ are defined at
and will thus satisfy (48.1). However, since vectors
every point («*) of the manifold, we can write
da’
‘dt
_ aAtde
dx" dt’
so that equations 48.1 assume the form
0A’
phale
i AZ \=
=— =)0;
ice¥" bed dt
130 GEOMETRY [CHaP. 3
This must be true for all curves passing through P, that is, for all values
of dx"/dt. Accordingly, the parallel vector field in E; satisfies the system
of equations
oA + [lato or A, =0.
Cx ak)
The converse follows, as previously, from the existence and uniqueness of
solutions of such systems of differential equations.
The condition for a parallel displacement of a covariant vector A; is
(49.2 =) te Ae er
dx dx?
Say
) Node cat
From (49.1) we see that
49.3 tee
( ) eet ds ds
(49.4) ~ByAtd? = 1.
Thus the vector A, with components A‘, is a unit vector. Moreover, A is
tangent to C, since its components 4‘, when the curve C is referred to a
rectangular cartesian reference frame Y, become A‘ = dy'/ds. These are
precisely the direction cosines of the tangent vector to the curve C. We
shall assume throughout this discussion that the curve C is of class C2,
so that it has a continuously turning tangent at all points of C.
Consider a pair of unit vectors A and p (with components 4‘ and py‘,
respectively) at any point P of C (Fig. 21). We suppose that A is tangent
SEc. 49] GEOMETRY OF SPACE CURVES 131
A+ dr
Fig. 21
(49.7) —,
w=-[i=je)
x OS
where x > 0 is so chosen as to make yw’ a unit vector.
the principal
The vector yw’, determined by the formula 49.7, is called
e of Cc
normal vector to the curve C at the point P, and ~ is the curvatur
at the point in question.
normal
The plane determined by the tangent vector A and the principal
vector p. is called the osculat ing plane to the curve C at P.
132 GEOMETRY [CHaP. 3
and we can treat this quadratic relation just as we did g,,4'A’ = 1 and
, WU
Ou ; : , OH j
deduce the orthogonality of vectors ae and pw’; that is, g,,u Fre 0.
Moreover, differentiating intrinsically the orthogonality relation 49.6,
we get
jae ny
8ij 5s fad T 833
rh aon ot A 3s = 9,
or
ieAne te
813 3s Sis 5 he
= — xg; jw?
— —— 0
where we used equation 49.7 and the quadratic relation 49.8. Thus
(49.9) ie,
Os
and, since g,,A‘A’? = 1, we can write (49.9) in the form
bu? ;
which shows that the vector — + xd?’ is orthogonal to A*. This result
s
shows that if we define a unit vector v, with components »’, by the for-
mula
(49.10) y= 2 (2 = xi’),
= : dx’
it follows that e€,;, = Vg é,;, 18 an absolute tensor, and hence the left-hand
member of (49.11) is an invariant. An algorithm of division suggests
that v* in (49.11) is determined by the formula
(49.12) ye = eA,
where A; and w; are the associated vectors g,,A* and g;,u*, and
elik = 1 tik
/&
is an absolute tensor. The validity of this expression follows from an
observation that (49.12) satisfies the conditions of orthogonality g,,A‘r’ =
0, g,;40°v? = 0, and the equation 49.11 determining the orientation of the
unit vector v relative to A and w. The number 7 appearing in equation
49.10 is called the torsion of C at P, and the vector v is the binormal.
In order to reconcile these definitions with the usual definitions of the
principal normal and curvature given in elementary vector analysis, we
recall the formula 46.6, 0A/0x* = A%.a,,
Fame)
and note that if the vector field
A is defined along C, we can write
@Ada’ ya da
DI ERO
: 7 ce eee OA ts
Using the definition of intrinsic derivative, —— = A%;—— , we can write
Os eas
the preceding result as
aA pA*
49.13 peony a,.
( ds Os
ds ,
dv) —=—a,=C,
dh OMe :
14 Gai
ds me ds Os
Ce
wr:
= pa,
where, in the last step, we have made use of the formula 49.7.
(50.2) as =v — xi.
Os
The first of these gives the rate of turning of the tangent vector A as the
point moves along the curve, and the second that of the principal normal
w. The third formula,
For the «**’s are constants in a cartesian system, hence ¢'}* = 0, and this is a tensor
: ik? . . Soh . .
equation!
Sec. 50] SERRET-FRENET FORMULAS 135
oe [iS —rp.
ds ds
Except for position of the curve C in space, the system 50.5 determines
the curve uniquely when continuous functions x(s) and a(s) are specified
along C.
We conclude this section by considering an example illustrating the use
of Frenet’s formulas. Consider a curve, defined in cylindrical coordinates
by equations
i = a
x? ==. 0s),
Teas
of arc in
This curve is a circle of radius a. The square of the element
cylindrical coordinates is
ds? = (dx)? + (a1)*(dx?)? + (dx*)’,
and it is easy to
so that g,, = 1, goo = (#')’, 833 = i 2, = 0, fi,
verify that the nonvanishing Christoffel symbols are
bal=-# (nd bl =
22
a,
12
—
21
C are A‘ = dzx'/ds,
=—,
foe
wit Ey
ds
(Medds (ada
\jk \22)” ds
2,a
> dA? (2 ede aga
Se 1? — = 0,
eds | \yihamedeetmal St ees
3 ke
a = aM (2|ige 0.
ds jk ds
C2] = 6 g* = KO.
Show that the tangent vector 4 at every point of C makes a constant angle with
the direction of the X-axis. Consider C also in the form y! = acos@, y? =
asin 0, y> = k0, where the coordinates y’ are rectangular cartesian.
2. Show that
OPA?
istieidss
s akae gem
Out = dr Pa
ie
pale ee deay Whas (x?2 + 7*)u
Das 8 es
St yt
a ae t(xA* — rv?) —— wt
3. Using results of Problem 1, show that the ratio of the curvature x to the
torsion + is a constant. Show from Frenet’s formulas that whenever 7/x =
constant, and the coordinates are cartesian, »* = cA? + b*, where c and b? are
constants. From this result it follows that A4*b* = constant, so that the curve
makes a constant angle with the lines whose direction ratios are b*. In other
words, the curve is a cylindrical helix. This theorem is due to Bertrand.
4. When C is specified in the form
Cc: yi =y(s),
where the y are orthogonal cartesian coordinates and s is the arc parameter,
show that
x = [(y")"F + (y)’F + [PF
SEC. 51] EQUATIONS OF A STRAIGHT LINE ibe
and
(yy (y?) (y3)’
rx = |(y4)"(y*)"(y*)” |. .
Gy) "G77"
y’ = (0),
Cc: (y? = (0),
y* = ka, .k = constant,
where o is the arc parameter of the directrix curve C’ in the y!y?-plane, so that
(ds)? = (dy)? + (dy)?. Note that (ds)? = (1 + k?)(do)? and show that
¢ wy ik
gp” y” (9)
p” yn (0) 1
PE OA
Pies Vie
igh
Lt k* >
s being the arc parameter. If the vector field A’ is parallel, then it follows
from Sec. 48 that 6A‘'/ds = 0, or
dA’ i dix?
atl as ae = 0.
GED ds ap ds
the straight line, so that the totality of tangent vectors A forms a parallel
vector field. Thus the field of tangent vector A* = dx‘/ds must satisfy
(51.1), and we have
Si eae
és | ds af) ds ds
The equation
ax’ |i ee dx?
(51.2)
ds” aB) ds ds
is the equation sought. In cartesian coordinates the Christoffel symbols
vanish and we obtain the familiar form of differential equations of straight
lines. From the geometric interpretation of the curvature x as a measure
of the rate of turning of the tangent line to a curve, we are led to define
the curvature of a straight line to be zero. This definition is consistent
with the first of Frenet’s formulas 50.1.
where u,! < u! < u,! and u,? < uw? < u,”, and the y’ are real functions
of class C1 in the region of definition of the independent parameters
u', u®. In order to reconcile these two different definitions we shall require
that the functions y‘(u1, u?) be such that the Jacobian matrix
be of rank two, so that not all the determinants of the second order
selected from this matrix vanish identically in the region of definition of
parameters u‘. This requirement ensures that it is possible to solve two
equations in (52.2) for wu! and u* in terms of some pair of variables y’,
and the substitution of these solutions in the remaining equation leads to
an equation of the form y* = y°(y', y*). It should be remarked that,
if any two determinants formed from the matrix 52.3 vanish identically,
then the third one also vanishes, provided that the surface S is nota
plane parallel to one of the coordinate planes.
Since u! and w? are independent variables, the locus defined by equations
52.2 is two-dimensional, and these equations give the coordinates y’ of a
point on the surface when u! and wv? are assigned particular values. This
point of view leads one to consider the surface as a two-dimensional
manifold S imbedded in a three-dimensional enveloping space E3. We
can also study surfaces without reference to the surrounding space, and
consider parameters u' and u? as coordinates of points in the surface.
A familiar example of this is the use of the latitude and longitude as
coordinates of points on the surface of the earth.
If we assign to wu! in (52.2) some fixed value u* = c (Fig. 22) we obtain
as a locus the one-dimensional manifold
y* = y(c, u’), (i aa is 2, 3):
We
which is a curve lying on the surface S defined by equations 52.2.
the v2-curve . Similarl y, setting vu? = constant in (52.2)
shall call this curve
u’ and
defines the w!-curve, along which only wu’ varies. By assigning to
on the surface,
w2 a succession of fixed values, we obtain a net of curves,
140 GEOMETRY [CHAP. 3
Fig. 22
(52.4) : weroet
Tian igFigg Te)
where the u*(a',u”) are of class C! and are such that the Jacobian
O(u, u? esti s: ; d
J= a does not vanish in some region of the variables #*, then one
Tih
can insert the values from (52.4) in (52.2) and obtain a different set of
parametric equations,
(62:5) y’ = f*(, u?), (i = 1, 2, 3),
where the u*’s are the Gaussian coordinates covering S. Viewed from
the surrounding space, the curve defined by (53.2) is a curve in a three-
dimensional Euclidean space and its element of arc is given by the formula
(53.3) ds? = dy’ dy’.
du* du?
= d,s, du du’,
where
dy’ oy’
53.5 aap = —
( ) Out au?
The expression for ds’, namely,
du* = a du’,
ou’
and hence (53.6) yields
If we set
: Ou duP
a.s 76 = A,,; —ya—
=F On due”
we see that the set of quantities a,, represents a symmetric covariant
tensor of rank two with respect to the admissible transformations 53.7
of surface coordinates. The fact that the a,, are components of a tensor
is also evident from (53.6), since ds* is an invariant and the quantities
a,3, are symmetric. The tensor a,; is called the covariant metric tensor
of the surface.
Sec. 53] INTRINSIC GEOMETRY 143
Since the form 53.6 is positive definite, the determinant
Gy, Aye
a= > 0,
Gs, Age “
and we can define the reciprocal tensor a*’ (see Sec. 30) by the formula
a“’a,, = 63. Thus we have
qu a 22. Ges gst
9 =a 12 a esa
a a a
The contravariant tensor a*’ is called the contravariant metric tensor.
We can repeat, almost verbatim, the contents of Sec. 44 concerning
metric properties of our two-dimensional space S. Thus the direction of a
linear element in the surface can be specified either by the direction
cosines dy'*/ds, (i = 1, 2, 3), or by the direction parameters
du*
(53.8)
53.8 jes Fe
For,
dy!_ dy‘du’
ds Ou" ds
and the du*/ds are uniquely determined when the direction cosines dy'|ds
are specified, and conversely. We define the length of the surface vector
A*, that is, the vector determined by A}(w’, u2) and A(u1, u?), by the
formula’
A = Va,,A7A?.
It follows from (53.6) that
du* du?
1 =a agp oe
ds ds
= ayaa,
where the subscripts | and 2 refer to the elements of arc of C, and C,,
respectively. Using the definition 53.8, we can write the unit vectors in
the directions of the tangents to C, and C, as
w= d,u* 2 _ dou"
ds,” # ds,
y?
Fig. 23
Sec. 54] ELEMENT OF SURFACE AREA 145
and
Oy? ay
ess B
(54.3) ohm 1
du" ine Out 5
Inserting in (54.2) the expressions from (54.3), we get
cos 9 = Oy’ 40 P
du* du?
and since
a Oy’ oy
du* du?
the foregoing expression can be written
(54.4) COs 0 = a,,A7y".
If the curves C, and C, are orthogonal,
(54.5) fAte 0.
In particular, if the surface vectors 4* and yw” are taken along the coor-
dinate curves (A) = Vay, A= Oust =); 1° 1/~/a9), then it follows
from (54.5) that the coordinate curves will form an orthogonal net if, and
only if, ay. = 0 at every point of the surface.
We can give a pictorial interpretation of these results in the manner
of Sec. 45. Thus, if r denotes the position vector of any point P on the
surface S, and the b, are the unit vectors directed along the orthogonal
coordinate axes Y, then equations 53.1 of the surface S can be written in
vector form (see Fig. 24) as
r(ui, u?) = by*(ul, uv).
Ye or
Vayl
Fig. 24
146 GEOMETRY [CHAP. 3
Or mor
ds?
= dr- dr = — - — du* du?
; : du* du? :
= 4, du* du®,
where
Or or
54.6 au = ——",
( ) ’ aut dub
Setting Or/du* = a,, where a, and a, are obviously tangent vectors to the
coordinate curves, we see that
In the notation of (54.3) the space components of a, and a, are &* and 7',
respectively.
We can define an element of area do of the surface S by the formula
do = |a; X a,| du du*,
This formula has precisely the same structure as the expression 44.11
for the volume element.
It follows from Sec. 40 that the skew-symmetric e-systems, in a two-
dimensional manifold, can be defined by the formulas
€,pA“u? = sin 8,
which is numerically equal to the area of the parallelogram constructed
on the unit vectors 4* and u*. It follows from this result that a necessary
and sufficient condition for the orthogonality of two surface unit-vectors A*
and y* is |e,,A*u"| = 1.
Sec. 55] FUNDAMENTAL CONCEPTS OF CALCULUS 147
Problems
1. Show that the cosine of the angle 6 between the coordinate curves wand u
on S is cos 9 = ays/ V.a4,499.
2. Find the element of area of the surface of the sphere of radius r if the
equations of the surface are given in the form:
yi =rsinuwicosuv, y2=rsinuwsinuw, y? =rcosu,
where the y’ are orthogonal cartesian coordinates. (Note that in this case
ayy = fetch ay = 0, Azo = r? sin? us)
t in the interval t, < t < to, vanishes for every choice of the function E(t)
of class C” in ty [t < be, and which is such that &(t,) = &(t.) = 0, then
M(t) is identically zero in the interval ty < t < tp.
We shall prove the lemma by assuming that M(t) # 0 and reaching a
contradiction. Assume M(t) #0 at some point t’ of h}<< t<h, and
suppose that M(t’) > 0. Since M(t) is continuous, there exists a number
such that M(t) > 0 in the interval (¢’ — 6,t’ + 0). Define a
6>0
function &(t) as follows:
E(t)=0, int, <¢t<7, wheres7, =?’ — 6,
Ef) =0, ntzst Sh, where tr, = t’ + 0,
["somear=["someoar>
te
ty
T2
Ti
0.
Thus we reach a
since the integrand is always positive in 7, <1? < 72.
¥ 0 is not tenable.
contradiction, and hence our assumption that M(t)
x = f(t), hitch,
called an extremal for the integral 56.1, such that J(x) for x = f(t) assumes
an extreme value in comparison with the values given to J by the admissible
functions in a sufficiently small A-neighborhood of the function x = f(t).
In other words, admissible functions x(t) are such that |z(t) — f(t)| <A
for t; <t<t,. We shall deduce next a necessary condition for an
extremum of J. Consider a function &(t) of class C?, such that &(t,) =
&(t.) = 0, and form a set of functions
Since &(t) satisfies the restrictions imposed on &(t) in the lemma of Sec.
55, we deduce from (56.4) that a necessary condition for an extremum
of (56.1) is that a(t) satisfy the differential equation
d’x dx
56.6 Fiz — + Fer — + Fu — Fr = 0,
ee dt? 7S tals
and &
where the subscripts denote the derivatives of F(t, x, 2) with ¢, x,
regarded as the independent variables. In order to determine x(t) we
must solve this ordinary differential equation subject to the end conditions
a(t,) = 2, and 2(t,) = 22. Equations 56.5 and 56.6 were first deduced
by Euler and are called Euler’s equations.
The expression (see equation 56.2)
d®
Wie)
(57.5) 7 = 0.
e=0
It follows that
te
(S7.6)oy =e } [(Foté! + Fin€!) +--+ 4+ (Fon€" + Fané")] dt = 0,
and the integration by parts gives
te
$b Fand®
tg
OJ = «|Fad
hy dt ty dt
“To ensure the independence of the integral 57.1 of special modes of parametriza-
tion, we suppose that F(t,x,) is positively homogeneous of degree one in the « (See
Sec. 43).
Sec. 57] EULER’S EQUATIONS 153
Since the &* are arbitrary and vanish at the end points of the interval,
we conclude from the fundamental lemma that
d
(57.7) Fi meine Gi= oa eee Pe -
or
Fa — @Fa3 — GF gta Fy, = 0.
relation
in which the variables are constrained by the
(57.10) A(t, x}, x?) = 0.
We suppose that the extremal,
xi=ai), t<t<t G=1,2),
154 GEOMETRY [CHAP. 3
satisfies the end condition of the type 57.8. When the constraining
condition 57.10 is written in the form
(57.11) $a,
y,2)=0
by setting « = t, y = x!,z = 2”, equation 57.11 can be thought to repre-
sent a surface referred to a set of cartesian xyz-axes. The extremal must
lie on this surface and we suppose that (57.11) can be solved for z in the
neighborhood of the extremal to yield a differentiable function
(57.12) z= f(@, y).
The substitution from (57.12) in (57.9) then yields an integral of the form
2
(713) J =| F(x, y, y’) dx,
xy
in which the variables x and y are independent, and we can obtain the
Euler equation by minimizing (57.13) on a set of admissible paths that
satisfy the end conditions y(x%,) = y;, y(%2) = yy. This is the problem of
the free extremum already considered in Sec. 56.
However, such reduction of the problem of constrained extremum to
the problem of a free extremum of the functional 57.13 is usually incon-
venient because an explicit solution 57.12 of equation 57.11 may prove
unwieldy. In this event we can follow a procedure similar to that of the
Lagrange multiplier method of obtaining the relative extreme values of
functions of several variables constrained by relations of the type 57.11.
We suppose that d¢/dz ¥ 0, so that it is theoretically possible fo obtain
the solution of (57.11). If this is so, (57.9) can be rewritten in the form
OF
ay a + Fete
Sec. 57] EULER’S EQUATIONS 155
F, + (F, dF ,
“)— r=
dF,’
0.
hh dx dx
Thus
dFy _ F,
(57.17) oe
TS
dx
dy + bf, = 9,
so that
(57.18) f= _ by
d:
extremal,
The expressions for f,, given by (57.18) and (57.17) along the
represent the same function of x; hence
oe ES = pp
(57.19) ee (2),
dy $.
*
where A(x) denotes the common value of the ratios.’
for the extremum
It follows from (57.19) that the necessary conditions
of the integral 57.9 are
which have the structure of Euler’s equations 57.7 for the variational
problem associated with the free extremum of the integral
2. Note the hint in Problem 1 ae show that the Euler equation 56.5 can be
written in the form
where g,; = 2); are specified functions of the variables x’. We suppose
that the form 58.1 is positive definite and the functions g;; are of class
C2. The length of a curve C, represented in R,, by equations
C: act a a(t), ty < t & tp,
is given by
fess SS
(58.2) S “AyJ gapt?a? dt, (x; P= .1, 225,57)
cs in R,,.
The extremals of the functional 58.2 will be termed geodesi
The function F of Sec. 57 in this case is
(58.3) F=V2,,0#,
and Fy. This
and, to form Euler’s equations 57.7, we need to compute F,,
computation is straightforward. We deduce from (58.3) that
anil a+ _ OLa8 a +B
= 5 (8ap@'®') auf toe
Fai
and
Fut = (gop)
See?
BastSof
Substituting these expressions in Euler’s equations yields
os gtak
d g ge x
28a8 ea?
4 (se) BCs SIN
dt\ds/dt 2ds/dt ;
158 GEOMETRY [CHAP. 3
and, carrying out the indicated differentiation, we obtain
a 2 2
O85 a 1 28up eegh — Sajt" s/dt .
ax? 20%) ds/dt
Since the second term in this equation can be written as a sum of two
terms, we have
; a 72 2
Ce ae (3 O86; Seng ciate? = Saji d s/dt z
ab= V gapiti? = 1,
dt
the system 58.5 simplifies to read
(58.6) Gast" + [oP jan? =:
In equation 58.6, dots denote the differentiation with respect to the arc
parameter s. ;
If we multiply equation 58.6 by the tensor g*’ and sum, we obtain a
simple form of the equations of geodesics in R,,.
Fig. 25
and
d’y®
= 0, for y = 2.
ds?
p. 175.
14 See, for example, L. P. Eisenhart, Differential Geometry (1940),
160 GEOMETRY [CHAP. 3
We thus obtain
o6 = As + B,
Y= As
sp an Cyo a C2,
and hence the curve is a helix, whose pitch is determined by C,. The
constant C, determines the origin for the arc parameter o.
We can show in a similar way that the geodesics on the surface of a
sphere are arcs of great circles. (See Problem 1.) However, as an illustra-
tion of the use of equations 57.23 we consider the functional
to —
d dx*
58.12
(58.12) ——— tO
nade +4 24()z'=0, (i 2)
(Le = 1,2,3),
where ds = V/#'## dt.
On eliminating A(t) from the first two equations in the set 58.12, we get
or
SEC. 58] GEODESICS IN &,, 161
Therefore
(58.13) a Peg he
ay 2
(58.14) CLI NE Ao
1 3
where C, is a constant.
From (58.13) and (58.14) we find
robes3 dx 1 i
Bias 3
ee 2
dx iE Teat akee2
(xt (x)?
Or
3)\ 2
Christoffel symbols |
ifvanish at P, then for that particular point
: ‘D2, B
(59.3) Lg poo = 0
dv* dv" — ul By) Av* dv"
15 See also Theorem I, Sec. 39.
Sec. 60] PARALLEL VECTOR FIELDS IN A SURFACE 163
We exhibit next a solution of this equation yielding a particular trans-
formation 59.1 to a coordinate system v% in which the Christoffel symbols
vanish at P. It is the second-degree polynomial
(59.4) WP YE pt ae |poh,
2 law P
where the wp* is the value of u* at P and the ff are the values of the
LM) Pp
Christoffel symbols at P. To verify that (59.4) satisfies (59.3), we compute
Ou" {x
f29).)) — = 0, — ‘
Ov" i Aw -
and
(59.6) Cie
2,,0
ee ye)
v* Ov" lau) Pe
From (59.4) we see that the point P, in new coordinates, is given by
ed
aOt = ae
a a y
(60.1) dt py dt
16 A derivation of explicit equations of transformation for this case, which include
(59.4) as a special case, was given by Levi-Civita in a paper entitled “Sur l’écart
géodésique,” Mathematische Annalen, 97 (1926-27), pp. 291-320.
164 GEOMETRY [CHapP. 3
This is identical in form with the left-hand member of equation 48.1
defining the parallel vector field along a space curve. Accordingly, we
take the differential equation
ds \ByJ ds
and if A* is taken to be the unit tangent vector to C so that
(Ca 12
(61.2) ds* = a,, dv" dv® = aa” y+ (vy1y2 (doy,
77,.2)2
v)*—a
so that
(v')?
Since this is identical with (61.3), the surfaces S, and S, are isometric.
It follows from discussion-in the next section that these surfaces are not
developable.
of the surface S, we can form the Christoffel symbols with respect to this
surface, and the corresponding Riemann tensor
(62.2) Rapys =
= a st a Pare ‘
[By,«] [Bd, «] [xy, A] [«d, A]
We recall that this tensor is skew-symmetric in the first two and last
two indices, so that, for the surface S,
(62.4) K= Ria :
a
curvature
where a = |a,,|, and call it the Gaussian curvature or the total
s a,, are involved in this
of the surface S. Since only metric coefficient
of the
definition, the properties described by K are intrinsic properties
- ryt
surface S.
If we introduce the two-dimensional e-tensors,
aa ap
is an invariant.
These equations show that the Gaussian curvature
the Euclidean plane, there
Now, when a surface S is isometric with
respec to which ay; = 42 = its
t
exists on S a coordinate system with
R,»,5 = 0 in this particular coor-
ay. = 0. It is obvious that in this case
vanish in every coordinate
dinate system, and since R,,,9 18 a tensor, it must
system.
at all points of the surface,
Conversely, if the Riemann tensor vanishes
there exist coordinate systems
Theorem II of Sec. 39 guarantees that
1, 412 = 0.
on the surface such that ay, = 422 =
168 GEOMETRY [CHAP. 3
Thus we have a
THEOREM. A necessary and sufficient condition that a surface S be
isometric with the Euclidean plane is that the Riemann tensor (or the
Gaussian curvature of S) be identically zero.
Consider next an invariant
(62.7) R=a"R,,,
where
(62.9) R= a@Rieaeo Th
and recalling (62.3) we see that (62.9) is equivalent to
Since
Aub es Ag2 qe = ay ae! ae
ie a a
we have
(62.11) R= —2 Rime,
a
Comparing (62.11) with (62.4) we see that
R= —2K.
1. Use formulas 62.2 and 62.4 to show that, if the system of coordinates is
- sates) +2
orthogonal, then
y¥? = flv),
y> =u,
where f; and fy are differentiable functions, is developable.
in the
6. Show that the formula for the Gaussian curvature K can be written
form
re 1 d |Ayo 9A, 1 “|
2Va\lawla,Va aw Va au
d |2 ay» 1 Gay Me au |
Se ies .
+ _——w ee
(63.3) I =
170 GEOMETRY [CHaP. 3
(63.4) — = x17,
where x, is a suitable scalar. In order to determine the direction of 7*
uniquely we choose 7% in the way analogous to the choice of the triad
of vectors in Sec. 49 (equation 49.11), namely, €,,A%7° = 1. This choice
of the orientation of A and y uniquely determines the sign of ,, and it
amounts to saying that the sine of the angle between A and n is +1. The
vector 7* is the unit surface vector orthogonal to the curve C, and the
scalar x, is called the geodesic curvature of C.
We recall that the equation of the geodesic on S (see equation 60.4)
can be written as 6A*/d6s = 0. Comparing this with (63.4) leads to the
conclusion that, if the geodesic curvature x, = 0, then the curve C is a
geodesic, and conversely. Hence the
THEOREM. A necessary and sufficient condition that a curve on a surface
S be a geodesic is that its geodesic curvature be zero.
As an illustration, we compute the geodesic curvature of the small
circle
Ge u*= constant =1,' > 0, u* = 7,
on the surface of the sphere (Fig. 26)
S: y' = acosu'cos v2
y? = acos ui sin uv?
y® = a sin ul.
If the arc-length s of C is measured from the plane u, = 0, we have
u® = s/(a cos up'), and the equations of C can be written in the form
(63.5) th ce Wes oh oe
a COS Uy’
From (63.5) we find that the components of the unit tangent vector
A* = du*/ds along C are
(63.6) oe ee
so that
bas dat |1 |
ga dul
ds ds ap
SEC. 64] SURFACES IN SPACE 171
y3
ey
Shona
Y 1
Fig. 26
and
6?
— =
di”
— +
Oa au
Qo) 25 —i() 2 \aaat
Os ds a ds 4 22
find, on referring to Problem 3 of Sec. 31, that ke = sin u' cos ul,
2
|= 0. Accordingly, formulas 63.4 yield
Ze
6, 1 62°
(63.7)
63.7 ee =x
On amg ee a
gl =—tanu,, —=x,7 me = 0.
7°
Since C is not a geodesic, x, ¥ 0, and we conclude that 7? = 0. But
1/a. Hence
is a unit vector so that a,,7%n° = 1, and we find that 7! =
the relation
equations 63.7 yield x, = (tan uy!)/a. In Sec. 71 we establish
of C.
between the geodesic curvature x, and the ordinary curvature x
dy* dy*
where g;; = ant Dal” and the set of equations 64.1 for the surface S
can be written as
[64.5] ee dae.
Ou”
The left-hand member of this expression is independent of the Greek
indices, and hence it is invariant relative to a change of the surface
coordinates u*. Since du* is an arbitrary surface vector, we conclude that
Ox'
(64.7) ou
is a covariant surface vector. On the other hand, if we change the space
coordinates, the du*, being a surface vector, is invariant relative to this
change, so that (64.7) must be a contravariant space vector. Hence we
can write (64.7) as
(64.8) “=
ou"
of
where the indices properly describe the tensor character of this set
quantities.
174 GEOMETRY [CHAP. 3
A simple geometrical significance of the set of quantities 64.8 can be
deduced from Fig. 27. Let r be the position vector of an arbitrary point
P on S. The point P is determined by a pair of Gaussian coordinates
(u!, u), or by a triplet of space coordinates (a1, x*, x*). Accordingly,
the vector r can be viewed as a function of the space variables x* satisfying
equations 64.4. Thus
(64,9) Or, aOreos
Ou* E Ox! Ou™
But dr/dx* are the base vectors b; at P, associated with the curvilinear
system X, whereas Or/du* are the base vectors a, atP relative to the
Gaussian system U.
Hence equations 64.9 yield
(64.10) a, = b; —
3 3 : Ox? ae a
It is clear from this representation that a, = — b, and a, = aA b,, so
dul u
that dx‘/du* = x,', (« = 1,2), are the contravariant components of
the surface base vectors a, referred to the base systems b;. Thus the
sets of quantities
4p ( Ox” =) ; Ox* =)
Uy: oe ae and x3: |—,-=-,->
du’ du: du du? du?’ du?
Fig. 27
SEC. 65] THE NORMAL LINE TO THE SURFACE 175
transform in a contravariant manner relative to the transformation of
space coordinates 2°.
We can also show that the three surface vectors
(64.11) aa!
aut
_ Ox! aut
du’ du*
But dx‘/du’ =z, and (64.11) yields, for i = 1, 2, 3, a! = (00° /du")z5.
This is the covariant law.
Let ds be an element of arc joining a pair of points P(u’, u*) and
P(u* + du}, v2 + du?) on S. The direction of the line element ds is given
by the direction parameters du*/ds = 4*. The same direction can be
specified by an observer in the enveloping space by means of three param-
eters dx'/ds = 2‘, and it follows from (64.5) that 2’ = x A*, This formula
tells us that any surface vector A* (that is, a vector lying in the tangent
plane to S) can be viewed as a space vector with components A‘ determined
by
(64.12) Ate AC
We shall refer to a vector A‘ determined by this formula as a tangent
vector to the surface S.
Fig. 28
If we define
ao ee ee
= 2\duz du® du® aut)’
so that (65.4) reads
(65.5) dn- dr = —b,, du* du’,
the left-hand member of (65.5), being the scalar product of two vectors,
is obviously an invariant; moreover, from symmetry with respect to «
and 8, it is clear that the coefficients of du* du’ in the right-hand member
of (65.5) define a covariant tensor of rank two. The quadratic form
(65.9) NE ap = €iintaXp-
Multiplying (65.9) through by «*’, and noting that e~Pe,, = 2, we get the
desired result
(65.10) iq t = ge Me 5 XaXp-
It is clear from the structure of this formula that n; is a space vector
which does not depend on the choice of surface coordinates. This fact
is also obvious from purely geometric considerations.
C=A xX B
Hence C, = 0, C, = 0, C; = ABsin §. Thus the C; define the vector
A* and
normal to the plane determined by A and B whose magnitude is AB |sin 6|. If
This result follows
BP are the surface components of A and B, then ABsin 6= €qpA=BP.
ely from the formula for the sine of the angle between two vectors given in
immediat
Sec. 54.
178 GEOMETRY [Cuap. 3
concept of tensor differentiation was introduced by A. J. McConnell,
whose elegant treatment of surfaces is followed closely in this and several
other sections of this chapter.”°
Let us consider a curve C lying on a given surface S and a vector “a
defined along C. If tis a parameter along C, we can compute the intrinsic
derivative 6A’/6t of A*, namely,
66.1
bai _ Ai [i\sda
Cae Ot dt alikI dt
Dt) =) ALB “
ee
; j a y
g\tJ dt dt a By dt
es
ee
ST eed Te
$f a di
i),dz® ft
(ree
@aslikl “dt calay)
ee
(6)_,dw’
* dt
If the field 7,’ is defined over the entire surface S, we can argue that,
os (tr (2)
since
(eo?) ;
nay :Ou’
: ;
9g jk om
a alny
| 26 By | ;:
a|
180 GEOMETRY [CHAP. 3
07x" a e {6
67.1 at = eat |)let \x5,
re) Fou" due 4 jk ne ala €
from which we deduce that
(67.2) oe = ee
-i(2. Or , On 2),
CPi
Ou" due ou? Ou"
note that the vectors n and a, = Or/du* are orthogonal and hence
e: = a, b,x;3
ou*
therefore
O’r Ox' Ob; ,;
bet ee
Ou" Ou du’ ~—-du?
pre Obs a,
“au? " Ox “Ay
= hb,
ae rat tsl) +
Ape
oe r)
where, in the last step, we made use of formula 46.4 for the derivative of
the base vector b,.
If we insert in the right-hand member of the foregoing expression from
equation 67.1, we get
(67.10) a
Parent = b,{ x a, e+ ae hal”?od
182 GEOMETRY [CHap.3
where nis the unit normal and r is the position vector of the point on the surface.
3. When the equation of a surface S, referred to a set of orthogonal cartesian
axes, is taken in the form y? = f(y', y”), we can write the parametric equations
of S in the form y' =uw,y? =v, y? = f(w, vu’). If partial derivatives of
fy’, y*) are denoted by fn =p, fe =] fry =" fyyt =5, fy: = ¢, then the
coefficients a,, in ds* = a, , du du are
a, =1+> p’, Q1. = pq, Qa. = 1 +4",
whereas the coefficients bap of the second fundamental form are
r Ss i
by =, SS by = —— -
Vit¢pt+”@ Vi+p?+¢@ Vitp+¢
Show this and compute a,, and b,, for the surface of the sphere
Show this. Use these formula to compute the a,, and b,, for the surface of
revolution y! = ul cos uv, y? = ul sin uw, y® = f(u!),
SEC. 68] THE INTEGRABILITY CONDITIONS 183
ha — bap’,
(68.1)
where
where R®,,., is the Riemann tensor of the second kind, formed with the
aid of the coefficients a,, of the first fundamental quadratic form. Equa-
tions 68.4 involve third partial derivatives of the coordinates x’, and we
shall assume from now on that the functions entering in (68.2) are of
class C*.
We shall see that the conditions of integrability 68.4 impose certain
restrictions on the possible choices of functions b,, and a,;. These
restrictive conditions are known as the equations of Gauss and Codazzi.
They will be derived in the following section.
(69.4) Xp = dapn’,
so that the substitution from (69.4) and (69.2) in (69.3) yields
bap + Big%a2ich = O,
and, since a,, = 2;;%,'%,’, we have
bap = —AgyCp”.
*® The treatment given here is patterned after A. J. McConnell, Absolute Differential
Calculus, pp. 201-205.
24
We recall that n,, i = Be
ont
oo :i nak,
j.
NS ay
SEC. 69] FORMULAS OF WEINGARTEN 185
Solving this equation for c,”, we get
Cp” = —a"'D 5;
(69.5) WalDy et
These are the Weingarten formulas which we will use in deriving the
Codazzi equations.
The equations we desire follow from the integrability conditions 68.4,
namely,
Hence
Since «, 6 assume values 1,2 and b,, = b,,, we see that there are
two independent equations of Codazzi and only one independent equation
of Gauss.?° The independent equations of Codazzi are
we get
db Obip |6 {i
69.12 —a2 _ —ah _ p, +b =0, «Ff,
Gua ous lap Maer
(no sum on @).
The equation of Gauss, on the other hand, is
b
(70.1) K —
5=
bir bo. ana bis"
7:
431499 — Ae a
Thus the Gaussian curvature is equal to the quotient of the discriminants of
the second and first fundamental quadratic forms.
We introduce next another important invariant H, called the mean
curvature of the surface. This is given by the formula
(70.2) H = 3a*b,,,
and we shall see in Sec. 72 that the invariants K and H are connected in a
remarkable way with the ordinary curvatures of certain curves formed by
taking normal sections of the surface.
*6 For a detailed discussion, see L. P. Eisenhart, Introduction to Differential Geometry,
pp. 218-221, where the case of cartesian variables x‘ is considered.
SEc. 71] CURVES ON A SURFACE 187
tS
\
Fig. 29
188 GEOMETRY [CHAP. 3
Fig. 30
190 GEOMETRY [CHAP. 3
Problems
1. Prove that the geodesic curvature x, and the curvature « of any surface
curve C are connected by the formula x, = sin 6, where @ is the angle between
the normal to the surface and the principal normal to C.
2. Consider the surface of the right circular cone
Sag) = wicosius
y2 = u sin u?,
Te.
and the curve
C2 fsa. a OnnSs
Write equations of C in the form ul = a, u* = s/a, where s is the arc parameter
and show with the aid of (71.6) that x, = V2/2a. Verify this result by (71.10).
3. Show that the parallels u’ = constant on a sufficiently smooth surface of
revolution
y> = uw cos u,
y? = usin v2
= f(u’)
are curves of constant geodesic curvature.
4. A curve C on a surface S is called an asymptotic line if b,,A*2°= 0 along
C. Show that the principal normal p to the asymptotic line is fnipent to S and
the binormal v is normal to S.
5. Show that the normal curvatures in the directions of the coordinate curves
are by/ay, and b59/d39.
6. Prove the theorem: If a curve is a geodesic on the surface, then either it isa
straight line or its principal normal is orthogonal to the surface at every point,
and conversely.
(7235) B — ab gd + ee 0,
a
where b = |b,,| and a = |a,,].
Since the Gaussian curvature K is given by
70.1] Kee
a
and the mean curvature H is
[70.2] H = 4a"b,,,
we see that equation 72.5 assumes the form
(72.6) — 2H + K=0.
The roots @ = xq) and 3 = x.) of (72.6) are called the principal curvatures
of the surface, and the directions 4% and 4%, corresponding to these
extreme values of x;,,, are the principal directions on the surfaces. We
leave it for the reader to show that these directions are real.
From (72.6) it is clear that the principal curvatures x.) and %.») are
related to the mean and Gaussian curvatures by the formulas
#1) + xX(2) = 2H,
(72.7) |
(1)%(2) = K.
From equation 72.3 it follows that the principal directions are determined
by
Hee = (1) Gap Aa) = 0,
(beg — %(2)4ap)A(2) = 0.
If the first of these equations is multiplied by 2%), the second by 4%),
and the results subtracted, we obtain
(72.8) (4¢2) — ay )GapAy
Alo)= 9.
192 GEOMETRY [CHAP. 3
bigdu? _ bapdul
aygdu® — dog du®
or
(72.10) (41412 — byp@y1)(du')® + (611422 — 9441) du’ du?
+ (by2422 — Ay2b2)(du*)? = 0.
At each point of S where either 6,, du* du? # 0 or b,; du* du® is not
proportional to a,,du* du’, equation 72.10 specifies two orthogonal
directions
(72.11) “~=¢,u',u),
u
(«= 1,2),
which coincide with directions of the principal curvatures.?”? Each equation
in (72.11) determines a family of curves on S covering the surface without
gaps. These two families of curves are orthogonal, and, if they are taken
as a parametric net on S, the first fundamental form has the form
(ds)? = G,,(dii*)? + Go.(diui?)*.
Accordingly, equation 72.10 in the coordinate system a* takes the form
—by2dy1(dit?)? + (by 1429 — by94,) di! dit? + by.d2.(di?)? = 0,
and its solutions are
ui = constant, u? = constant.
If we take di! ¥ 0 and di? = 0, we see that 5,, = 0, since 4,, ¥ 0. Thus
a necessary condition for the net of lines of curvature to be orthogonal
*7 We exclude those points on S at which x, =0 or x) = %@). See concluding
remarks in Sec, 71.
SEC. 72] PRINCIPAL CURVATURES OF A SURFACE 193
is that ay. = by = 0. Conversely, if a,, = by. = 0, then (72.10) has
solutions ut = constant, vu? = constant, so that the coordinate lines are
the lines of curvature. Hence a
THEOREM. A necessary and sufficient condition for the cootdinate net
on a surface S (other than a plane or a sphere) to be the net of lines of
curvature is that ay. = by, = 0 at allpoints of S.
We note that for every orthogonal net on a plane or a sphere a,. =
by. = 0.
Formula 71.9, for the normal curvature x,,), when the coordinate
system is taken to be the net of lines of curvature becomes
oe by,(du’)? a5 byo(du*)?
- dy,(du’)” + ay(du*)”
If we set dul = 0, du? ¥ 0, and du* = 0, du ¥ 0, we get
are two directions for which x;,) = 0. A point is parabolic if one of the
values 4) OF xg) is zero. In the special case x1) = %(2), all values of xn)
are equal and such points are called spherical. In the neighborhood of
a spherical point the surface looks like a sphere, and we can prove that
if all points of S are spherical, then the surface S is a sphere. In some
books spherical points are also called umbilical.
Problems
Yee
yaa,
y® = f(u', uv’).
3. Show that the helicoid
y =u cos u*,
y*? = uv sin v2,
y> = au
with f(r) of class C*. Prove that the lines of curvature on S are the meridians
¢ = constant and the parallels r = constant.
7. Refer to Problem 6 and show that the points on a surface of revolution
S for which f’f” > 0 are elliptic; those for which f’f” < 0 are hyperbolic; and
if f” = 0, then S is a cone.
SEC. 73] PARALLEL SURFACES 195
8. Let the vector equation of a curve C drawn on a smooth surface S be
r =r(s). Ifn(s) is a unit normal to S at a given point of C, and v is a parameter
measuring the distance along n, the vector equation R(s,v) = r(s) + vn(s)
defines a ruled surface S’. Prove that S’ is developable if C is a line of curvature,
and conversely. Outline of solution: Denote the coefficients in the second
fundamental form of S’ by dg: Compute the d,, from the formula
2R
dp = ea0
Fs N,
where v! = s, v? = v, and N is the unit normal to S’. Show that d,. = 0. The
developability condition d,,dy. — dj,” = 0 implies then that d), = 0. But, along
C, N = dr/ds x n; hence
dn dr
1 cae ae P n = 0.
[64.10] a, = b, og
Ou"
For simplicity in writing we introduce the notations (cf. Sec. 64)
(73.3) Ge = Ya + hn’,
so that
(73.5) 70, = 0,
On the other hand, the unit normal #; to S is orthogonal to the base
vectors y/, on S, so that
(73.6) ¥,n, = 0.
We conclude from (73.5) and (73.6) that the vectors n; and A, are
collinear, and, since they are unit vectors, n; = fj;.
The metric coefficients a,, of S are given by*®
ayp = Yn Gp's
H
a
(73:8)
, nin p =
5 aise a” bys fa" bay :
==" 0 ap’ BAs
since yiy', = a,,,- On the other hand, the Gauss equations 69.10 require
that
since
H = 3a”b,>.
[70.2]
the right-hand member of (73.9)
The substitution in the first term in
from (73.8) yields
(73.10) nin', = —Kazg + 2Hb«p,
in the form
and we can thus write (73.7)
so that
(73.14)
Seay Taia Fe2hH
:
°
pe H—hK
1 + h?K — 2hH
From the first of these elegant formulas it follows that when S is a develop-
able surface, then the parallel surfaces S$ are also developable.
Problems
1. If Sis a surface of revolution, show that a parallel surface S is also a surface
of revolution. Hint: Consider y! = u' cos u?, y® = wu! sin u?, u3 = f(u?).
2. Show that the principal radii R,,, of normal curvature of S are related to
the principal radii R,,) of a parallel surface S by
Ree
a
hs
Fig. 31
[63.4]
63.4 ry
a geodesic,
where ~, is the geodesic curvature of C. Moreover, if C is
then x, = 0 at all points of C, and conversely.
paragraph
Since 7% is orthogonal to /®, it follows from the concluding
of Sec. 54 that ¢,97%4" = 1, or
ee €aph?.
But from (63.4)
. on
pasill ar R.
or ;
on
Kg
os €
€ap
44Os
(74.1)
curve C gives
The integration of this expression over the
On
[ dss =| [/: phi —ree ds,
€,,A°
(74.2)
if they can be mapped into one another
2° Two regions are said to be homeomorphic
in a continuous one-to-one manner.
200 GEOMETRY [CHaP. 3
A(B)
Fig. 32
and when the line integral in the right-hand member of (74.2) is trans-
formed into a surface integral by Green’s formula, we find that*°
gp — pada Q9,.
t=1
of n sides,
When the vector A is propagated along a geodesic polygon
spherica l excess of the polygon®*
entirely similar computations yield for the
gy —p=2r—- D9,
i=1
and since the Gaussian curvature K for the sphere is equal to 1/R*, we
can write
ar 2H,
(74.5) K To eet
This formula can be generalized to obtain the Gauss-Bonnet formula
74.3 for the case where C in Fig. 31 is a geodesic polygon. Thus, if the
region D is subdivided by small geodesic polygons into subregions of
areas do,, the familiar procedures of integral calculus applied to (74.5)
yield
(74.6) |i:Kdo =2n — > 6.
This formula coincides with (74.3), since x, = 0 when C is a geodesic
polygon.
‘Formula 74.6, first obtained by Gauss, was generalized by Bonnet* to
yield the result (74.3), which, as we have already noted follows directly
from (74.2) on application of Green’s formula.
The left-hand member {|K do in (74.6) is called the integral curvature
D
of D. It turns out that the integral curvature is a topological invariant.
Two surfaces are said to be topologically equivalent if they can be mapped
into one another by a continuous one-to-one transformation. It can be
shown by using (74.3) that [| « do = 47 for all regular surfaces
D
topologically equivalent to a sphere and {|Kdo = 0 for all regular
D
surfaces topologically equivalent to a torus.**
s= |“Vg,Ada'ldty(dx'|dt) dt
manifolds
is the length of the curve C. There are definitions of metric
the element of arc in the form
which are not based on the expression for
75.1, but they need not concern us here (see Sec. 43).
on of the curve,
The vector 4‘ = dx'/ds will be said to define the directi
a unit vector. The
and it is clear that g,,A‘A? = 1, so that the vector Ai is
length of any vector A’ is given by the formul a
thes V 9,,A'Al
ensional manifold is a
The notion of the angular metric in an n-dim
of the angle in the three-dimensional
direct generalization of the definition
case.
the cosine of the angle
If A‘ and w* are two unit vectors, we define
between them by the formula
(75.3) cos 0 = g,,A‘u’.
204 GEOMETRY [CHAP. 3
It is not clear from this definition that the angle 6 is necessarily real. We
shall prove, however, that this is always so if the form g,; dx‘ dx’ is
positive definite. The proof follows at once from the Cauchy-Schwarz
inequality
(gisAin’)
(SijsAA Sie’)
and, since A’ and qm’ are unit vectors, we have (g,;A‘u’)® < 1, which
states that the angle @ in formula 75.3 is real.
We define the volume element in R,, by the formula
anes Vg dx! dx® ++» dx",
Now equations 75.5 together with 75.6 and 75.7 require that
pe Ox Ox
(75.8) ne ee
eu Ou! ©
The set of 4m(m + 1) partial differential equations 75.8 in n variables ae
will not be expected to possess a solution unless n > 3m(m + 1); thus,
if m= 2,n> 3; if m=3,n> 6, and so on. This estimate, however,
does not constitute a proof that an m-dimensional variety can be imbedded
in E,, whenever n > }m(m + 1). It is possible, however, to prove that a
neighborhood of R,, can be imbedded in E,, if n > 4m(m + 1). Con-
cerning the global imbedding of the whole of R,, in E,, almost no general
results are known. There are a few special theorems on the imbedding
of two-dimensional varieties with special topological properties. These
refer almost wholly to convex two-dimensional manifolds.** The problems
on imbedding now lie at the forefront of researches in geometry.
(77.1) Fi ay
dt
provided that our units are so chosen as to make the proportionality
constant equal to one.
If we postulate the invariance of mass, then equation 77.1 can be
written in the familiar form
(712) F = ma.
We note from (77.1) that, if F = 0, then d(mv)/dt = 0, so that my =
constant, and hence v is a constant vector. Thus the first law is a con-
sequence of the second.
SEC. 78] EQUATIONS OF MOTION OF A PARTICLE 209
The concept of mass can obviously be defined with the aid of the second
law in terms of force and acceleration. There were numerous attempts to
define mass and force independently of one another. The most familiar
of such definitions is due to Ernst Mach,! who formulated a definition of
mass with the aid of Newton’s third law of motion. In our opinion a fine-
grained analysis of Mach’s definition of mass reveals certain logical
difficulties which cannot be resolved by appealing to the third law alone.
For this reason it seems best to leave one of the fundamental building
blocks of mechanics (mass or force) undefined and admit it in the science
of mechanics on the same basis as the “God-given integers” in mathe-
matics.
The third law of motion states that accelerations always occur in pairs.
In terms of force we may say that, if a force acts on a given body, the body
itself exerts an equal and oppositely directed force on some other body.
Newton called the two aspects of the force action and reaction, whence the
usual statements of the law.
The entity of mass entering in the formulation of Newtonian laws is
sometimes called the inertial mass (or simply inertia) to distinguish it from
the gravitational mass M entering in the Newtonian law of gravitation.
This law states that the force of attraction between a pair of particles is
proportional to the product of their masses, is inversely proportional to
the square of the distance r between them, and is directed along the line
joining the particles. In symbols,
(77.3) F=k )
r
M, to
where k is a universal constant and r is a vector directed from mass
mass M,.
If it is assumed (as it is usually done) that the gravitational and inertial
ng
masses are equal, the law 77.3 furnishes a practical means for compari
masses with the aid of beam balances.
ng of
In order to develop the science of mechanics of a universe consisti
an laws the
more than two particles, it is necessary to adjoin to Newtoni
ions re-
principle of superposition of effects and make further assumpt
garding the nature of constraints.
by a vector Tr.
Let the position of a moving particle P be determined
of r are denoted by
If the curvilinear coordinates of the terminal point
survey of it is contained in
1B. Mach, The Science of Mechanics. An interesting
R. B. Lindsay and H. Margen au, Foundat ions of Physics.
210 ANALYTICAL MECHANICS [CHAP. 4
a(t), then the equations of the path C of the particle can be written in the
form
(78.1) Cr at = (5),
182
(78.2) eea
v= 2
(78.5) 4 Ook a
= F; dx’,
where the F; = g,,F’ are the covariant components of the vector F. We
shall suppose that, in general, the functions F'(x), defining the vector field
Sec. 78] EQUATIONS OF MOTION OF A PARTICLE 211
F, belong to class C1. The work done in displacing a particle along the
trajectory C, joining a pair of points P, and Pg, is the line integral
P2
(78.6) W=| F,dz’. :
Hea
Making use of Newton’s second law of motion 78.4, we can write 78.6
in the form
Pe i
w=| poy dl
Pi ot
(78.7) te ieee
=| ee v’ dt.
t ot
But
O(g;v'v’) , ou
Si Fieet 2
and, since g,,v'v’ is an invariant,
6(g;;v'v’) = d inj
ayes = iF (g,;0'v’),
and hence
d ee Oe
FACou, ) == 2 283;Tian
mea ron ds
w={ m © (g.viv’) dt
te d Bade
78.8
hE, ty 2 at
sli
ane Py
=i 7, — 13,
where
2
m °% v
T=—¢g,vv’
y o4 =
; —
> .
ing the
We have the result that the work done by the force F; in displac
differen ce of the
particle from the point P, to the point P, is equal to the
the beginni ng of
values of the quantity T= jmv* at the end and at
is exactly one-
the displacement. We define the quantity T = 4mv*, which
particle.
half of the vis viva of Huygens, as the kinetic energy of the
be enunci ated as a
The statement embodied in the formula 78.8 can
The work done in displacing a particle along its trajectory
TueoreM.
e.
is equal to the change in the kinetic energy of the particl
212 ANALYTICAL MECHANICS [CHAP. 4
It may happen that the force field F; is such that the integral 78.6 is
independent of the path. In this event the integrand F, dx’ is an exact
differential,
(78.9) dW = F, dz’,
of the work function W. The negative of the work function W is called the
force potential or potential energy. We denote the potential energy by the
symbol V, and conclude from (78.9) that
(78.10) Ep
The fields of force for which potential functions exist are called conservative.
There is a simple criterion for a field of force F; to be conservative. We
state it asa
‘THEOREM. A necessary and sufficient condition that a force field F,,
defined in a simply connected region, be conservative is that F; ;= F; ;.
The proof of this theorem follows immediately from the observation
that a necessary and sufficient condition for the expression F; dx‘ to be an
exact differential of a single-valued function V is that
(78.11) a
On gx
since these derivatives are assumed to be continuous functions.2. But
Pye {lF
Ox? ij
k\ . ies
and, since biis symmetric in 7 and 7, we conclude that the condition
78.11 is completely equivalent to the one stated in the theorem.
As a corollary we observe that a parallel force field (Sec. 48) is necessarily
conservative, since the condition for a vector field F; to be parallel is
F; ; => 0.
(79.1) T= m
=
railed
eat
Bit hs
2
If we differentiate (79.1) with respect to z we obtain
since #? = v'.
OT/dx' = mg;,;%. The derivative of this expression with respect to f is
) — meet + got)
4 O8s5 oe oi
sige
d (oT) _ <j
d ( =) oT OV
79.3 = || a
dt \oz* Ox Ox'
( )
or
the form
we can write equation 79.4 in
Noman
CayLue(LS) et
eae All aa!
214 ANALYTICAL MECHANICS [CHAP. 4
F = F'e,,
where the e,’s are unit vectors codirectional with the base vectors a,.
(See Sec. 45.) Since F = F’a, and a,-a; = g;;, the physical components
F are related to the tensor components F’ by the formula
Fi=vV¢g,,F', (no sum).
Problems
d
(Bi 7 [(e" sin aye].
Deduce these expressions from formula 78.3 and also from Lagrangean equations
We
mn
Hint: F; = ma, and T = *(5) Bicep
mids mgis .,.,x),
2. Use Lagrangean equations to show that, if a particle is not subjected to the
action of forces, then its trajectory is given by y* = a‘t + b*, where the a* and
b? are constants and the y* are orthogonal cartesian coordinates.
3. Find, with the aid of Lagrangean equations, the trajectory of a particle
moving in a uniform gravitational field. Hint: T = }my'y' and V = mgy’,
where 7? is normal to the plane of the earth.
4. Deduce from Newtonian equations the equation of energy, T+ V =A,
where h is a constant. Hint: Show that d7/dt = ma,v’ = —dV/dt.
5. Prove that, if a particle moves so that its velocity is constant in magnitude,
then its acceleration vector is either orthogonal to the velocity vector, or it is
zero. Hint: Compute the intrinsic derivatives of v? = g;,v'v’.
: d oT oT
6. We have shown in Sec. 79 that —— — — is a covariant vector F;
dt ox* Ox?
whenever T(x, @) is an invariant defined by (79.1). Prove more generally that if
Seger : . : d aw ow
W(a, ©) is an invariant, then both @W/ dé? and — —— — —— are covariant
dt oa Ox?
vectors. Hint: Let a = a*(q',q?,q°) be an admissible transformation of
coordinates. Compute #', show that 03*/dg? = axt/ dq’, and observe that
the invariance of W(x, #) requires that W(x, ¢) = W[x(q), #(q)] = W(q, 9).
Sec. 80] APPLICATIONS OF LAGRANGEAN EQUATIONS 215
80.1 eee an
d aT| OT
( ) aS Ox’
(80.3) va =H H= ol,
which states that the acceleration vector a lies in the osculating plane of the
curve. Moreover, the component of a in the tangential direction is equal
to the time rate of change of speed v, whereas the component in the
direction of the principal normal is v?/R, where R = 1/x is the radius of
curvature of C.
The force F = ma acting on a particle of mass m moving along C is
determined by
(80.7) F‘'=m a + mxvu'.
It should be observed that F" is the resultant of all external forces that act
on the particle and thus F includes the reaction R of the curve on the
particle. Since F lies in the osculating plane of the curve, the component
of all external forces normal to this plane is zero. This condition enables
us to compute reaction R in the general case. In mechanics the curve C is
said to be smooth if the reaction R is normal* to C, that is, if R‘A; = 0:
If R = 0 the curve C is called the natural trajectory of the particle.
As an illustration, let a bead of mass m slide under gravity along a
smooth curve C lying in the vertical Y1 Y-plane (Fig. 34).
The force F acting on m is
F = mg +R,
where R is the pressure exerted by the curve on the particle and mg is the
gravitational force. Since the curve is smooth, R is normal to C. If « is
the angle between the direction of R and the positive Y?-axis, the com-
ponents of F in the directions of the tangent A and the principal normal
ware
Fy = —megsina, Fi, = —mgcos«+ R.
* This is equivalent to saying that the frictional force is zero. The term “smooth”
employed in mechanics is different from the term “smooth” used in geometry, where a
“smooth curve” is one with a continuously turning tangent.
Sec. 80] APPLICATIONS OF LAGRANGEAN EQUATIONS 217
>y!
Fig. 34
(b) mo = —me
where
s+s=zg,
4a A
the general solution of which is
Fig. 35
Sec. 80] APPLICATIONS OF LAGRANGEAN EQUATIONS 219
Problems
1. Deduce the differential equations for a simple pendulum of length 1, and
show that for small oscillations the period is 27/ V g/l.
2. Derive the equations of motion for a particle moving under gravity on a
smooth helix:
=) @ COS.0) y* = asin 6, OS 3 Vide).
Note that since the helix is smooth, the reaction R is normal to the helix and
hence the component of the resultant force F in the tangential direction is equal
to the component of the gravitational force mg in that direction. The latter
component can be computed from gravitational potential V = mgy®. Further-
more, by (80.9) the energy equation in this case yields dm(v? — v9”) =
mg(y® —"").
3. If a particle of mass m moves on a smooth parabola with its axis vertical
and concavity downwards, show that the reaction R = xm(v2 — v’), where v’
is the velocity of m for which this parabola is the natural trajectory and v is the
velocity in constrained motion.
4. Motion of a Particle on a Surface. Let the equations of a regular
surface S be given in a parametric form as
us on au F
p= = = 2 a a’, (a = 1,2),
din) Oulasas,
or
(80.11) p= 20%,
where v% = ui".
yields
The acceleration a’ = 6v‘/6¢; hence equation 80.11
pene ot a OD a uae
mor ot
e the base vectors a
4 See equation 64.5. The reader should take care not to confus
nents a* used in this section.
used in Chapter 3 with the acceleration compo
220 ANALYTICAL MECHANICS [CHAP. 4
or
(80.12) a’ = x,'a* + ai gv7v",
where a® = 6v*/dt.
If we make use of the Gauss formula
[67.7] r= bane
equation 80.12 reads
a’ = 2,'a* +. v°%,)n'.
Since F* = ma’, we have
since the surface vectors x,’ are orthogonal to n; and n'n; = 1. The com-
ponents of F in the plane tangent to S, on the other hand, are given by
a, /F; = ma,,
and set F,, = x,’F;, we obtain a pair of Newtonian equations
(80.16) F, = ma,,
relating the surface force vector F,, to the surface acceleration vector a,
Sec. 80] APPLICATIONS OF LAGRANGEAN EQUATIONS 221
m m sakes
T= —a,,0°v" = — ORO
2 2
We obtain, as in Sec. 79,
dt bs
If we recall that
[71.6] OER
és
the tangent plane, and x,
where 77% is the unit normal to the trajectory in
is the geodesic curvature, we can write
fg ==, dv Ag + 2
vn a
We a")
so that
2 ds
that
If follows from this result (cf. equation 80.7)
Fe = ae + 2Tx,7",
ds
hes identically, then dT/ds = 0
where T = mv?/2. If the vector F% vanis
first of these equations states that
and x, = 0 along the trajectory. The
ctory is a geodesic by the theorem
= constant, and, ifv # 0, then the traje
of Sec. 63.
222 ANALYTICAL MECHANICS [CHAP. 4
y3
Fig. 36
(80.19) o=
ee EO wae ns gr
i+; Sasiyerin GLE wera
(80.20) :
d 2h
77
—(r°0) ) = 0.
The elimination of 6 from the first of equations 80.20 with the aid of
(80.21) gives
4 a
which has a unique solution when the initial position r = 1 and the
initial velocity 7 = vy of the particle are specified.
If our particle is constrained to move ona horizontal circle r = constant,
6? =
(80.22) requires that h? = gr*/2a, and equation 80.21 then shows that
g/2a, so that the angular velocity 6 is independent of the radius of the
circle. When the path of the particle is the meridional line 6 = constant,
we get from 80.20 the equation
n R
The integration of equation 80.22 and the calculation of the reactio
to the parabol oid is tedious. To com-
required to constrain the motion
in which F =
pute the magnitude of reaction, we need equation 80.15,
mg +R.
illustration by the
If we replace the surface of the paraboloid in this
the proble m of a spheri cal pendulum.
surface of the sphere, we have
for the spheri cal pendu lum can be
The solution of equations of motion
functi ons. When the surface is a cylinder
obtained with the aid of elliptic
easy.°
r =a, the integration of equations of motion is
in E. T. Whittaker’s Analytical
5 Interested readers are referred to pp. 99-109
Press (1917), where the motion of the particle on a surface is
Mechanics, Cambridge
analyzed in a different way.
224 ANALYTICAL MECHANICS [CHAP. 4
Problem
be ees
i} — (v’)* sin u cos ul = —5,
ma
ne A “4° A F.2
i sin? 2 + 2i4u? sin 2 cos ul = —> .
ma?
Solve these equations for the case when F* = 0, and show that the trajectory is
an arc of a great circle, and the speed v = const.
Hint: The first integral of the second equation is u? sin® u} = constant. Use
this result in the first equation and observe that v® = a®{(w)? + (@*)? sin® uv’).
C: x = x(t), iS tsk,
(81.1) 6a = oe -e = Ete,
de e=0
i = (=) :
de 0
where
(=) 0) ee ee a OF i
de /o de sp MOD at|
Thus
(81.2) OF = we ba,
Ox"
(81.3) Ea ee
dt dt
the variation.
so that the variation of the derivative is the derivative of
, 2%", t) of 2n + 1
Clearly, if we have a function F(a!,..., 2", #,...
C2, we can write
variables x‘, #¢ = dx‘/dt, and t, which is of class
OF OF
F OF = — 6x" — dz.
oy Ox" + 5g i
226 ANALYTICAL MECHANICS [CHAP. 4
where the functional arguments 2'(t), t, < t < t,, belonged to the h-
neighborhood of an extremal of J. That is, we considered the behavior of
the integralJ along the varied paths Z(t, ©) = 2° + e&‘(t). Making use of
equation 81.4 of this section and referring to formula 57.6, we see that
formula 57.6 can be written
te / \
OT Ole.
(82.1) OT = — 6a’ + — 62’,
Ox" Ox"
where «' = x'(t) are the coordinates of the particle along the trajectory and
a’ + 6x? are the coordinates along a varied path beginning at P, at time t,
and ending at Py, at time tg.
It will be shown next that this principle is equivalent to Lagrangean
is
equations of motion 79.2, and hence to Newtonian laws. The proof
simple. Substituting (82.1) in (82.2) yields
fe Ox"
parts,
Integrating the first term under the integral sign of (82.3) by
t
Aaiap
e
}t TE eeaa!
te
4, “fs
dtp at Oe
and &(t,) vanishing,
and, since x(t.) = dx'(t,) = 0 by virtue of & (ty)
equation 82.3 becomes
-420)da‘ dt=0.
te
(82.4) i}fat
Ox’ = at 0a"
ty
“() -4 =
dt\axz!) axt
follow at once from the formulation 82.7.
(83.1) ola
dt
es
where v’ is the velocity and a, is the acceleration of the particle.
For a conservative field of force, ma; = F; = —0V/dx’, and we can
write (83.1) as
Tai OV dx’
dt — Ox dt’
or
(eon i — 7h
Py
(84.1) A= i my + ds,
Py
ibe dx dx?
t(P) iodibads
SEC. 84] PRINCIPLE OF LEAST ACTION 231
and, since
ein dx dx!
~ 959 Gr at’
we have
t( Ps»)
(84.2) A -| 27 at.
(Py)
This integral has a physical meaning only when evaluated over the tra-
jectory C, but its value can be computed along any varied path joining
the points P, and P,. Let us consider a particular set of admissible paths
C’ along which the function T + V, for each value of parameter ¢, has the
same constant value h. The functional A so determined is called the action
integral, and concerning it we can formulate
THE PRINCIPLE OF LEAST ACTION.® Of all curves C’ passing through P,
and P, in the neighborhood of the trajectory C, which are traversed at a rate
such that, for each C’, for every value of t, T + V =h, that one for which
the action integral A is stationary is the trajectory of the particle.
When stated in the form of the variational equation, this principle reads
5 i(Py) oT dt = 0,
t( P2)
(84.3)
with the auxiliary condition
(84.4) T+VvV-h=0), enC
It is important to recognize that in this instance we cannot determine the
extremals of the action integral by setting F in the Euler equations 57.7
equal to 27, because of the auxiliary condition (84.4). Since Tis a function
of the velocity v, and V is a function of position alone, the times t(P,) —
t(P,) required to traverse the varied paths C’ will differ in general. Thus
the upper limit ¢(P,) in the integral 84.4 is not fixed. In this case we have
the problem in the calculus of variations with variable end points and with
one auxiliary condition 84.4. The procedure employed in solving this
problem makes use of Lagrange’s method of multipliers for a problem
with nonholonomic constraints, which we briefly indicate. (Compare
Sec. 57.)
We construct a function F = 27 + Ad, where ¢ = T+ V —A, and
determine the solution of the system of four equations
oF_ a (28jicwo
dx’ = dt \0z"
eee OIENE
T+ V—h=0.
6 Strictly speaking this principle should be called the principle of stationary action.
PE ANALYTICAL MECHANICS [CHAP. 4
(84.5) a (i = 1, 2, 3).
Gi\0u /e Ode Ox’
These are precisely the Lagrangean equations of motion.
A different and somewhat more illuminating mode of attack on this
problem is to reduce it to a consideration of the variational problem wiih
fixed end points by a change of variable. Since the kinetic energy
84.6
(84.6) pele
ae
= lanl ds.
2(h— V)
Consequently the action integral 84.2 can be written®
and, since the limits of integration in (84.8) are fixed, we see that the
determination of the trajectory is equivalent to finding the geodesics in a
three-dimensional Riemannian manifold with the arc element
(84.9) dS? = 2m(h — V)g,, dx! dx’.
* See Sec. 88 below and O. Bolza, Vorlesungen iiber Variationsrechnung, p. 586.
* The form 84.7 of the action integral was used by Jacobi. See a discussion of this
integral and its generalizations in C. Carathéodory’s Variationsrechnung, pp. 255, 290.
SEc. 85] SYSTEMS OF PARTICLES 233
If we form Euler’s equations
Fa —_— a pert — 0,
We have already remarked (in Sec. 77) that the passage from mechanics
of a single particle to mechanics of material bodies can be accomplished by
introducing certain assumptions regarding the nature of constraining
forces operating on particles making up the body. In some dynamical
problems the change of shape of the body is so slight that one is justified in
supposing that the particles remain at fixed distances from one another.
This assumption leads to the dynamics of rigid bodies. If a body suffers
nonnegligible deformations we can postulate, with varying degrees of
realism, the nature of constraining forces and thus arrive at the dynamics
of elastic bodies, ideal fluids, viscoplastic media, and so on. The assump-
tions concerning the nature of constitutive forces permit us to characterize
the positions of a large number of material particles in terms of relatively
few descriptive parameters. Thus a thin rigid rod of length /, moving in
space, requires only five parameters for the determination of its position.
These can be taken as space coordinates of its center of mass and two
direction ratios of one of the ends relative to the center of mass. The
choice of descriptive parameters is not unique, and they clearly need not
have the dimensions of length. A bead sliding on a curved wire requires
only one parameter for the description of its location, say the distance from
some fixed point on the wire; a particle moving on the surface is located
unambiguously by a pair of Gaussian coordinates. Whatever is the nature
234 ANALYTICAL MECHANICS [CHAP. 4
(85.2) tn ee A(een):
where we introduced the time parameter ¢ which may enter in the problem
explicitly if one deals with moving constraints.® If t does not enter ex-
plicitly in equations 85.2, the dynamical system is called a natural system.
We will suppose that the functions 2/,) = x!,(q, t) are of class C? in
the region of definition of the variables qg‘ and ¢ and that the Jacobian
matrix (02'/dq’) is of rank n [cf. (75.5)].
The velocities of the particles are given by differentiating equations
85.2 with respect to time. Thus
and assume, as in Sec. 85, that the functions 2’(q, t) are of class C*. The
velocity «” of any point of the body is given by
oe0q’ da!
dt
BeOt
etches asa G=1, ae
0q’- Ot
where the g’ are generalized velocities.
Let the system in question be natural, holonomic, with n degrees of
freedom, so that the relations
where mis the mass of the particle located at the point 2” and the summation
(or integration) is carried over the entire region occupied by the body.
The g,. in (86.4) are the components of the metric tensor associated with
the coordinate system X covering Es.
Sec. 86] EQUATIONS IN GENERALIZED COORDINATES 237
If we insert in (86.4) the values of #’ from (86.2), we obtain"
GLeon wa:
T= 4) mg,,— —4q‘q'
i Brea aaa
= ha;,g'q’,
where soe
(86.5) T = 4a,,q°q'
is an invariant, and the quantities a,, are symmetric, we conclude that the
a;; are components of a covariant tensor of rank two with respect to a class
of admissible transformations 86.3 of generalized coordinates. We note
that, since the kinetic energy T is a positive form in the velocities q’,
|a;;| > 0, and we can construct the reciprocal tensor a’.
If we carry out a computation, in every detail identical with that of
Sec. 79, by using the expression for the kinetic energy in the form 86.5,
we obtain the formula
Ox"
Q; ty > ae eat?
a 0q
(86.11) 4%) = fe ‘
dt \dq' 0q'
are known as Lagrangean equations in generalized coordinates. They yield
a system of n second-order ordinary differential equations for the general-
ized coordinates q’. The solutions of these equations in the form
Cc: g= dt)
represent the dynamical trajectory of the system.
If there exists a function V(q1,...,q”), such that
3 Ga =O
ar ag? *agit
AES Ol Pamolny
GEEwOLY, rd yeb.
(86.13) dt agi! tea)?
=.(ot '
dt\dq'
But, since L = T — V, and the potential energy V is not a function of the
qs
i (6T+ 0, 6q") dt = 0,
ta as AS
(86.14)
ty
and, if the force field Q, is conservative, the principle can be stated in the
form
te
) { Ldt=0.
th
ld, lav
seh,
m)r250 % 6 cos Jsant
— rd? sin
dt r 00’
d lor
—(r24 sin? 0) | = — ==
m| 5 0% si | rsin@ a’
whereas in cylindrical coordinates, with ds? = (dr)* + r°(d6)? + (dz)?, they are
oV
mi — r62) = ——,
or
Bas al
at” poe
In formula 87.2 the 6g’’s are arbitrary, and they are necessarily consistent
with constraints imposed on the system, since the coordinates q° are
independent.
If we insert expressions from (87.2) in (87.1), we get
Ox 59°
(87.3) W, = =F, —
q
= Q; 6q’,
where the last step makes use of the definition of the generalized force Q;.
It follows from this formula that one can calculate the generalized forces
Q;, acting on the system, by computing the work W, produced by dis-
placing the system through a virtual displacement dq’ ¥ 0, (j fixed), and
with 6g’ = 0,1 #7. Then Q, = W,/dg’. We shall resort to this method
of computing generalized forces in the illustrative examples of Sec. 89.
c,.g' 6t = 0,
But, 6g’ = (d/dt) 6g’, and the integration by parts of the first term in the
integrand of (88.5) gives (cf. 82.3)
Siang OF '
88.6 { (£5 = 0,
o “dt
( ) t, \Oq* dtdg’ 2: )04
e n generalized co-
The two sets of equations (88.8) and (88.10) involv
..-,9"). By adjoining
ordinates gq’ and m Lagrangean multipliers 2*(q',
244 ANALYTICAL MECHANICS [CHAP. 4
(88.14) a eo F0
sO that Cy = Qa, Cig = —1.
SEc. 88] NONHOLONOMIC SYSTEMS 245
Equations (88.11) then yield
ar--d
ee aT ee ia
a6. di a0
(88.15) ae .
an
Biing 4toOlesd
oT 1 Ya
sya
ee CENA Qs
Now, the work W done by the gravitational force alone, when the
center of mass moves through a distance x is W = amg sin ¢. Hence V =
—a«mg sin ¢, and
= ——
oV = 0, = —
oV
— = meg
;
sin
Q; 36 Q» an & $ @.
which, when compared with (88.11), show that the generalized reactions
R, are
The term k2mg sin ¢/(a? + k’), in the second of equations 88.17, represents
the frictional force F opposing the component mg sin # of the gravi-
F =
tational force along the plane. If the cylinder is solid, k? = a?/2, and
force F = uN, where uu is the
img sing. The magnitude of frictional
coefficient of friction and N is the pressure of the cylinder on the plane.
Since N = mg cos ¢, we conclude that « = F/N = } tan ¢.
the
As another illustration of the use of equations 88.11 we consider
brachistochrone problem in a resistin g medium. ’®
of Variations,” American
15 See G. A. Bliss, ‘“The Problem of Lagrange in the Calculus
Calculus of Variations (1962), pp.
Journal of Mathematics, 52 (1930) and L. A. Pars,
241-243.
246 ANALYTICAL MECHANICS [CHAP. 4
te x Ze 1\2
(88.20) " =i] dt = {Alin Pahited
a ee
ty 2, U 2 D
G=F+A¢
88.23
(88.23) dGy
ae mie aGy
- — G,=0,
SEc. 88] NONHOLONOMIC SYSTEMS 247
and, since G does not contain y, we conclude from the first of equations
88.23 that G,, = a or
ae Sa rn, lo’ = 0:
dx
or
(88.25) 5) tral
Vi+(y'?
We thus have the system of three equations 88.19, 88.24, 88.25 for the
determination of y(x), v(x), and A(x). We can rewrite them as
d dv is giv sey
ds ds
(88.26) H dy =Ag+a,
ds
te =H,
ds
Since dy/dv = dy/ds - ds/dv, we find with the aid of the first two equations
in (88.26) that
dy __w(Ag +a)
(88.29)
dv g(Ag +a) — RH’
the right-hand member of which is a known function of v. On performing
quadrature we then get
(88.30) y =fiv, a, b) + ¢,
where c is a constant. Equation 88.30 is one of the desired equations in
(88.28). To obtain x = 2(v), we note that dx/dv = dx/ds - ds/dv and, since
dx/ds = vie (dy/ds)? and both dy/ds and ds/dv are determined by the
first two equations in (88.26), we see that dx/ds is also a known function of
v. The reader will check that
dx bv
dv g(Ag+a)—RH’
so that
(88.31) x = f,(v, a, b) + d,
where d is a constant. The constants of integration in (88.30) and (88.31)
must be determined for the initial conditions. To make the problem
physically meaningful we must impose some restrictions on the relative
magnitudes of R(v) and the gravitational force g, as, for example, R < g
for all relevant values of v. .
Problems
Fig. 38
suppose that the pendulum is set in vibration in some plane which we take
as the Y1Y*-plane. (See Fig. 38.)
In order to form Lagrangean equations
(89.1) alee) ~ ag
we need the expression for the kinetic energy
(89.2) T = hmy'y.
However,
and hence the generalized force Q = —mg sing/I. Thus equation 89.1
yields
(89.4) gt sin 4=O)
250 ANALYTICAL MECHANICS [CHaP. 4
Fig. 39
and, since for small displacements sin 6 = 6, for small vibrations we have
ink diz 0:
where k? = g//. The solution of this equation is g = acos (kt + «). The
solution of (89.4) can be expressed in terms of elliptic integrals of the
first kind.
We turn next to a more interesting problem of a double pendulum.
Consider an arrangement of particles shown in Fig. 39, where we suppose
that the masses m, and m, are supported by inextensible light cords of
lengths /, and /,, respectively. The pendulum is assumed to vibrate in one
plane, and we take as our generalized coordinates the quantities 6 and ¢,
which give the angular deviations of the cords of lengths /, and /, from the
vertical.
The equations connecting the coordinates (y,, y,”) and (y,!, y”), of the
masses m, and m,, with generalized coordinates g' = @ and g* = ¢ are
yy = 1, sin q’,
y= 1, cosq’,
Yo = i; sin q + l, sin q’,
d
7 {(m, + my)(11)°g* + melylog? cos (q? — q*)}
and we suppose that, in the neighborhood of the point q' = 0, the a’”’s do
do not vary appreciably, so that they can be regarded as constants.
The Lagrangean equations 86.12 now yield the system ofn simultaneous
second-order ordinary differential equations with constant coefficients
4,4’ + big’ = 9.
Instead of integrating this coupled system directly we can simplify the
problem by introducing a new set of independent variables g”, the so-called
normal coordinates, which are related linearly to the coordinates g’ in such
a way that the quadratic forms 89.6 and 89.7 reduce simultaneously”* to a
sum of squares. We then have
T= (GEE = (gy ae oe ce (gay
All the coefficients of the q’’s in (89.8) are nonnegative since the quadratic
form 89.6 is necessarily nonnegative if the potential energy V has a
minimum at q’ = 0.
The Lagrangean equations now become
PM 2 OP, 2 242
Manin ance Atal
Thus the desired transformation 89.11 is
ph eo
=4/8 toy),
(89.13) a
2 -<, 5 a2 PN) 9),
254 ANALYTICAL MECHANICS [CHAP. 4
#4 +2 —/2)e¢=0, y+ FQ + ./2)y= 0.
Solving these, we get
(89.14) % =2.¢;.C0S (Ak €a)} Y = C3 COS (Ast + Cy),
where
qo — V/2 fi:
O O
q q
9 = ~2q, |
q, =— V2q,
2g +g + “Eq =0,
(89.15)
PichiG ch Sahen,
in which the variables g,' and g,? are coupled. We assume solutions of
(89.15) in the form
(89.16) gi=a,e*” q*?=a,e%
and determine / so that equations 89.15 are satisfied. On substituting
(89.16) in (89.15) we get two homogeneous equations
Ao wedis 2
On expanding this determinant, we find that
P= E24 V2),
which yields two values A, = (g//)(2 — V2), A? = (g/)(2-+ /2) corre-
sponding to the grave and rapid modes found previously. Thus the solution
(89.16) can be written
qi = cet" rs c,ettt,
q@ ae HD cet42t ae af cet,
as in (89.13).
Problems
39,
1. Find the normal modes of vibration for the double pendulum in Fig.
assuming that /, = /2, but my, # mg.
surface
2. A particle of mass m oscillates about the lowest point of a smooth
orthogon al cartesian and
z = 4(ax® + 2hey + by”), where the coordinates are
256 ANALYTICAL MECHANICS [CHAP. 4
the z-axis is directed vertically up. We suppose that the vertical component of
the velocity is small, so that T = 4m(a? + y). The potential V = mgz =
(mg/2)(ax® + 2hey + by”). Obtain equations of motion, determine their solu-
tions in the form x = a,e*, y = a,e™, and conclude that if V = min at x = 0,
y =0,thena >0,b >0, ab —h? > 0.
3. Let the particle in the problem at the end of Sec. 80 be acted on by the
force of gravity, so that F; = mga sin u’, F, = 0. (Note that the work OW done
in a small displacement dy? is 6W = —mg dy? = mga sin u’ du’.) Show that the
motion, when the particle passes through the highest and lowest points on the
sphere, is along an arc of a great circle. A complete discussion of this problem
is involved. See P. Appell, Mécanique rationelle, 1, Chapter 13, especially Sec.
277. See also a discussion of the spherical pendulum in J. L. Synge and B. A.
Griffith, Principles of Mechanics.
4. Let the particle in the preceding problem execute small oscillations about
the lower pole of the sphere. Consider projection of this motion on the plane
tangent to the pole and discuss the motion.
Hint: Set u’ = 7 — (r/a), and deduce equations
- aes
Fr
Pose ide eles a?
rii + 2ru = 0.
ALi: = Li
dt
and, if we recall the definition 90.3 and formula 90.5, we obtain a set of n
first-order equations,
dpi
(90.7)
90.7 —if) = aC
—H,i, Mb, )
ss on Wk
(90.8) H=q——L=g——T+V.
q' do Odd iy
0g’
and hence (90.8) reduces to
H=T-4+ VS.
Thus # is the total energy of the system.
The variables
1G. :
(90.9) dD; = ae = aif
0g’
are called the generalized momenta, and we note that the square of the
magnitude of the vector p, is
(90.10) Pp? = app; = A A;0;9°9'
= AGG’ = P28be
,
¥ nls + (r6)*] = $4;;9'9’
where
ue Ae
90.11) dritips
—=, et* —=-4,
dl p, dp, _ £2
1 Ps" _ yp), dp tet)
ae
( ) dt m dt mr dt mr ”) dt 4
The last of these equations, combined with the second, yields
d 24
a r’@) = 0,
—(mr*é)
SEC. 91] NEWTONIAN LAW OF
GRAVITATION 259
which is a statement of Kepler’s second law of planetary motion. It is
not difficult to show by using the remaining equations in (90.11) that if
V = —m/r, the orbit is a conic section (cf. Sec. 97).
”
Problems
1. If a particle of mass m is constrained to move on a smooth surface, show
that the system of Hamilton’s equations is
du* aH dp* eH
PID ape DogeTas Bee eed)
i ee
and g as independent, and show that
ae
tmOp, fOP
is
Ni Pe ct eh
oq", q
0g’) dii=D,
where m, and m, are the masses of the particles and rj, is the vector from
P, to P,. The constant of proportionality y depends on the choice of units;
in the cgs system its value is found to be 6.664 x 10-*, and its physical
dimensions are M-1! L? T-2. In our work we shall make y = 1, by a
suitable choice of units of measure, so that
(OIE) F=
We observe first that the law of gravitation 91.1 refers to two particles,
and, since in dynamics one usually deals with continuous distributions of
matter, it is necessary to generalize it. Thus one can subdivide the bodies
into small parts, replace each part by an equivalent material particle, add
the forces corresponding to discrete particles, and pass to the limit as the
number of subdivisions is increased indefinitely. This procedure for two
bodies 7, and 7, leads to the formula
19 We recall that the moment of force F relative to the origin, acting at a point
determined by the position vector r, is L=rxF or, in terms of components
L; = e:;y'F*.
SEC. 91] NEWTONIAN LAW OF GRAVITATION 261
y}
Fig. 41
(91.3) ee)
pa] | PrpaYo! a— t') 4, ie
(91-4) Lm | ikpipseiayse Ya
= = Ye dt, dtp.
We prove next that, as 7, and 7, are allowed to shrink toward P, and Ps,
respectively (or, even if 7, alone is allowed to shrink to zero) the resultant
moment L, tends to zero and equation 91.3 specializes to the law in the
form 91.1;
We choose the origin O of the coordinate system at P,, and let 7, shrink
toward O and 7, toward P.(y2!, Yo”, ¥2*). Since p; and p2 in equations 91.3
and 91.4 are nonnegative functions, the first mean value theorem for
integrals is applicable and we obtain
F= oa
wine
= ie
pip2 7; dt2,
and
eye eye
pe [ein | [ | roe tee
r U6| T2
This is precisely the law of gravitation 91.1 for two particles located at
(0, 0, 0) and (y2", Yo", Yo").
It follows from the foregoing that a material body interacting with a
point mass produces no resultant moment L. Moreover, direct calcu-
lations show that this is also true when the point mass is replaced by a
sphere 7 whose density p is a continuous function of the radius alone. The
resultant force F, exerted by the body on the sphere, turns out to be the
same as that produced by the body acting on a point mass m =e dr,
located at the center of the sphere.?° :
Consider next a body 7 with piecewise continuous density p and let
Py /y2, y°) be a fixed point either within or outside 7. The gravitational
potential V(P) at the point P due to 7 is defined by the integral
where r = J (y! — &)? + (y? — &)? + (y® — &)? is the distance between
Py’, y*, y®) and the variable point (1, €, €°) associated with the volume
element d7(&) of r. The integral 91.5, as we shall presently see, defines a
differentiable function V(y', y?, y°) for all locations of P.
If P is outside the body, the integral (91.5) is proper and we can compute
as many derivatives of V as desired by differentiating (91.5) under the
integral sign with respect to the parameters y‘. In particular,
(91.6) — = -F,,
(91.9) VV =0,
The integral with the subscript 7 is evaluated over the volume 7, whereas
the integral in the right-hand member of (92.1) measures the flux of the
vector quantity F over the surface &.
We recall from elementary vector analysis that, in orthogonal cartesian
coordinates, the divergence of F is given by the formula
OF = Or OF
222 iV Ee eae
sos Oy: andy bay?
If the components of F relative to an arbitrary curvilinear coordinate
system X are denoted by F’, then the covariant derivative of F° is
I i \ px
yaa —— OF” _ i \ps >
Ox? kj J
and we observe that the invariant F’. in cartesian coordinates reduces to
the right-hand member of (92.2), and hence it represents the divergence of
the vector field F. In addition,
From this theorem two other theorems (usually attributed to Green) can
be derived easily.
Let u(x, x, x°) and v(a!, 2, 2°) be two scalar functions of class C2 in +
and of class C1 in the closed region = + 7. We denote the gradients of u
and v by u; and »v,, respectively, so that
where
n° Ve= yn= ae
on
Interchanging u and v in equation 92.5 and subtracting the resulting
formula from equation 92.5 yields a symmetrical form of Green’s theorem
Theorems stated in equations 92.3, 92.4, 92.5, and 92.6 are, perhaps,
the ones most frequently used in mathematical physics.
The Laplacian of v,
(92.7) V0 = £70;55
when written out explicitly in terms of the Christoffel symbols associated
weet)
with the curvilinear coordinates x‘ covering Es, is
a > ae
and the divergence of the vector F" is
eI IGES
aoe
k\ dv
uilox
; OF’ jal
(92.9) Fi,=—+ ale
Ox j
266 ANALYTICAL MECHANICS [CHAP. 4
Formulas 92.8 and 92.9 can be written in different forms, which fre-
quently are more convenient in computations. Equation 31.10 yields
i| ) ra
31.10 || Ete
Sore lid = date
and hence the divergence F', in (92.9), can be written as
I
O. as 0
——a ‘|F,:
Ps Ox? a (2 = Vg
or ¢
(92.10) Fi,=
plet(yek)
eaaceee
If we set in this formula F’ = g”(dv/0x’), we get
1 a(/g g? dv/dx’)
(92.11) Vv = gv J ie ae Cae
Vg Ox
We turn next to a consideration of Stokes’s theorem which permits us
to express certain surface integrals in terms of line integrals.
Let a portion of regular surface & be bounded by a closed regular curve
C, and let F be any vector function of class C1 defined on & and on C.
The theorem of Stokes states that
ey i) e3
7) 0 0
(92.13) curl F = By) By? By ;
TEs eRe
the e; being the unit base vectors in a cartesian frame. The determinant
in 92.13 can be written as a symbolic vector product V x F.
We consider the covariant derivative F;; of the vector F; and form a
contravariant vector
(92.14) Gi = —c*F,
It is readily checked that in cartesian coordinates equation 92.14 reduces
to 92.13, and we define the vector G to be the curl of F.
Sec. 92] INTEGRAL TRANSFORMATION THEOREMS 267
Since n+ curl F = n,G' = —e’’*F, ,n,, and the components of the unit
tangent vector A are dx'/ds, we may rewrite equation 92.12 as
(92.15) -| e*F
,.n,do =| foe ds. 1
5 Cans
1. Prove that
}v,n' do = |v dr,
Ji T
y Lied dv d (1 av
OO = TlaN ae) 7 BON26) |
where F, and F, are the physical components of the vector F, that is,
F = Fir, + F.6,,
av
it. |
tere
1a eV
ar 2a Toe”
where F = F,r, + F,0, + F,2, and ry, 9), 2 are unit vectors, so that F,, F,,
and F, are the physical components of F.
(c) In spherical coordinates with ds® = (dr)? + (0)? + 9° sin? 6 (dd),
1 a(r2F,) 1 a(sin 6F,) 1 aF,
a ~ 2 ar r sin 0 ld) rsin@ a’
dv dv
: 1 7)(gee= 1 a
a(sin esS| i 1 a
ets V £9 ae V e353Ee
where the a; are the unit base vectors and F = Fla, + Fa, + Fag.
4. Show that the contravariant components of the curl of a vector F are:
i (oF, oF \ 1 perl ere =i
I OF, OF;:
lat ~ a) vp lon sat) Vg \ot axe)
5. Prove that under suitable restrictions on continuity the curl of a gradient
vector vanishes identically.
6. In orthogonal curvilinear coordinates,
- 1 1
Seu? §22 = 99> &§33 = 733°
=p" =0, i #j, and &u
Siti
fe)
If we set ds* = e,*(dz!)? + e,%(dx?)® + e,*(dx3)*, so that 23, = e17, So0 = es",
§33 = eas then
k
(a) [j,k] = 9, - = 1) i, j, k distinct,
i
ave a ey de;
—
[ii, i]= e;aa fi | ee
\i/ _ oOloge;
[ij, 1] = —[ii,j] = €% yi?
Fig. 42
We thus have
continuous
The result embodied in formula 93.2 can be easily generalized to
er such distrib utions nowher e meet the
distributions of matter whenev
is a standar d one. The contrib ution to the
surface ©. The procedure
element p dr, contain ed within 7, is
flux integral from the mass
[p-n da =| 7"r
ao,
> pay
y within & is
and the contribution from all masses contained entirel
interior to x. Since
where |denotes the volume integral over all bodies
r never vanishes, so that the
all masses are assumed to be interior to x,
270 ANALYTICAL MECHANICS [CHAP. 4
integrand in (93.3) is continuous, and hence one can interchange the order
of integration to obtain
cos 4 do
But the integral | = 47, since it represents the flux due to a
x
unit mass contained within &. Hence
(93.6) i
x (F -n);do = 47m,
Fig. 43
SEC, 94] GREEN’S THIRD IDENTITY 271
where m is the total mass within & and the subscript i refers to the flux
produced by the masses inside &. On the other hand, the net flux over 2’
produced by all masses outside & is
(93.7) Le -n), do = 0,
for the flux cone from any point outside & cuts &’ twice.
Now, if we let &’ and &” approach &, the right-hand members in (93.6)
and (93.7) do not change, whereas the left-hand member of (93.6) becomes
the flux integral over © produced by the masses within & and the left-hand
member of (93.7) represents the flux over & due to all masses exterior to x.
Thus the total flux produced by a distribution of masses within & is
However, formula 91.6 states that F = —VV, and thus (93.10) is equiv-
alent to
(93.11) V2V = —4rp.
Thus at all points interior to the body 7, the gravitational potential
satisfies the Poisson equation. We note in conclusion that formula
dr
[91.5] V(P) = [=
gives a solution of equation 93.11 at all points in rT.
10V 2)
lel (2Nk
+ (on On dg,
where n is the unit exterior normal to the surface & + o. Since, however,
on o the normal n is directed toward P,
I | | i]
nN=)
eg
ee lie
Sade a7aS
hes S
tel
ee
I | ey.) +. = i3
Sa?
ae= Bas
seed Seo
Fig. 44
Sec. 94] GREEN’S THIRD IDENTITY aE:
where V is the mean value of V over the sphere o, and w denotes the solid
angle.
On letting 6 — 0, the right-hand member of (94.2) yields —47V(P), and
it follows from (94.1) that 5
(94.8) a = 0.
x On
la 92.6 on setting vu= |
The condition 94.8 follows directly from the formu
ied by every harmonic function.
and v = V. It is anecessary condition satisf
274 ANALYTICAL MECHANICS [CHAP. 4
Show that formula 94.7 is valid in an infinite region 7 exterior to a closed surface
= whenever Vis regular at infinity. Hint: Apply formula 94.7 to a finite region
bounded by = and by a sphere S of radius R so large that S encloses &.
We have just shown that the solution of the interior problem of Dirichlet
in Laplace’s equation
V7u = 0 in 7,
(95.1)
u=f(P) on,
when it exists, is necessarily unique. Also the solution of the Neumann
interior problem
V*v =0 in 7,
(95.3) fedo = 0.
w(P’) = — : pad
so that W(P’), and hence G(P, P’), is uniquely determined by properties
(a) and (5). We call G(P, P’) Green’s function for the region r.
We show next how Green’s function can be used to construct an
explicit integral formula solving the Dirichlet problem in Poisson’s
equation
(95.4) V?V = —4rp iy;
V =f (P) on &.
Fig. 45
Sec. 95] FUNCTIONS OF GREEN AND NEUMANN 277
The integral formula will include the solution of the boundary value
problem 95.1 as a special case.
Green’s symmetrical formula,
+| 7 eal)
Baus) do.
a or
dV/dn =
In writing (95.6) we observed that, since n is an exterior normal,
.
—aV/ar and dG/@n = —AG/er on o.
= r? dw, where dw is
Since @V/@r and w are continuous on 6 and do
second integral on the
an element of solid angle, it is obvious that the
right in (95.6) tends to zero as 6—0. Similarly,
[ vo ae =0 as 60,
a or .
whereas
[ oO ae Ag V(P) as 6-0.
a if
get
Accordingly, on letting 6 — 0 in (95.6), we
aG do,
Vi"
(95.7) 4nV(P) = I4nGp dr —|
278 ANALYTICAL MECHANICS [CHAP. 4
To apply this formula we must first obtain Green’s function G for the
region 7, that is, we must solve the special Dirichlet problem
V’w = 0 in 7,
w= — on».
V(P) = a { gN do.
4n Jz
Fig. 46
When a positive unit charge is placed at P(z, y, z) (Fig. 46) and a negative
unit charge at the mirror image Q(z, y, —2), the electrostatic potential G
produced by these charges at P’(é, 7, ¢) is
(96.1) econpep
te
where r= PP’ =V(E—2P? +(7-y? + (6 — 2
and
r= OP =VE— 2 + (g— yh + (E+ 2.
Obviously, G = 0 on the planez = 0, and since w(P’) = —1/r’ is harmonic
for z > 0, equation 96.1 gives the desired Green’s function for the region
z > 0.
On the plane z = 0,
ie E oF Dat |
on OC Ico rat ral Ico
95.8,
and after performing simple calculations and substituting in formula
we obtain the solution of the Dirichle t problem 95.1 for the region z > 0
in the form of Poisson’s integral
= at 5 on S
r op
Fig. 47
SEc. 97] THE PROBLEM OF TWO BODIES 281
and formula 95.8 yields the solution of the Dirichlet problem 95.1 for the
sphere in the form of the Poisson integral
9
r4 “
pee) 20%
7 207 2 2 ! ! 1
where cos y = cos 6 cos 6’ + sin @ sin 6’ cos (¢’ — ¢), 9 being the co-
latitude and ¢ the longitude of P.
Green’s function for the region 2? + y? + 2 > R is obtained from
(96.3) by interchanging the roles of P and Q.
Problems
1. Show by using (96.4) that for every position ofP in the interior of the sphere
1 R? — p
(tej
2 Nhe
47R Ss r3
fixed
Hint: Take the z-axis along OP, so that r2 = R? — 2Rpcos@ + p*. For a
is fixed. Express do = R® sin 6 dé d in spherical coordinate s
position of P, p
and evaluate the integral.
S is
2. Show that the solution of the exterior problem 95.1 for the sphere
given by
—= |
VP) = 4nR f(P) eet tay
Js re ‘
where p = OP >R.
3. Deduce (96.5) from (96.4). Hint: Let y be the angle between OP and OP’
when P’ is on S.
(97.2) i ce PO ig tations)
mM, + Mg
Clearly, the wu’ lie on the line joining the points (x,") and (a,"), and our
choice of the generalized coordinates is then as follows:
g=¥, g@=y, g@=y, qg = ul, g@=uw, ge =u.
If we solve equations 97.1 and 97.2 for the 2,‘ and 2’, we obtain
a ; ye mra a
m m
(97.3) 1 2
pt yt ds FL ay y?
3
m, + Ms,
and these equations enable us to determine the positional coordinates 2‘
in terms of the generalized coordinates g’.
BXac
y?
Fig. 48
SEc. 97] THE PROBLEM OF TWO BODIES 283
This particular choice of generalized coordinates is made with a view
toward obtaining a simple expression for the potential energy V of our
system of particles. Indeed, since the magnitude of the force of attraction
F is given by F = mym,/r?, where r is the distance between the particles,
the potential energy V is
MM, _ mM,
r (2) ie! ao})? a (x, on 2) ee (x,° = Nal
1 oer ee ML) a
—(m,
5 1 + m,)u'u’
2 + =ea 12 ity’.
yy
ey be (i= d8208))
(97.4) m, + Mm,
i’ = 0, (i-==1132;,3).
of
Equations 97.4 are the differential equations characterizing the motion
the motion of the mass m, relative to m, is
our system. We note first that
toward it
the same as though the mass m, were fixed and m, attracted
l is [(m, + m,)/m,|V . This follows at once
with a force whose potentia
form
from the first three of equations 97.4 if we rewrite them in the
m, + m, 0V
(97.5) mi = — m, oy
the action of
Thus our problem is reduced to a study of motion under
97.4 states that the center of
central forces. The second set of equations
mass moves in a straight line with constant velocit y.
under the assumption
We shall carry out the integration of equations 97.4
than m, (the mass of the
that m, (the mass of the sun) is much larger
close to the mass m,
earth). If m, > m, the center of mass u’ will lie very
with those of the mass
and hence the coordinates u’ will nearly coincide
of equati ons 97.4 we conclude
m,. Thus 2,’ = u’, and from the second set
284 ANALYTICAL MECHANICS [CHAP. 4
y2
Fig. 49
mF
— mar? = — MyM»
r ’
d Py
or 28)
—(m,r*6)'= 0,
Or
(97.6) caer
F— rO° + ies 0,
r°é = h,
rat a0.
h Jo
Since df/dt = df/d6 - d6/dt, we have the relation d/dt = (h/r*)(d/d6), and,
making use of this in the first equation in (97.6), we get
r? d6 \r? dO err
or multiplying by 7°,
d (h dr h?
Feu ee = 0),
ie!) dé \r* d6 r :
d'u my
de" hh?”
whose solution is
v= a — ecos(§ — «)],
or
if
(97.8) r
~ 1—ecos(6 — a)
y?
m9
>
Fig. 50
We thus see that the orbit is a conic section (Fig. 50) whose eccentricity
is e, with the position of the apse line determined by «. The constant « is
known as the perihelion constant. We shall not go to the trouble of
determining these constants”* in terms of the initial position and velocity
of mass m,, since the main object of this section is to obtain formula 97.8
for the purpose of comparing it with the corresponding equation of the
orbit in the relativistic dynamics.
8 See P. Appell, Mecanique rationelle, vol. 1, Chapter 11, and J. L- Synge and
B. A. Griffith, Principles of Mechanics (1959),pp. 160-169.
5)
RELATIVISTIC MECHANICS
1. Physical laws and principles have the same form in all Galilean systems;
that is, reference systems that move relative to one another with uniform
velocities.
2. The speed of light in free space has the same constant value in all
inertial systems.
It follows from these equations that the origin of the system Y moves _
relative to the system Y with constant velocity. To see this, note that the
coordinates of the origin O of the system Y are (0, 0, 0), and hence the
trajectory of the origin O relative to Y is given by (99.1) as
7 = at,
C:
i = a,'t.
Hence dy'/di = «,'/«,4 = constant.
It can be shown in a similar way that the coordinate planes move with
constant velocity, so that the reference frames Y and Y are Galilean.
Let us suppose next that a spherical pulse of light is sent out from the
point P(y', y?, y?) of the system Y at the time t. According to Einstein’s
second postulate, light travels with constant speed c in all directions;
hence in dt seconds a photon starting from the point (y’) will be at the
point (y’ + dy’), and
(99.2) dy* dy =c? dt?.
Relative to an observer located in the Y-system, the light pulse originates
at the point (v1, 7’, 7°), and his equation for the spherical wave front,
di seconds later, is
(99.3) ay dy = Cat
Now if we substitute in (99.3) from (99.1) and compare the result
with (99.2), we find that a particular set of equations
7 = k(y* = vt),
y =y",
(99.4) yy
We note that the determinant of coefficients of this form has the value
—cg.
The foregoing formulas can be cast in a symmetric form by setting
f= 2*; then
(99.6) do® =a,,dx* dx’, (4,8 = 1,2, 3,4),
where
a;; = —fij> GJ = il, I, 3),
and require that the form 99.6 be invariant under the class of transforma-
tions 99.7, we can formulate the calculus of tensors as we did in Chapter 2.
Problems
1. Show, with the aid of equations 99.4, that events that are simultaneous
from the point of view of an observer in the Y-system are not in general simul-
taneous in the Y-system.
2. Discuss the slowing down of moving clocks.
3. Differentiate equations 99.4, and establish the relations between the com-
ponents of velocity w* of a moving point, as measured by an observer in the
Y-system, with the corresponding quantities w* measured in the Y-system.
dy} wi +0 dy* Ww
Ans. (Coo)
dt 1+ (6/c)w’ dt k(1 + (B/c)w)’
4. With the aid of the formulas given in Problem 3, show that, if # and v are
both less than c, then w/c < 1. Thus, if v = 0.9c, w = 0.9c, then w = 0.994c
instead of 1.8c given by the usual law of composition of velocities.
5. The expression arctanh w/c is sometimes called the rapidity. Show that
the usual law of composition of velocities is obeyed by the rapidities. Thus
Ww Ww v
arctanh — = arctanh — — arctanh-.
G Cc c
so that
day ee2 dx dx’
(100.2) (<<)
dt ee iba di
=e
(100.3) ec
c differentiation as
and define the operations of covariant and intrinsi
s us to define the Minkowski
was done in Chapters 2 and 3. This permit
acceleration vector f* by the formula
a 2a B da’
oe a
(100.5) se bo ais Z (a, B,y =1,...,4).
do? \By) do do’
so that do? = c? di? — dy’ dy’,
If our local reference frame X is cartesian
accel erati on relative to it are
the components /* of the Minkowski
h. (ar
patdo? ya = Sy
cdi\do
di\do/do
1 d’y"
ela
dx* dx 1 v®
dx* eS ER are
Sa
5 For,
dt Vey Ver —v?
do (c? dt® — gi; dx* dxi)t
yp? = yp? = 0, zt = ¢ a local system.
in
294 RELATIVISTIC MECHANICS [CHAP. 5
so that
(iain?) 3),
Cc
fi =0, since y =7.
We shall show next how Newton’s second law can be written in an
invariant form relative to all Galilean reference frames. Consider the
formula suggested by Newton’s second law,
Fe pa rg Oe OEY
(exes
a) dt
Fo
= — (mott") —
at : Line
1 | a
= ————_—(m, —
Jc? — v2 ot do,
ale
faa
Bp ae
ot Vet — v? dt
sepa $ (—Fe =)
c/1 — pr ot\/1 — g? dt]
where we made use of the relation 100.2, and set 6 = v/c. If we define
Tae
100.6 “fie treat ies:Sl, dx* E
( ) Pp c’ Ot dt
and since, in the local coordinate system Y,8B =Oandm= Mo;
ay”
100.7 Fe — Mo
( ) ec dé
= mof*
This has the form of Newton’s second law used in classical mechanics.
We see that the invariant m, is the mass of the particle P referred to a
local reference frame. It is called the rest (or proper) mass of the particle.
Sec. 101] EINSTEIN’S ENERGY EQUATION 295
Since equation 100.7 is a tensor equation, we can write the force equation
as
Fr = mf,
which is valid in all Galilean reference frames. ’
We shall rewrite (100.6) in the form
(100.8) gS (=),
where v* = dzx*/dt, and F* = c2/ 1 — f? F*, and shall take it as the
equation of motion of a particle in the restricted theory of relativity.
T_T x| mo dv =| m= 4(=)
vo v at dt
t i 72,0
= |meee a
dled
(101.1) T=|
Po
, we get®
and insert for the ¥, from equation 100.8
P i
zie Fax =| ee
P
Po Fed ig tf
{|4 ( 1 dv
jodx viSS i ta
— |———— ] v —_ em
=m
Joldt\/1—p dt dtvi—p?dat
derivative reduces to the
is cartesian, the intrinsic
6 Since the reference frame
ordinary derivative.
296 RELATIVISTIC MECHANICS [CHaP. 5
But
Phe MO) 2 v'Dto a ot zt
— ’ D = 3
P Cima. dt
hence vi(dx'/dt) = Bc?, and BB = (v'/c*)(dv'/dt). Substituting these ex-
pressions in the integral, we get
iyoa
A ree lia ="
mi l ca rt a el
gales
= myc!| (ema a
Thus
=n le)
Moc
+ constant.
e (1 i By
=(m — m))c’.
Thus
(101.2) m= Mp) a =
Cc
We see that the mass m depends on the kinetic energy. If this result is
assumed to hold in dissipative systems, then the decrease in mass m must
be accounted for by the loss of energy by radiation.’
We see from the foregoing that the principles of conservation of energy
and conservation of mass, which appeared to be quite distinct in the
classical theory, can be united into one law in the restricted theory.
We also see from equation 101.2 that, if a particle takes up an amount
of energy AT, then its inertial mass m is increased by an amount A7/c?.
"In vol. 41 (1935) of the Bulletin of the American Mathematical Society, Einstein
gave an elementary derivation of this mass-energy relation by basing his considerations
on the principles of conservation of energy and momentum. For a definitive treatment
see J. L. Synge, Relativity: The Special Theory (1956), Chapter VI.
SEc. 102] RESTRICTED ‘THEORY 288 |
in which myc? appears as the intrinsic energy and T as the kinetic energy.
reduces to
where G;’ is the Einstein tensor defined in Sec. 38. If we contract Gr.
we get the equation R — 34R = 0, so that R= 0. Accordingly,
(103.4) Ry = Rie = 0,
where R,, is the Ricci tensor. These equations include the flat manifold
of restricted theory and admit the case for which the components of the
curvature tensor do not vanish.
0 Fe aol Gl be negative.
>
where we write |g| since the determinant of the form 103.1 may
of the coefficients in
10 This terminology can be justified by examining the form
equation 84.9 in a related problem in Newtoni an mechani cs.
making use of equations of
11 This equation is suggested by a chain of reasoning
=0, where GY? = —pu'w’ with uv‘ = dx'/dt. A delightful
motion in the form G,;
s, Mathema tics of Relativity
account of this approach is contained in G. Y. Rainich’
(1950). See also Problem 2, Sec. 38.
it can be shown that there are four
12 These equations are not independent, and
on, The Mathematical Theory
relations connecting them. See, for example, A. S. Eddingt the calculations
, has no bearing on
of Relativity, 2d ed. (1924), p. 115. This fact, however
given below.
300 RELATIVISTIC MECHANICS [CHaAP. 5
It is obvious from the foregoing that the system of ten nonlinear partial
differential equations (see Sec. 38)
Ry=0
for the ten unknown functions g;; is extremely complicated.’* The general
solution of this system is not known, and one is obliged to seek particular
solutions, essentially by trial, and use Newtonian mechanics as a guide
in selecting sensible forms for the coefficients g;,. Once a set of g,,’s
satisfying equations 103.4 is found, we can form the equations of geodesics
103.3, and if the solution of equations 103.3 agrees to the first order of
small quantities with the corresponding situations in Newtonian theory,
all is well.
We shall illustrate this procedure in Sec. 104, where we will obtain the
Schwarzschild" solution of the gravitational equations 103.4.
Before we proceed to that topic we note that equations 103.3 can be
written in a neat form,
(103.5) xd? = 0,
where #' = dx'/ds. If we regard the vector dx'/ds = i’ as the tangent
vector, then equations 103.5, or A',d’ = 0, are precisely the equations for
the parallel displacement of the tangent vector 4’ along a geodesic. Our
problem has thus been reduced to the solution of a deceptively simple-
looking system
R,, = 0,
7 ea
zt’ = 0,
with which we will occupy ourselves in Secs. 104 and 105.
hh = eos ho = Be
= * Ati! cin?
—e'*r’ sin’ 8,
& = 211822833844 =
and the contravariant tensor g” is given by the matrix
Ca 0 0 0
fe
Aoner may 0
(g’) = 1
0 0 = 0
r® sin? 6
0 0 0 ea,
A ee oe
It is easy to verify that distinct, nonvanishing Christoffel’s symbols are
14 ae - \22 3 93). :
A = —rsin® Oe“, lsat= —sin 6 cos 6, ee a ae ie
ron ER) fn
" aatdat at ij) * Maj) lial lil at”
Sec. 104] SPHERICALLY SYMMETRIC STATIC FIELD 303
and obtain after tedious but simple calculations the following set of
differential equations:
so that
A = —yu + constant.
However, as r—> ©, A andy tend to zero; hence,
Ar) = —H(0).
Epuation 104.6 thus becomes
(104.9) e*(1 + ry’) = 1.
We set
ev=y,
and equation 104.9 becomes
ytry’ =1.
Integrating this first-order linear equation, we get
(104.10) et err oS
105,
where 2m is a constant of integration. We shall identify m, in Sec.
with the mass of the sun.
ns
It is easily checked that the solution just obtained satisfies all equatio
we get the
in our system. Inserting e-* =e" = y in equation 104.4,
desired quadratic form
(104.11) ds* = —y“(dr)e — (db)? — FP sin? (dg)? + y(d0)*
2m vanishes, y = 1,
where y = 1 — 2m/r. If the constant of integration
ted theory. For
and the resulting manifold is the flat manifold of restric
m 0, the manifold is curved.
304 RELATIVISTIC MECHANICS [CHAP. 5
The reader may feel uneasy about the Schwarzschild solution of
Einstein’s gravitational equations, since it was obtained on the basis of
several fortuitous guesses with one eye cocked on results of the classical
theory. He may feel that a different mode of attack might yield a different
solution. That this is not so was shown’? by G. D. Birkhoff, who demon-
strated that all spherically symmetric static solutions of the gravitational
equations R,; = 0, which yield a flat metric at infinity (that is, the one
characterized by equation 104.2), are equivalent to the Schwarzschild
solution. Thus the solution obtained previously is of interest because
it is the only static solution of our equations satisfying specified boundary
conditions at infinity.
ogds” [Nee
12) ds ds
de
21
IEE >
ds ds (33/ ds his. a
or
2 2
(105.1) a0, Behe aoe 6(42) =)
dss aTasias S
In a similar way we form equations for i = 1, 3,4. The results are
a 1 2 2 2 2
(105.2) an au( a) — yr (2)— yrsin® 0(22) + raat) = 0,
ds*> 2y dr\ds ds ds 2 dr \ds.
2
(105.3) as SMC Leta ery 8G las
ds" rds ds ds ds
(105.4) So a andar
2 -
ds’ y drds ds
*° G. D. Birkhoff, Relativity and Modern Physics, p. 253.
*® For an elegant treatment of planetary orbits by means of Lagrangean equations
see J. L. Synge, Relativity: The General Theory (1960), pp. 289-298.
Sec. 105] PLANETARY ORBITS 305
The last of these equations can be written
dt ldydt_
2
ds* yds ds
:
or
(105.5) a(
ay
“)
di
a
ds or
We will prove that the analytic solution of equation 105.1, satisfying
the initial condition d0/ds =0, when 6 = 7/2, is 0(s) = 7/2. Since
d0/ds = (d6/dt)(dt/ds), and dt/ds # 0, this is equivalent to showing that
the trajectory of the particle lies in the plane 0 = 7/2, provided that the
initial component d6/dt of the velocity, in the direction of increasing 6,
vanishes. We thus assume that the solution 6(s) can be represented by
the series
Since d6/ds =0, when 6 =7/2, equation 105.1 for = 7/2 gives
(d?6/ds*)) = 0.
To obtain (d?6/ds*), we differentiate equation 105.1, and insert in the
result the values 0= 7/2, d6/ds = 0, and d?6/ds? = 0. We find d°6/ds* = 0.
In this manner we can show that 6(s) in (105.6) is 0(s) = (8)o = 7/2.
The corresponding result in the Newtonian case is obvious since, under
the assumption of the central field of force, there can be no component
of force at right angles to the plane of motion. Thus, if the motion had
once started in the plane § = 7/2, it would continue in that plane. If
we insert the solution 6 = 7/2 of equation 105.1 in equation 105.3, we get
2
(105.7) d biieeadr dp ih 0;
ds are rasds
and integrating equations 105.5 and 105.7 we obtain
dr 1 du ee an eng a
a7 aes a ahag) kag
and equation 105.12 reduces to
(105.15 ) ah ok
Sec. 105] PLANETARY ORBITS 307
where we write ¢@ for the angular variable @ used in that section and
introduce the gravitational constant k = 6.7 x 10-° and
m, = 1.98 x 10* pr «
is the mass of the sun. Because of our choice of units for the velocity
of light, we note that far away from gravitating matter
du a km, - 3kmyu"
dd? h? Ce
where m, = 1.98 x 10*8 gr (mass ofthe sun), k = 6.7 x 10-* gr-* cm?/sec?,
r= 3 eO cnt secs
For the motion of Mercury the term km,/h? is of the order 10°”,
whereas 3km,u?/c? is of the order 10-71. These estimates justify us in
attempting to solve equation 105.13 by a method of successive approxi-
mations sketched in the following section.
d°u
(106.1) +u= ne + 3h*u’),
d¢”
deduced in Sec. 105, can be integrated in closed form with the aid of
elliptic functions, but the solution obtained in this way does not lend
itself to a convenient comparison with the corresponding result obtained
in Sec. 97 on the basis of the Newtonian theory.
We noted in Sec. 105 that the magnitude of the term 3/?u?, appearing
in the right-hand member of equation 106.1, is small compared with
unity, and this justifies us in attempting to obtain a solution of this equation
by the method of perturbations. Accordingly, we neglect the small term
3mu* and obtain for our first approximation wu, the Newtonian equation
au, m
ples gate
the solution of which is
m . 6m°
= ie + i ecos(¢ — w)
ame. 3m?
ieee creat aie Ae ewes
SP Ss + SF e cos ($ — 0).
d¢” h*
The solution of this linear equation is clearly made up of the solution
u, and the solution of
du
d¢”
3
u = ST ecos ($ — 0).
up in
The result of easy calculations gives us the second approximation
the form
(106.4) w= ml
2
+ ecos(¢ — w)+ we ed sin (d — 0) }
steps in
It will suffice for our purposes to terminate the sequence of
and to regard u,
the scheme of successive approximations at this stage
ntly high degree
as representing the solution of equation 106.1 to a sufficie
of accuracy. If we set
3m?
(106.5) bo = 72
(106.7) a a omrad.
Equation 106.6 represents a closed orbit, only approximately elliptical
in shape, because dw is a function of ¢. Since u = 1/r, we have
TiAl 2
> Eee es
1+ ecos(¢
— w — da)’
_ 64m? 67m
"alae
a es
In this expression m is the mass of the sun.
For Mercury the quantity « works out to be 4.90 x 10-* rad. This
angle is very small, but the observational data on the location of Mercury
during the last century are available, and since this planet has a period
of 88 days, it completes 415 revolutions per century. Thus the cumulative
advance of the perihelion in 100 years should amount to 415e = 2.04 x
10-4 rad = 42" of arc. For planets other than Mercury the corresponding
advance is too small for accurate experimental determination. Thus for
Venus it is only 9”, for Earth 4”, and for Mars 1”.
The actual path of Mercury about the sun is not an ellipse, of course,
because of the perturbing effects of other planets. We are not in reality
dealing with a two-body problem. However, perturbations due to other
planets can be taken into account and the deviations from an elliptical °
path calculated. Such calculations have been performed with great care,
and it has been found that the advance of Mercury’s perihelion should
amount to about 42” of arc per century. The Newtonian theory is unable
to account for the advance of this amount, and the remarkably close
““ For a different way of deducing the value of « see J. L. Synge, Relativity: The
General Theory (1960), pp. 294-296, and G. Y. Rainich, Mathematics of Relativity
(1950), p. 162.
Sec. 107] CONCLUDING REMARKS 311
agreement between the relativistic calculations and the best observed
value can hardly be viewed as fortuitous.!®
It is worth noting that the calculations based on the restricted theory
of relativity also give a precessional effect when one assumes that a
particle moves in a field of force with potential V = km/r. However,
the precession based on such calculations yields results that are not as
close to the observed value as those furnished by the general theory.
y}
Fig. 51
where the c, are the orthonormal base vectors associated with the frame Y.
The location of the same point P in the region 7 is determined by the
vector
(109.5) gat ey Me
ES)
Ox?
and, if we set
(109.11) 25 — hy = 2¢,;
Since (109.12) is an invariant and €;; = e;,, we conclude that the set of |
functions ¢,,(x, ¢) represents a tensor Ey with respect to a class of ad-
missible transformations of coordinates X, with the basis a,, covering the
region tT). The same set of functions ¢;,(x, t) also determines a tensor E
with respect to a set of transformations of coordinates determined by the
basis b; of the final state 7. In the notation of the concluding paragraph
of Sec. 45, the tensor E, is specified by the multilinear form E, = e,,a‘a’,
whereas the tensor E is determined by E = ¢,,b’b’. Thus the operations of
covariant differentiation and those of raising and lowering indices on the
Sec. 110] GEOMETRIC INTERPRETATION OF FE, AND E 317
j j = j
aj — = E's
Cin = €&k and Lex
h
However, the two sets of functions «,’ so computed in general are distinct,
and to indicate the origin of the set €,’ obtained with the use of the tensor
h,;, we shall write
hei, = ox
pi ldtlisa |dro| = ds — dS
|dro| dSo
we have
and hence formula 110.2 can be written with the aid of (110.4) and (110.5)
as
2€;; 5
110.6 —" __. = (1 + e,(1 + e,) cos 6,; — cos G;,.
on V hiehss ;
Since 0,,° = 6,; = 0 for i = j, equation 110.6 yields
2€;3 ~
7% at =>
(
1 -- e;
) Te 1.2
or
where we recalled (110.7). If 2e,;<« 1 and the angle «,; is small, we have
an approximate equality, «,;— 2e¢,;. Thus the functions ¢«,; for i #/j
provide a measure of the decrease in the initially right angle between the
arc elements parallel to the vectors a; and a;,. The components e,; for
i ¥ j are called shearing components of the strain tensor Ep, and the com- .
ponents e,; for i = j are the normal components of E,.
Quite analogous interpretations can be provided for the functions «,;
when these are viewed as components of the tensor E = e,,b’b’. Indeed,
if we now define the elongation e as the change in length per unit final
length |dr| of the arc element so that
ds — dsp
C=.
ds
Sec. 111] STRAIN QUADRATIC 319
the calculations similar to those that have led to formulas 110.7 and 110.8
now yield
(110.9) ee
Sui
and
The defining formula 109.12 for components ¢;; of the strain tensor
E = «,,;b'b’ can be written as
00 —e hs = 0,
on on
or
at each point P(x) of the region r. In order to reduce this system 111.3 to
the form 13.10 considered in Sec. 13, we multiply (111.3) by g*, sum on i,
and. obtain
(111.4) (c* — <d)# =0,
where
(111.5) ej = gike,;.
The system 111.4 has three nontrivial solutions 2/,), Aig), Ais) @ = 1, 2, 3),
corresponding to the roots e, of the cubic
(111.6) je? — €6;'| = —F + He — Be + 0; = 0.
The coefficients #,; in this cubic are the invariants
B, =e, + 2 + €s,
(117) Oy = €n€5 + €3€) + €1€0,
Oy = €1€n€5.
It was shown in Secs. 13-15 that the roots e; are necessarily real and
the directions 2{,), Aig), Ais) associated with them are orthogonal.
The quadratic form 111.2, where we regard the 4’ as the running co-
ordinates, reduces to the canonical form
which, at each point P(«), represents a quadric surface with the 4‘ as the
running coordinates. The principal directions 2',, coincide with the axes
of the quadric 111.9, and it follows from (111.8) that the strain tensor e¢;,,
when referred to the frame Y, has the form
€y 0 0
0 €.. Dale
WET Tips
From the geometrical significance of components «;;, i # j (see equation
110.10), it follows that the principal directions are those orthogonal directions
in the undeformed state which remain orthogonal after deformation.
The strains €,, €,, €; are termed the principal strains.
Sec. 111] STRAIN QUADRATIC a2i
The invariants #,, 35, #3 defined by (111.7) play an important role in
the construction of models of continuous media. If we expand the
determinant in (111.6) and equate the coefficients of like powers of € in the
result, we find a
ye 2 pee, er:
O,=& “I €5 + €5 =e,
2 2 3 3
Si SG 4 €g € Swe a ceo 1 ml ek oe
en is od Por eal ot 2| = 5)Omer a>
icing
€ 1
(111.10)
2 3 3
1 8
en he
2 2 a (pues Sat,
Ds Sai Gq Coun Cg || ae 3!
Oe
apy~t
€ ~jpee
3 3 3 :
ey E5 €3
We will see in the following section how these invariants enter in the
expression for the ratio of the volume elements dry and dr of the initial
and deformed states.
We could have equally well considered the quadratic form
with A,’ = dzx'/ds, specifying the direction of the vector dro of the initial
state, and with ¢,,’s regarded as components of the strain tensor Ey =
€ja'a’.
For the determination of principal directions we now have the set of
equations of the type 111.4 in which
(111.12) ef he);
O= OP Sg tae €o(Yo)”
€y°(Yo')® + aM
the basis of a
when the principal directions Aj,1), Jia), Aiyg) ate taken as
suitable orthogonal frame Yo in 7».
tions e,° along the
It follows from formulas 110.7 that the elonga
principal directions are
(111.14) of = ad
dS
e T+ 268 — 1,
$22 MECHANICS OF CONTINUOUS MEDIA [CHAP. 6
whereas the elongations e,, reckoned per unit length in the final state
(cf. equation 110.9), are
Problem
Show that
O° + 49° + 128,°
se 29,9 + 495° + 89,9’
ore J. + 605°
2 1 +2049 + 492° + 895°’
ov 3
03 = 0
1 = 2),° = 43,° + 804° :
where h = |h,;| and g = |g;,| are the determinants of the quadratic forms
dsj? =h,,dx' dvi and ds? = g,, dx’ dx’.
Thus
(112.1) — = Vh/g,
The set of functions h,,(x) can be regarded as components of the tensor
H = h,,b’b’ defined in the space of the variables 2’ in the final state, so that
g" hy = hi
and
Sixhj* = hj.
Sec. 112] DISTORTION OF VOLUME ELEMENTS Eve
We conclude that
|gixh;"| = |h,,|,
so that
g |h,;'| = bh. 7
dt
(112.3) V6; i 2e;|.
dr a
free
=
1 20,
— 1 = O,.
Thus approximately
(112.5) i A
dr
volume, and, for this
This represents the change in volume per unit
figure s promi nentl y in linear elasticity
reason, #, is called the dilatation. It
and hydrodynamics.
Formula 112.3 can be cast in the form
dt 1
(112.6) —————————— — oo
dr
324 MECHANICS OF CONTINUOUS MEDIA [CHAP. 6
dt — dt) go 0
(T1237)
dry :
and, since for small deformations ¢,° = «;,, 3,°=%,, both formulas
112.5 and 112.7 give the same value for the dilatation.
Problem
Obtain formulas 112.3 and 112.6 directly from (111.14) and (111.15).
(113.1) o=r—7,
and denote the components of & relative to the basis a; by u‘ and its com-
ponents relative to the basis b; by w’. Thus
(11372) FE = wa; c= eb
(113.3) b; = a; + =
is the covariant derivative of u; with respect to the metric h,; of the initial
state and
Ow; k
( ) ie Ox" tida
is the covariant derivative of w; with respect to the metric g;; of the final
state. The prescripts on the Christoffel symbols in (113.7) and (113.8)
indicate that these symbols in (113.7) are constructed from the tensor hi;
whereas those in (113.8) from the g;,’s.
If we insert from the first of formulas 113.6 in 113.4, we get
1
2€,5 = (uj\.a 2 Up) 58") + (ag aun; dias ay) ;)
a Lk ke k
= Uj {Uz,)ja ° a + 6; Unis 7 0; Uni
I Ujk Ugg + Vays + Uji
since a’.a® = hh”
Thus
m
oe (petelee)
(113.3) and write
Ox' Ox’?
The substitution into this expression from the second of the formulas in
(113.6) yields
(113.10) 2€;3 = Wig. t Wis — WhWae, 5:
¢«;; from
Formulas 113.9 enable us to compute the strain components
to the basis a, of the initial state.
components u; of the vector & referred
ents of E relative
Formulas 113.10, on the other hand, involve the compon
s €,; are
to the basis b, of the final state. Alternatively, when the function
differen tial equatio ns serving to
specified, equations 113.9 and 113.10 are
determine the components of the displac ement vector =
326 MECHANICS OF CONTINUOUS MEDIA — [Cuap. 6
where the e,, are the infinitesimal components of the strain tensor e,;. In
classical theory of elasticity, the strain tensor e,; is taken in the form
(113.13). The strain invariant 3, = e,, + es, + e353, as follows from
(113.13), is then equal to divergence of the displacement vector u,, and
hence the dilation 3, = (dr — drt))/dr = u',.
where the Riemann tensor R},, is formed from the metric coefficients h,;.
If we recall that (see 110.1)
hoo ae DE
tj fone) ij?
compute the Christoffel symbols needed in (114.1) in terms of the g,; and
e,;, and make use of the fact that the Riemann tensor R;;,, based on the
g,;8 also vanishes, we get the condition
ae
(114.2) €ijet + h (€ 5xp ita — € 51p€ixa) = 0,
where
Cpr fuutee ot Ck,dl,Ty Stick et
ijn = “in,31 e5,4 — “3.0
and
Mae
ap
>
h
H* being the cofactor? of h,, in |h;,|.
If we linearize (114.2) by dropping terms involving the products of the
€ijn. We get Saint Venant’s compatibility equations
xe T do
the surface elements do; are denoted by T, and we take as their positive
directions the directions of the exterior normals to the volume element.
We can write
(115.2) T = —71b,,
where the b, are base vectors directed along the coordinate lines and the
v
The quadric surface 115.9 was introduced by Cauchy, and it is called the
stress quadric.
It is obvious from (115.10) that the components 7,;;, for i ¥ j, vanish
when a suitable reference frame is chosen at P. The components 7},
T22, T33 are called the normal components of stress, and the remaining ones
are shears.
By analogy with formulas 111.10, we can write down the expressions
for the stress invariants ©,. They are
We apply next the condition that the resultant moment of the body and
surface forces vanishes. If r= /'b, is the position vector of the point P’(«)
relative to some point P, the component of the moment (F x r) dr in the
direction of the unit vector Ais F x r- A dr. The component of the moment
due to the surface forces T is T x r-Ado. Recalling (Sec. 49) the ex-
pression for the triple scalar product
ACB EC: a Apic*
enables us to write
| EeE'lia® dt + i éieT'l'*do = 0.
i S
The substitution in the surface integral from (115.4) and the application of
the divergence theorem yields
| eae le a) 2|dro 0,
V
SINCE €;4,m = 0.
If we carry out the indicated covariant differentiation and make use of
equations 116.3, we get
i Ak Loh OT = U,
Vv
where 7? = 7. On the surface X of the body where stress vectors T' are
assigned,
7), le
Since all equations in Secs. 115 and 116 appear in tensor form, they are
valid in all admissible reference frames. In particular, in the reference
frame X of the initial state 7), the covariant derivatives in (116.8) are
taken with respect to the metric coefficients /,; and the a’ and F’ are com-
ponents of the acceleration and force vectors relative to the basis of the
initial state.
neighborhood of P, and denote the vector P,P’ (not shown in Fig. 51) by
Ge yetine
B(x, t) = Ga, t) + OE(@, 0),
where the variation
(117.1) 6B =F —&
ed
or the virtual displacement of P, is an arbitrary vector PP’ in the neighbor-
hood of P. We consider the variation of vectors only in the final state +
Sec. 117] VIRTUAL WORK 5
and we shall say that the variations of vectors and tensors associated with
the points Py of the initial state 75 is zero.
We suppose that & is of class C? and define the variation of 0&/dx’ by
LES AS MUL — a;
fi oie Or .
we obtain, on utilizing the distributive property of the symbol 6,
6 (=)
Ox"
= (bam a.) = dbp
for da, = 0, since points in the initial state are not varied. Thus
The metric coefficients of the final state are given by g,; = b,*b; and
we find, as in Sec. 81,
If we form the inner product ofthe vector (0&),; with both members of the
equilibrium equation
(117.8) 74 = —pF",
where F’ is the body force per unit mass (cf. 116.3), and integrate over the
body 7, we get
where we transformed the volume integral over 7 into the surface integral
over surface & bounding vr.
Since ry; = T’ by (115.4) and
CLigei2) OW = | Oe,;dr.
This term has a simple mechanical interpretation when the virtual dis-
placements (68), are the actual displacements (d&), that take place in a body
whose motion is governed by equations 116.8. In this event we write
(117.3) as
(117.14) dK =f pa‘(d&),
dr.
dK -| pa'v, dr dt.
The line integral [a dq’ then represents the work done along the path
y
C by the generalized forces Q,;. Ordinarily, such integrals depend on the
path C associated with a given process.
We shall suppose that a particle of mass dm = p dz, where p is the density
and dris the volume element, may acquire energies other than mechanical,
and we shall represent such accretions of energy in the form
(118.2) OE = ELG Soon, Gas uer ee wy Ce). 0g"
If dE includes all energies other than mechanical, the total amount of
energy acquired by the particle is determined by the integral ;
(118.4) (i fh tle «
where v is the velocity of the element of mass dm. The amount K = 3v”
represents the kinetic energy per unit mass, so that the total energy
OK:
(118.5) 6K + 6U = 6W + OE,
(118.7) 6S dm = 2
of heat acquired
T being the absolute temperature and 5Q an increment
by the element of mass dm.
brium, the kinetic
When the medium is in the state of mechanical equili
the form
energy K vanishes, and the law (118.6) assumes
Some bodies possess the property of recovering their original size and
shape when the impressed forces producing deformations are removed.
The media of which such bodies are composed are called elastic. In con-
structing a model of an elastic body we shall suppose that all processes
taking place in such a body are reversible, but we do not assume that the
body is necessarily in the state of thermal equilibrium. Thus our thermo-
elastic model will take account of the effects of temperature on defor-
mations.
As our points of departure we take the First Law of Thermodynamics in
the form [cf. (118.8)]
(119.1) 6U = 60+6W
in which
is given by (117.12).
We also write the relation 118.7 in the form
6Q =| TOS dm
or 7
(119.3) dQ =|rosp dr.
If uw denotes the internal energy U per unit mass of the body, then
(119.4) dU =| du dm =p du dr,
a rs)
(119.12) Ti=p ae
Thus either uw or ¢ (when they exist) can be used to deduce the stress-
strain relations. When the process is adiabatic, S = constant and hence
6Q =0 by (119.3). It is then more convenient to use u as a stress-
potential. In the isothermal case, T = constant and ¢ appears to be more
suitable.
We say that an elastic medium is homogeneous whenever the coordinates
a do not appear explicitly in (119.7) or (119.11). The medium is isotropic
when all parameters in the set {c} are scalars, so that the values in {c} are
independent of the choices of the reference frames ¥. When the medium is
both homogeneous and isotropic, the parameters {c} have constant values
throughout the medium.
If we consider a homogeneous elastic medium and suppose that ¢(e€,;, T)
is an analytic function of the e,,; and of AT = T — Ty, where Ty is the
temperature of the initial state, we can expand ¢ in powers of e,; and AT.
When the initial state of the body is that corresponding to «,;; = 0 and
T,; = 0, the expansions will begin with the second-order terms, so that
(a).
Formula 120.2 then follows on substituting this result in
342 MECHANICS OF CONTINUOUS MEDIA [CHAP. 6
where 0 = ge,; = e;', and use the equilibrium equations 116.3 in the
form
(121.2) gry, + F; = 0.
we can write down the linearized differential equations of equilibrium, in
terms of the displacement vector u’, by recalling that (equation 113.13)
(121.3) C5 = (Ui, + Y;,,)-
The computation proceeds as follows. The substitution from (121.1)
into (121.2) yields
SU ne= V7Ui,
we get
ao
(121.6) A+, + uV'u, + F, = 0.
in the classical theory of elasticity.
These are the celebrated Navier equations
344 MECHANICS OF CONTINUOUS MEDIA [CuapP. 6
(12221) 7! = —pg",
It follows from (122.1) that the hydrostatic pressure p is related to the
stress invariant @ = g,,7"’ (see equation 115.11) by the formula
(122.2) p= —eyjr'.
——,
Fig. 53
their momentum to the particles in the interior. In this way the fluid
between the plates is set in motion, and experiments show that the re-
tarding force per unit area of the plate, exerted on the plate by the fluid,
is proportional to its velocity and inversely proportional to the distance
between the plates. The proportionality constant in this relation is the
measure of viscosity of the fluid.
We shall say that the fluid is viscous if the stress tensor for a fluid in
motion has the form
zi = —pg" =- 7,
(122.3)
where the nonvanishing tensor ft” is the tensor of viscous stresses. The
fluid is called ideal if t’ = 0.
The mass-conservation principle of mechanics requires that
dm = po a7) = p dr,
where po(2, to) is the density of matter in the volume element dt) =
Vh dx! dx? dx® of the initial state and p(2, 1) is the density in the volume
element dr = J/g dx! da dx at time t. Thus
(122.4)
which we can also write as
= dt»
= pl®, to) _ dt a
_ p(&, t)cee
(122.5)
t) — p(%, to),
The numerator in the left-hand member of (122.5), Ap = p(x,
Ar = t — fo,
represents the change in density in the small interval of time
whereas the right-hand member
— dt. go
dt py (112.7)
dt
346 MECHANICS OF CONTINUOUS MEDIA [CHAP. 6
(122.6) —_-f,.-.-#
and since u* = v’At, where the v’ denote the components of velocity d&/dt,
we conclude from (122.6) that —(1/p)(dp/dt) = v';. We thus get the
continuity equation in the form
(227) 1d dtig ; i)
thal
pdt
We recall that p(z, t) is a function of the coordinates x’ in the reference
frame X in which the x’ are independent of t. If Y is a fixed reference
frame (cf. Sec. 109) in which the coordinates y’ of the particle are given by
—dp = (2)
{— Op
ee
dt Ot /yifixed dy’
Equation 122.7 then assumes the form
Opa eens
—F 4 —F vi + pri, =0,
Dihendivlaty bert
or
ap .
(T2988 ) ed ‘) =0.
y + (pv'),
(123.3)
1
nant ; ae):
pdt
deduced in Sec. 122. The remaining equation, known as the equation of
state, is furnished by the thermodynamical equations 118.7 and 118.8,
which, for reversible processes in the fluid, we write in the form
(123.7) Fj te
LC
pe at
when we note equations 123.1.
We shall suppose that the components e,; of the deformation tensor ina
fluid are so small that they are represented with sufficiently high accuracy
by linearized formulas 113.9 or 113.10, which coalesce in the infinitesimal
theory. Thus we write
2e55= Wig + Wy,
and
de;; d
2—2 = —(w;.; + W;.1)
dt ai! STE
com-
where the e,; are the linearized components ¢,;. Since the velocity
conclud e from the equation just written that
ponents v; = dw,/dt, we
de;; I
(123.8) B(0;,5 + 03,1):
dt
348 MECHANICS OF CONTINUOUS MEDIA _ [CuapP. 6
dW = — "vi,dtdm,
P
(123.14) p =f).
Equation 123.14 is the desired equation of state needed to complete the
system of four equations 123.2, 123.4 for the determination of the five
unknown functions v', v?, v3, p, and p.
*° These functions may (and usually do) depend on physical or chemical constants
which characterize properties of a specific fluid.
SEc. 124] VISCOUS FLUIDS 349
When viscous fluid is in motion, the components 7,; of the stress tensor
have the form
(124.1) 7) = —pg + 2%,
where the t”’, as noted in Sec. 122, are associated with viscous stresses.
As in Sec. 123, we limit ourselves to the consideration of small displace-
ments and write formula 123.8 as
(124.2) 6: = 3(0,5 + 2;
where é,; = de,,/dt are components of the strain velocity tensor.
The construction of models of viscous fluids and the formulation of
complete systems of equations now call for the introduction of additional
assumptions about the nature of viscous stresses. The latter must ob-
viously depend on the strain velocities é;; and, to a first-order approxi-
mation, it is natural to suppose that
(124.3) pi = ciklg
The coefficients c’”*' are the coefficients of viscosity, which depend on the
properties of a specific fluid under consideration. The linear law (124.3)
is quite analogous to the generalized Hooke’s law (119.14).
If the fluid is both homogeneous and isotropic, the number of inde-
pendent viscosity coefficients reduces to two, and the relation 124.3
assumes the form [cf. (121.1)]
(124.4) 13 = Avi
gt + Que",
where 4 and yw are constants and v'", is the divergence of the velocity field.
Accordingly, the complete stress tensor which includes the effects of
viscosity and hydrostatic pressure can be written in the form
(124.5) Ti; = —P8ig + Bis + 2Meis,
where # = v', = gié,,.
We recall next the equations of motion 116.8, and write them in the
covariant form
(124.6) Brie = pla; — Fi),
and substitute in (124.6) for the 7,; from (124.5). The result" is Navier’s
equations offluid motion
(124.7) (A+ md, + mein — Px = PCAs — F;),
dv' aut dv
11 Note that a i= —Raima;
= — + p'F yi, so that a; = i
ar v;,;0?.
350 MECHANICS OF CONTINUOUS MEDIA [CHAP. 6
or, in vector form,
(A+ u)V3 + uV2v — Vp = pla — F).
The set of three Navier’s equations 124.7 involves five unknowns: vt"(z, f),
(i = 1, 2, 3), p(x, t), and p(x, t). To complete the system, we adjoin (as in
the case of ideal fluids) the equation of state and the continuity equation
123.3. For incompressible fluids dp/dt = 0, and hence v'; = # = 0 by
(123.3). Accordingly, for incompressible fluids, equation 124.7 yields
(124.8) MEV; i — p= pla; — F;).
Furthermore if the fluid is ideal, « = 0 and equations 124.8 reduce to
Euler’s equations (123.2).
Stokes simplified equations 124.7 by introducing a hypothesis to the
effect that the mean pressurep in a viscous fluid is given by the same formula
122.2 as in the case of fluids at rest. This assumption leads to the con-
clusion that the constants A and w are not independent. Indeed, from
(124.1)
tig = Tig + P8is;
hence
Bolg = 8 ta Pees
= —3p + 3p = 0,
if we use formula 122.2. But since 1,; is given by (124.4), we have upon
multiplying those equations by g”,
Agitg,0 + u2g%é,, = 0,
or
(3A + 2u)d = 0.
Thus
(124.9) 3A + 2u =0.
As a consequence of this relation, equations 124.7 depend only on one
viscosity coefficient “, and the substitution from (124.9) in (124.7) yields
the set of Navier-Stokes’s hydrodynamical equations
0;
(124.10) Ue vzp + i 98 = ees = p(a; os F;).
SOs eos
If the fluid is ideal we get, on setting w = 0 and
Ov; ‘
a; = — a v; iv’,
a
the Eulerian hydrodynamical equations
(124.11) EP a ee
Ot NPL
Ot p Ox" =
for ideal compressible fluids.
Sec. 124] VISCOUS FLUIDS 351
If the motion is slow, the term v,,v’ can be disregarded, and then
a = Ov'lor.
Problems
i
? =
garg
— —— =
— Oaladwbereis. le
0 /h c= Pile
in
14 We
posi leaee es ed old lagu ka bare
where » = //p is the kinematic viscosity.
3. Show that the equation of continuity can be written
a . Apv')
apn ant + pv Aver
_, Alog Vg = 0
Op A(pv*) yi
a :
or ox?
and in spherical polar coordinates [gy, = 1, g22 = (a1)?, g33 = (x1)? sin? x*] is
a Apu’) 2v}
— “Sa at v2 cot xe] = 0.
pr aire te)
5. The curly of the velocity field v is equal to twice the angular velocity of
rotation. The vector w such that curly = 2w is called the vorticity vector.
Show that w', = 0. Hint: wt = —}e'!0;». .
6. If the vorticity vector w* = 0, the motion is called irrotational. Show that,
if the motion is irrotational, the velocity vector v is the gradient of the velocity
potential ®.
7. Write out the approximate equations of motion of a viscous fluid when the
motion is slow.
352 MECHANICS OF CONTINUOUS MEDIA [Cnap. 6
353
r i Cee! ‘ee
a7 de
= a + x Ls ; § : Lies
~ J a | ais
thal < <3 (RE gree te tna
~ ~ j ‘4 neat = , a ‘ Poa
oA¥s mAs Ribahen b refs :
7
"teinals lh sd vhiehons ae thyc%emer
t alors 4
- ey « wea © ii a b
net essaane
4.4 a
INDEX
APES:
pPust grit
*
TY
7 ~ x
“
<2
2 ity Mey
; ‘ "4 a
fi : im|
ut ' ' if 10; 28 ty Sictoedie
Uh <vitiviley x lo wwiisois¥
_— ey okt on oF
ee a ag. OY ae ey . rac}Rericar ne
fielded a
ter Rl’ Weteestineia lary
Magnesite
an
0856, oF, pee ao ee
Os; ive feutsy +
vit 2 =
ied .
Jo andbe
7
Introduction to Vector
and Tensor Analysis
By ROBERT C. WREDE, San Jose State College. This
book is a careful presentation of the fundame ideas
used in developing geometry, analysis, and linear
It is centered around four pivotal ideas: historic:
tive and motivation; the interrelationships of
geometry; a structure that stresses proof, and centr
the idea of transformation; and the use of sur
ventions and other notational devices. These
troduced early and referred to all through thet
possible the section on special relativity, in
geometric structure of the theory.
sa,