Sie sind auf Seite 1von 49

C H R

Vector Analysis, Vector Operators,


and Transformations

5 1 INTRODUCTION

Because mathematics is an integral part of physics, it is essential that we have a clear under-
>tanding of vectors and are capable of using them frequently and with ease. To make sure that
ve understand vector terminology and symbolism, we start with a brief review of vector analy-
sis, establishing a thorough understanding of vector operators such as vector differential oper-
ators (gradient, divergence, and curl). In the process of doing this, we will make sure to use ex-
ifflp-totornsMtom t&OTtti m C\MO\ medma Mty w mil aflita U m of
matrices in coordinate transformations.

5 2 VECTOR PROPERTIES

Most of the physical quantities we encounter in physics and engineering may be classed as one
of two types: scalar or vector. A scalar quantity is completely specified by stating its magni-
:ude only, together with units, if any. Examples of such quantities are mass, volume, energy,
:ime, and number. These quantities may be treated as ordinary numbers and may be added, sub-
r
jacted, multiplied, or divided by simple arithmetic rules. A vector quantity is completely spec-
ified by stating both magnitude and direction. Examples include displacement, velocity, accel-
eration, force, and electric field strength, to name a few. Vector quantities follow the rules of
vector algebra.
The vectors may be treated by either a geometrical or an analytical approach (see Sec-
tion 5.3); each has its advantages and disadvantages. We will briefly discuss both approaches in
this chapter.

139
140 Vector Analysis, Vector Operators, and Transformations Chap. 5

(a) Representation of vector quantities A and B; the dot


indicates a point of application

s = -4

s=4

(b) Four equal vectors (c) Multiplication of a vector A by a scalar J

(d) Addition of two vectors A and B; R = A + B

Figure 5.1 Geometrical approach to vectors.

Geometrical Approach

Geometrically, a vector is represented by a directed line segment with an arrowhead. In gen-


eral, we will use boldface letters (A, B, . . . , R) to denote different vectors, as shown in
Fig. 5.1(a). If we are concerned with the magnitudes of the vectors, these are represented as A,
B, C, ...,R or |A|, |B|, |C|, .. . , |R|.
Figure 5.1 illustrates the geometrical approach to analyzing the vectors.
Sec. 5.3 Vector Addition: Analytical Treatment 141

(e) Resultant of three vectors A, B, and C; R = A + B + C

(f) Subtraction of a vector B from vector A ; R = A - B # B - A

F i g u r e 5.1 (continued)

5.3 VECTOR ADDITION: ANALYTICAL TREATMENT

One advantage of vector formulation as applied to any particular situation is that vectors can be
used without reference to any particular coordinate system. But there are two main reasons that
eventually compel us to use a proper set of coordinate system: (1) the geometrical method of
obtaining the resultant of several vectors is cumbersome and not very accurate, and (2) to make
a proper interpretation, it is always helpful to present the results in a suitable set of coordinate
system. An alternative is to use an analytical method using the components of vectors.
Again, an analytical approach uses the components of vectors. Figure 5.2 illustrates the
components of a vector R in two dimensions. The two quantities Rx and Ry are the rectangular
components of vector R, and the magnitudes of these are given by (see Fig. 5.2)

R=R cos and R, = R sin (5.1)


142 Vector Analysis, Vector Operators, and Transformations Chap. 5

, x o\+ RXH x

Figure 5.2 Rectangular components Rx and Ry of a vector R.

If R and 6 are given, we can find Rx and Ry. On the other hand, if Rx and Ry are given, R
and 6 can be calculated by the relations
R
R = VRi + Rl and tan 6 = (5.2)
Thus a vector is completely defined in a plane (two dimensions) provided we know R and 6 or
Rx and Ry.
We shall use the component method for adding several vectors. Suppose we have three
vectors R b R2, and R 3 [see Fig. 5.3(a)], making angles 6U 62, and G3, respectively, with the X-
+

(a)

Figure 5.3 Rectangular components of vectors R,, R2, and R3


Sec. 5.3 Vector Addition: Analytical Treatment 143

axis [Fig. 5.3(b)]. We are interested in finding the resultant. First, we draw the XYcoordinate [as
in Fig. 5.3(b)], and then, at point O, draw Ru R2, and R3 parallel to the vectors in part (a) Next
each vector is resolved into its components as shown. Let Rx and RY denote the sum of the X and
Y components. Thus

(5.3a)
1=1

Rr = Rly + R2y + R3y = 2 R (5.3b)

where Rix - R1 cos 0,, RXy - Rx sin 8X, and the summation sign indicates the sum of all the
components from i = 1 to i = 3. Treating Rx and RY as single components, we find the resul-
tant to be

Ry
and tan 8 = - ^ (5.4)

It may be pointed out that we need not draw the diagram in Fig. 5.3(b). The calculations are sim-
ple, quick, and accurate and may be extended to any number of vectors.
We can extend this procedure to n vectors in three dimensions. As shown in Fig. 5.4 vec-
tor R may be resolved into three components: Rx, Ry, and Rz along the X-, Y-, and Z-axes, re-
spectively. Thus the resultant R of vectors R1; R 2 , . . . , Rn may be written as

+R3+ (5.5)

D
Figure 5.4 Rectangular components
Rx, Ry, and Rz of a vector R in three
dimensions.
144 Vector Analysis, Vector Operators, and Transformations Chap. 5

n
where
1=1

with similar expressions for RY and Rz. Referring to Fig. 5.4, we obtain the resultant

R = R2y + R\ (5.7)

Rv
t m 6 (5.8)
Rz

tan <f> = (5.9)

Since the components of a vector equally define a vector, it should be possible to corre-
late the two by extending the analytical method to the geometrical method discussed in the pre-
vious section. Also,

A = (Ax, AJ in two dimensions (5.10)

A = (Ax, A , Az) in three dimensions (5.11)

Using Eqs. (5.10) and (5.11), we may write the properties of vectors in component form as
follows:
Equality of vectors:

A = B or [Ax, Ay, Az] = [Bx, By, Bz] (5.12a)


which means

Ax = Bx, Ay = By Az = Bz (5.12b)
Scalar multiplication:

sA = s[Ax, Ay, Az] = [sAx, sAy, sAz] (5.13)


Null vector: The null vector has a zero magnitude and undefined direction

0 = [0, 0, 0] (5.14)
Vector addition:

R = A + B = [Rx, Ry, Rz] (5.15)


where

Bx, Ry = Ay + By, R2 = AZ + Bz (5.16)


Commutative law:

A + B = B + A (5.17)
Sec. 5.4 Scalar and Vector Products of Vectors 145

Associative law:

A + (B + C) = (A + B) + C (5.18)
and similarly,

(ns)A = (ns)[Ax, Ay, Az] = n[sAx, sAy, sAz]

= n(sA) (5.19)
Distributive law:

(n + s)A = nA + sA (5.20)
and

s(A + B) = sA + sK (5.21)

5.4 SCALAR AND VECTOR PRODUCTS OF VECTORS

Unlike vector addition and subtraction, there are several ways of defining multiplication of two
vectors. Two types of vector products are commonly used and we shall discuss them at some
length; they are called (1) the scalar product of two vectors and (2) the vector product of two
vectors. A third and fourth product type we shall also introduce are (3) the scalar triple product
and (4) the vector triple product.

Scalar Product (or Dot Product)

The scalar or dot product of two vectors A and B is defined to be a scalar quantity S obtained
by taking the magnitude of A multiplied by the magnitude of B and then multiplied by the co-
sine of the angle between these two vectors; that is,

S = A B = |A||B| COS(A, B) = AB cos 8, for 0 < 0 < v (5.22)


where A B reads as A dot B and is defined to be a scalar quantity of magnitude AB cos 0 = S.
The angle 0is the smaller of the two angles between the two vectors, as shown in Fig. 5.5. There
are two alternatives for defining the scalar products of two vectors, as illustrated in Fig. 5.5(a)
and (b): either by projection of A on B or B on A. In either case (as shown),

A B = AB cos 0 (5.23)
Let us consider some special cases of the scalar products.

The dot product is zero. Suppose pp the scalar orr the dot product of two vectors is zero;
s A
that is, A B
B== 0.
0 This
This is
is possible
possible if
if att least
l t one off three
th dii
conditions i satisfied:
is fi A = 0, B =
0. or A is perpendicular to B (that is, 0 = 90). If A is perpendicular to B, vector A is said to be
orthogonal to vector B. Thus it is possible to obtain a scalar product of zero, even though nei-
ther of the two vectors is zero.
146 Vector Analysis, Vector Operators, and Transformations Chap. 5

A B = IAI(IBI cos 0) = AB cos 9

BcosO
(a)

A B = (IAI cos 0) IBI = AB cos 6

Figure 5.5 Scalar product of two vec-


tors A and B; A B.

The scalar product is commutative. It is easy to show that


A B = B A (5.24)
That is, the order of multiplication is not important:
B A = |B||A| COS(- 0) = AB cos 9 = A B

Two vectors are equal. Suppose B = A; then cos(A, B) = cos 0 = 1; hence


A B = A - A = |A|2 = A2 (5.25)

Law of cosines. The proof of the law of cosines becomes trivial if we make use of the
dot product. In Fig. 5.6, let
R = A+ B

(a)
Figure 5.6 (a) Law of cosines as applied to R = A + B. (b) Law of cosines as
applied to D = A - B.
Sec. 5.4 Scalar and Vector Products of Vectors 147

R2 = R R = (A + B) (A + B)

= A A + B B + 2AB
= A2 + B2 + 2ABcos(A, B)
That is,

R2 = A2 + B2 + lABcosd (5.26)
Since cos 9 = COS(TT 4>) = - c o s 4>,

R2 = A2 + B2 - 2ABcos<f> (5.27)
Similarly,

D = A - B
^nd D2 = A2 + B2 - 2AB cos 9 (5.28)
Since cos 9 = cos(77 - <j>) = cos <j>,

D2 = A2 + B2 + 2ABcos(f) (5.29)

Vector Product (or Cross Product)

The vector or cross product of two vectors A and B is defined to be a vector C.The magnitude
?i this vector is obtained by taking the magnitude of A multiplied by the magnitude of B and
:hen multiplied by the sine of the angle between A and B, while the direction of C is such that
:; is perpendicular to both A and B: that is, C is perpendicular to the plane containing both A
md B. We denote the cross product as

C - A x B (5.30)

vhere A X B reads as "A cross B " and is as shown in Fig. 5.7. Thus the magnitude of C is
C = |C| = |A||B|sin(A, B) = AB sin d, for 0 < 0 < IT (5.31)
Hence C will be zero if A = 0, B = 0, or the angle 9 is zero. The direction of C is given by the
nght-hand rule or right-hand screw, as illustrated in Fig. 5.7. By drawing simple diagrams, we
;an draw several conclusions about the cross product.

Commutative and distributive properties. Because of the right-hand rule con-


vention used in defining the vector product, we may conclude (from Fig. 5.7)
A x B = -B x A (5.32)

That is, vector multiplication is not commutative; it is anticommutative. From Eq. (5.32), it is
also clear that

A x A = 0 (5.33)
148 Vector Analysis, Vector Operators, and Transformations Chap. 5

Figure 5.7 Cross product of vectors


A and B, illustrating the direction of C
by the right-hand rule.

That is, the vector product of any vector with itself is zero. Also, by definition, sin 6 = sin 0 =
0; that is, A X A = 0, and if the angle between A and B is zero, then |A X B| = AB sin 0 =
0. Note that the right side of Eq. (5.33) is a null vector, which obeys the rule
A + 0 = A, A x 0 = 0, A 0 = 0 (5.34)

The vector product does obey the distributive law:

A x ( B + C) = A x B + A x C (5.35)
and
s(A x B) = (sA) x B = A x OB) (5.36)

Area of a parallelogram. In many situations it is desirable to know the orientation


of the area, and for this purpose the vector product is convenient to use. Consider vectors B and
C, which form the two sides of a triangle OBC or two adjacent sides of a parallelogram OBDC,
as shown in Fig. 5.8. The area A of the parallelogram is

A = 2(area of triangle OBC)


= 2[2 base X height] = 2[{ Bh]

= BC sin 6= |B||c|sin0

or, if we represent the area by a vector quantity A, we may write

A = |A| = |B X C| = BC sin(B, C) = BC sin 8 (5.37)


Sec. 5.4 Scalar and Vector Products of Vectors
149

A =B x C

C
D

ilSlS

O ^ B
Figure 5.8 Area of a parallelogram.

A - SCsinfl /
= area /

Figure 5.9 Vector representation of


the area by A that is normal to the
plane of vectors B and C.

Thus the area of a parallelogram of sides B and C is B X C. Since the direction of B X C is nor-
mal or perpendicular to the plane of the parallelogram, vector A is as shown in Fig. 5.9. That is,
the area is assigned a direction of A.

Law of sines. Consider a triangle formed by vectors A, B, and C such that

C = A+ B (5.38)
Take the vector product of both the sides by A; that is

AxC=AxA+AxB (5.39)
Since A X A = 0, Eq. (5.39) takes the form

A x C=AxB

or AC sin(A, C) = AB sin(A, B)

_ t. sin(A, C) sin(A, B)
That is,
B C
150 Vector Analysis, Vector Operators, and Transformations Chap. 5

sin(A, C) sin(A, B) sin(B, C)


or, in general, (5.40)
B C
which is the law of sines.

Scalar Triple Product: Volume of a Parallelepiped

The quantity (A x B) C is a scalar quantity. We can show that if A, B, and C are the sides of
a parallelepiped then such a product represents the volume of the parallelepiped. From Fig. 5.10,
the volume V is
V = (base area)(height) = |A X B\h
= |A X B||C| sin[(A X B), C]
or V = (A x B) C (5.41)
If the vectors in the product follow a cyclic order, we may write V to be
V = (A x B) C = (B x C) A = (C X A) B (5.42)
or, since a scalar product is commutative,
V = A (B x C) = B (C x A) = C (A x B) (5.43)
That is, the scalar triple product remains unchanged if the vectors are interchanged in a cyclic
order. We may also show that
A (B x C) = - A (C x B) (5.44)

Vector Triple Product

There are two different ways in which three vectors may be multiplied so that the resulting prod-
uct is a vector in each case. One such product is
A(B C) = B(C A) = C(A B) (5.45)

Ax B

h = ICI sin [(A X B), C]

W;i l \ X HI

Figure 5.10 Volume of a parallele-


piped in terms of sides A, B, and C.
Sec. 5.5 Unit Vectors or Base Vectors 151

The interpretation of this is: (B C) is a scalar; hence A(B C) is vector A multiplied by a scalar
quantity (B C). The direction of the new vector is that of A. Similarly, B(C A) is a vector in
the direction of B, and C(A B) is a vector in the direction of C.
The second vector triple product is

A x (B X C) (5.46)

B x C) is a vector perpendicular to the plane of B and C. Therefore, A X (B X C) is a vector


perpendicular to both A and (B X C). Hence A X (B X C) is a vector in the plane of B and C.
A similar argument applies to other triple products. We may also show that
A X B X C = B(A C) - C(A B) (5.47)

5 5 UNIT VECTORS OR BASE VECTORS


As the name indicates, a unit vector is a vector of unit magnitude and has a direction. We shall
iiscuss the concept of a unit vector in this section and use it frequently throughout the text. For
i Cartesian or rectangular coordinate system, unit vectors are a set of three mutually perpen-
dicular (or orthogonal) unit vectors, one for each dimension. These are also called unit coordi-
nate vectors. For a rectangular coordinate system the unit vectors are denoted by i, j , k, where
i is along the X-axis, j is along the F-axis, and is along the Z-axis (shown in Fig. 5.11), form-
Lag a right-handed triad. (Several other notations are used for unit vectors, such as x, y, z, or
v . e,, or u,, where i = 1, 2, 3.) Sometimes unit vectors are represented simply by i, j , and k.
Thus, by definition,

PI = 111 = (5.48)
in component form

T = [1, 0, 0], J = [0, 1, 0], = [0, 0, 1] (5.49)

3--

2--

j
-H-

Figure 5.11 Unit vector!, j , it of a


rectangular (or Cartesian coordinate)
system.
152 Vector Analysis, Vector Operators, and Transformations Chap. 5

Figure 5.12 Components of a vector


A = A). + Aj + A$L.

Using this definition, we can write vector A in component form as

A = [Ax, Ay, Az] = [A, 0, 0] + [0, Ay, 0] + [0, 0, Az]

= Ax[l, 0, 0] + Ay(0, 1, 0] + Az[0, 0, 1]


that is, A = Axi + Ayj + A z (5.50)
where Ax, Ay, and Az are the components of A along the three axes, as shown in Fig. 5.12. As is
clear, the multiplication of i with Ax gives direction to the scalar without changing its mag-
nitude, and similarly for the other two terms, Ayj and Azk. That is, the multiplication of a scalar
with a unit vector gives direction to the scalar (the direction of the unit vector) without chang-
ing its magnitude. Equation (5.50) has an interesting interpretation, as shown in Fig. 5.12. The
tail of the vector is at O, while the head, which is at P, may be located by first going along path I
along Oa equal to Axi, followed by path II along ab equal to A y j, and followed by path III along
bP equal to Azk. In going from O to P, it is not necessary to follow this particular sequence of
paths I, II, III along the axes X, Y, Z; any other sequence(F, Z, X) or (Z, X, Y)will give the
same result, that is, will reach point P after starting from point O.
Using the preceding definition of unit vectors and the definition of a scalar product,

A B = AB cos 6
and of a vector product

A X B = C and \C\ = C = AB sin 0

we can arrive at the following results, remembering that unit vectors are orthogonal:

l l = ] - j = k - k = l (5.51)
Sec. 5.5 Unit Vectors or Base Vectors 153

i'J =J-k = k - i = j - i = k ' j = i - k = 0 (5.52)


i x i = j x j = xll = o (5.53)

(5.54)

We shall prove two of these results; the remaining ones can be proved similarly.
t -I = [i||i|cosO = (1)(1)(1) = 1

| l x t | = |l||!|sin0o =
Scalar and vector products may be rewritten in a more compact form by making use of
unit vector notation and their properties. Using Eq.(5.50), we may write
A = Ax\ + Ay] + (5.55)
B = Bxi (5.56)

Addition and substraction of these two vectors may be written as


A + B = (Ax + Bji + (Ay + By)j + (Az + B$) (5.57)
A - B = (A, - Bxt + (Ay - Sp] + (Az - Bfi) (5.58)

The scalar product of the two vectors may be evaluated as

A B = (A,t + Ay] + Afr (Bx\ + ByJ + B)


i 1) + AyB/j J) + Afifi ) + Afifi X J) + Af X )
tf X t) + A/Z(J ) + AAtf t) + A ^ ])
Using the results of E q s . (5.51) a n d (5.52), ( ! ] ) = = 1 and (i j ) = = 0, w e get

A B = Afix + AyBy + Az (5.59)

Let us now calculate the cross product:

A X B = (A,! + Ay\ + A) X (Bx\ + By] + B)

= AJB/i X 1) + Afifi X }) + Afiz$. Xfc)+ Afifi X }) + AA(i X fc)


d x t) + (A/?z(j x fc) + AA( X i) + A^/fi X ])
154 Vector Analysis, Vector Operators, and Transformations Chap. 5

Using the results ofEqs. (5.53) and (5.54), (1 x i ) = = O a n d ( l X j ) = k = - < j x ? ) , a n d


so on, we get

A X B = AjBJO) + AyByiQ) + Afiz{<3) + A^/k") + A ^ - j )


s\ A A /v

+ AyBx(-k) + AyBz(i) + A^O) + ^ / ^ i )

= \{AyBz - Afiy) + ]{AZBX - Afiz) + k ( A ^ - AyBx) (5.60)

which may be written in a convenient form as (using the definition of determinants)


Az Ax Ax Ay
A x B = i y **z + k (5.61)
Bz Bx

or in a still simpler form as

1 J K
A x B = Ax Ay A, (5.62)
Br fi B

Note that if A = B or A = sB, where s is a scalar number, s can be taken out of the determinant,
thereby making two rows identical; hence the whole determinant will be zero. This is as it should
be for 6 = 0, that is, when the two vectors are parallel.
Finally, let us once again consider vector A as shown in Fig. 5.13. The component of this
vector along the X-axis is given by

Ax = A cos 6

Another way of writing this is to take the dot product of A with i ; that is,

A i = |A||i| cos 8 = A cos 6

which is the same as the preceding equation. Thus

A = A i (5.63)

Figure 5.13 Component Ax of vector


A long the X-axis.
Sec. 5.6 Directional Cosines 155

Figure 5.14 Component An of


vector A along the /V-axis.

The result of Eq. (5.63) may be written in a general form. Suppose we want to find component
A,, of A along an arbitrary axis N that has a unit vector en along this axis, as shown in Fig. 5.14.
We may write
An = A eB (5.64)

5 6 DIRECTIONAL COSINES

Let us start with vector A expressed in the form of Eq. (5.50) written in a slightly different
form as

A = AJ = A ^ f c ) = AeA (5.65)

where eA is a unit vector in the direction of A. AJA is equal to the cosine of the angle between
A and the X-axis. Thus

= cos(A, X) = cos(A, i) = a (5.66)

= c o s ( A , Y ) = cos(A,j) = /3 (5.67)
A
A
1 = cos(A, Z) = cos(A, k) = y (5.68)

where a, /3, and y are called the directional cosines of the line representing A. Thus Eq. (5.65)
may be written as
A = A(ai + /3j yk) = Ae (5.69)
That is, e + yk (5.70)
A = <*t +
which expresses the unit vector eA along A in terms of the directional cosines of A and the unit
vectors. From Eq. (5.65), we may also write

(5.71)

By definition, A A = A2; hence Eq. (5.71) yields

= 1 (5.72)
* ) *)"(*
a2 + (32 + y2 = 1 (5.73)
That is, the sum of the squares of the directional cosines of any line is equal to 1.
156 Vector Analysis, Vector Operators, and Transformations Chap. 5

Example 5.1
For the vectors A = (6,4,-7) andB = (2,-2,3), calculate the expressions (a) to (j) given
below. The two vectors are expressed in matrix form. Make the calculations and then
express the results in terms of unit vectors where possible.
6 2

Solution A := 4 B : = -2
-7 3

(a) A (a) A -t- B = | 2 A +- B = 8 - \ + 2 j - 4 k


-4

(b) A-B (b) A- B= -j - io-k


-10

(c) 4-A + 2-B 28


(c) 4-A + 2-B = 12
4-A + 2-B = 28-i +- 12-j - 22-k
-22

(d) 4-A - 2 B 20
(d) 4-A - 2-B = 20 4-A - 2 + B = 20-i + 20-j - 34-1
-34

(e) Calculate the magnitudes


A, B, and the sum of the (e) A- A = 101 B B = 17
squares of A and B.

^ A - A = 10.05 ^ B - B = 4.123

A := 10.05 B := 4.123

A 2 +- B 2 = 118.002

(f) Calculate the dot product of the


A-B = 1 7
two vectors. (f)

(g) Calculate the angle 8 between


A-B
the two vectors by making use 9 := acos e = 1.994-rad
of the relation JA-A-JB-B
e = 114.221-deg
A-B = ABcos(9)
Sec. 5.6 Directional Cosines 157

(h) Calculate the cross product of ( h )


the two vectors.

A X B = -2 i - 32-j - 20-k

(i) Calculate the component of B (*>


Ba=B-cos(6)
along the direction of vector A
(9 is as calculated above).
Ba := B-cos(e) Ba = -1.692

Alternate approach: A
Bal := B
A Bal = - 1 . 6 9 2

(j) Calculate the directional cosines of (j) Ax := 6 Ay := 4 Az := - 7


A and B. (The components are
from the values of the vectors.)
A := A / 1 0 1
a and (3 each gives three
directional cosine component:
0.597
for each vector A and B. A
a= 0.398
a :=
A 1-0.697

Bx := 2 By := - 2 Bz := 3

Bx
B := By B = 4.123
Bz
0.485
As shown, the sum of the squares
of the three directional cosines of P= -0.485
0.728
each vector should be unity, and
the sum of the two should be 2.
( a ) =i ( B ) =i

EXERCISE 5.1 Repeat the calculations for the vectors A = (3,4,-9) and B = (4,-3,6).
158 Vector Analysis, Vector Operators, and Transformations Chap. 5

5.7 VECTOR CALCULUS

Differentiation of Vectors

Let us consider a vector that is a function of a scalar quantity, say s, where s may be time /,
angle 0, or some other quantity. An example is the velocity v(t) of a particle or the position r(0)
of a particle as a function of an angle. The vector A, which is a function of time t, may be writ-
ten algebraically in component form as
A = A(f) = Ax(tfi + Ay(f)j + Az{t)k (5.74)
The derivative of A with respect to t is defined in a manner similar to the derivative of a scalar
function. That is,
dk A(f
= limit ^ P.)
dt A'^o
This is illustrated in Fig. 5.15, where AA or AA/Ar becomes tangent to the curve as A? ap-
proaches zero. If A is replaced by r, then Ar/Af represents the average velocity, while drldt is
the instantaneous velocity. We can also write the derivative of a vector in terms of its compo-
nents as
dA y dA ^ dA^,
(5.76)
dt dt' d ' dt I d tl +' d dt *t i +' dt
As an example, if A = r, the displacement, velocity, and acceleration vectors are
r = xi + yj + zfi (5.77)
v = r = ii + j j + zlt (5.78)
a = r = v = xi + y] + zk (5.79)

AA = A(f + Af) - A(f)

Figure 5.15 Derivative of A with re-


spect to t.
Sec. 5.7 Vector Calculus 159

Note that the magnitudes of the velocity and the acceleration are

v = = Vx 2 + v' (5.80)
2 2
a = = Vx + y + z (5.81)
Thus the differentiation of a vector follows the same procedure used in the differentiation
of a scalar function. We can extend these rules to the following particular cases:

(5.82)
ds ds
A J A

f
ds +
fir
a\ (5.83)

dA n -u dsA . dB
B T A (5.84)
ds ds ' ds

(5.85)
ds
It is important to note that in the last equation the order of the factors in the cross product must
not be changed.
We can use these results to discuss the concepts of relative position, velocity, and accel-
eration. Let point P\ be at a distance rt and point P2 at a distance r 2 , both with respect to the ori-
gin O (Fig. 5.16). The relative position of P2 with respect to P} is r21 and is given by

(5.86)

Since points P, and P2 are in motion, the relative velocity of P2 with respect to Px is

_ dr2l _ dx2
Vrel
~ ^/T ~ltt ~ ~dt
That is, V
21 r
21 r
2 r
i (5.87)

Figure 5.16 Relative position of P2


with respect to Px is Tj r2.
160 Vector Analysis, Vector Operators, and Transformations Chap. 5

and, similarly, the relative acceleration of point P2 with respect to point P{ is


a
2i = V
2i = *u = r2 - r'i (5.88)

Tangential and Normal Components of Acceleration

By definition, for a particle moving in a curved path, its velocity vector v is equal to the prod-
uct of the speed v and a unit vector u, (in the direction of the tangent); that is,

V = VU, (5.89)
As the particle moves, the speed as well as the direction may change; hence the acceleration of
the particle is written as
dv d(vut) dv du,

du,
or a = vu, + v- (5.90)
dt
Since the unit vector u, is of constant magnitude, the derivative dut/dt means that the direction
of u, is changing with time. As shown in Fig. 5.17(a), initially the particle is at point P and
in time At it travels a distance A.? reaching a point Q. Let the unit vectors at P and Q be u, and
u't. As shown in Fig. 5.17(b), the two unit vectors differ by an angle Ad; the magnitude of the
diffence in the two unit vectors is

lAfi,l = |uf' - uj = 2siny


As A 6 approches zero, the right side approaches A0; hence
Au,
limit = 1
A0->O AG

Figure 5.17 (a) Unit vector for the


motion of a particle along a curved
path, (b) The difference Au, between
(a) (b) unit vectors u, and u,' in part (a).
Sec. 5.7 Vector Calculus 161

Also in the limit, Au, becomes perpendicular to u r and is called the unit normal vector un, as
shown in Fig. 5.17(a); that is,

a. = (5.9D
dO
Now, using the chain rule, we may write
du1_du1d6 _ A dd(h
dt d6 dt "ds dt
But dsldt = v and dsldO = p = the radius of curvature of the path; thus

(5.92)
dt
Substituting this in Eq. (5.90) yields

a = uu, + un (5.93)
P
Thus the acceleration is described in terms of two components, tangential and normal, where

(5.94)
dt1

(5.95)
The normal component of the acceleration is always directed toward the concave side of the path
and is called the centripetal acceleration. The magnitude of the acceleration a is
,4\ 1/2
(5.96)

Integration of Vectors

In three-dimensional motion, we must distinguish between two types of functions or fields.


Scalar point functions are of the form
f(r)=f(x,y,z) (5.97)
while vector point functions are of the form
A(r) = A(JC, y, z) = [Ax(x, y, z), Ay(x, y, z), Az(x, y, z)] (5.98)
A typical example of a scalar point function is the potential energy function V(r) = V(x, y, z) of
a particle, and that of a vector point function is the electric field intensity E(JC, y, z). These func-
:ions may also be functions of time. In mechanics, the corresponding examples are that of den-
>ity and velocity.
Consider a curve C in space. Suppose a vector point function A is defined at all the points
along this curve. The line integral of A along the curve C is defined as (see Fig. 5.18)

A dv = line integral A along C (5.99)


162 Vector Analysis, Vector Operators, and Transformations Chap. 5

Figure 5.18 Line integral of A along


the curve C is the sum of the quantities
A-dr.

and if r is a position vector,


r = jri + vj + zk (5.100)
then dr = dxi + dyj + , (5.101)
where d*, dy, and Jz are the differences in the two ends of the line segment. Thus, if
A=AX\+Ay] +Azk (5.102)
we may write the line integral as

f A dr = f (Axdx + Ayrfy + Azdz) (5.103)


Jr Jr

An alternative way of expressing the line integral is in terms of s, where s is the distance mea-
sured along the curve from some fixed point, as shown in Fig. 5.18. Thus, if 6 is the angle be-
tween A and the tangent to the curve at each point,

A dr = A cos (8) ds (5.104)

If we know A and cos 9 as functions of s, the line integral may be evaluated.


It is possible that we know A(r), but r is a known function of parameter s; that is, r = r(s),
where s is the distance measured along the curve (s could be time t or some other quantity). In
such cases, the line integral can be evaluated as

(" A rfr = ("

Ax + Ay + Ax ds (5.105)

This and other methods of evaluating the line integral will be illustrated in subsequent examples.
Sec. 5.7 Vector Calculus 163

Example 5.2
Calculate the line integral of F = an + bxyj from (-R, 0) to (+/J, 0) along the semicircle shown ir
Fig. Ex. 5.2 using the parameter d.

(-R, 0) ( + R, 0) X Figure Ex. 5.2

Solution
As the force F is applied, the
F=axi-)-bxyj
angle 9 changes from n to 0 and the radius
vector sweeps out a semicircle as shown.
Express x and y in terms of R and 9. x(e)=R-cos(6) dx=-Rsin(G)de
In terms of R and 9, the x and y
force components, Fx and Fy (F = Fxi + y(9)=R-sin(9) dy=Rcos(6)d9
Fyj), take the form Fx=a-x=a-R-cos(9) Fy=bx-y=b-R cos(6)-sin(9)
After substituting for Fx, Fy, dx, and dy,
the line integral A takes the form Fdr= Fx dx -|- Fy dy

-a-R2cos(G)-sin(9) + b-R3-cos(9)2-sin(e))de
(i)

Because of the symmetry,


we can integrate from A=2
' (-a-R2-cos(9)-sin(e) + b-R 3 cos(9) 2 -sin(9)) de
%I2 to 0 and multiply by 2. The (ii)
resulting integral, which is equal
to the work done by the force, is
A=aR 2 --bR 3
as shown.
164 Vector Analysis, Vector Operators, and Transformations Chap. 5

Alternative approach

The alternative approach to the integration above is to substitute u = cos(9) in Eq. (i), as
shown below. The integration limits used are from 0 to 1 instead of from 1 to 1.

ei :=o cos(ei)2 = cos(ei)3 =


u=cos(8) du=-sin(9)-d6
62 :=- cos(92)2=0 cos(82)3 =
2 1
63 : = jt cos(83)2 = cos(93) 3 =-l A= 2(aR 2 u-bR 3 u 2 )du
0

Integration yields A=aR 2 --bR 3

5.8 VECTOR DIFFERENTIAL OPERATORS: GRADIENT,


DIVERGENCE, AND CURL

We are now ready to introduce new tools of mathematics that will enable us to study physics in
depth. It is not expected that one should grasp these concepts in the first reading; but as they are
applied, one becomes familiar with them and can appreciate their usefulness. These new tools
are the vector differential operators, and we will study them under the headings gradient, di-
vergence, and curl.
The vector differential operator denoted by grad or V (del) is not a vector; it is a vector
operator. This vector operator, in Cartesian coordinates, is represented by

_ 4. d /> d f, d / 3 5 3
grad = V = l h j + k = , , (5.106)
dx dy dz \dx dy dz,
We will write this operator in other coordinates later. When this operates on a scalar function,
it forms a vector. When it operates on a product of functions, it must be treated as a differential
operator.
We can perform three different operations with this operator:

1. The gradient operator operates on a scalar function u to form grad u or VM.


2. When the gradient operator performs a scalar product with another vector function to form
V A or div A, the result is called the divergence of A. This is a scalar quantity.
3. When the gradient operator performs a vector product with another vector function to
form V X A or curl A, the result is called the curl of A or rot (meaning "rotation") of A.
This is a vector quantity.

The div A and curl A will be discussed shortly. For the time being, let us concentrate on un-
derstanding the meaning and usefulness of the grad u.
Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 165

Gradient Operator (grad u = Vu)

Consider a scalar function u that is an explicit function of the coordinates x, y, and z, that is, u
u(x, y, z) and this function is continuous and single valued. This scalar function has three com-
ponents, which may be considered to be the components of a vector called grad u or VM (del u).
That is, even though u is a scalar, grad u is a vector with three components given by

* du du f. du du du du
grad u = = \ hk = , , (5.107)
dx dy dz dx dy dz

Physical interpretation of gradient. Suppose a particle described by a point func-


tion u(x, y, z) moves from a point r = (x, y, z) to a nearby point r + dx = (x + dx, y + dy, z +
dz). The change in the point function is

u(x + dx, y + dy, z + dz) u{x, y, z)

As dx > 0, dy > 0, dz > 0, the differential change du in u takes the form

du , du du
du = dx + dy + dz (5.108)
dx dy dz
which may be written as

du * du / du.
du = i + k {dx\ + dy] + dzk)
dx dy dz
The first term on the right is guard u and the second term is dr. Hence, we may write this equa-
tion as
du grad u dx (5.109)

Thus this equation allows us to define the change in the function u induced by the changes
in its variables. Actually, Eq. (5.109) is the definition of the vector operator gradient. It states
that grad u is a vector such that the change du in u, for an arbitrary small change of position
dx, is given by the relation in Eq. (5.109). Equation (5.109) may also be written as

du = |grad u\\dx\ cos 6 = |grad u\ dr cos d (5.110)

du will be maximum when grad u and dx are in the same direction so that cos 6 = cos 0 =
1. and
(du)m = I grad u\ dr when grad u is parallel to dx

Thus |grad u\ = ' ' (5.111)

Grad u is in the direction in which the change in u is most rapid, and its magnitude is the
directional derivative of u, that is, the rate of increase of u per unit distance in that
direction.
166 Vector Analysis, Vector Operators, and Transformations Chap. 5

Let us consider a line path or a surface for which u = constant. Let dr be directed tan-
gentially along this path of constant u so that du = 0. From Eq. (5.109), since neither grad u
nor dr is, in general, zero, they must be normal to each other so that cos 90 = 0 gives du = 0.
Thus grad u is normal to the line {in two dimensions) or to the surface {in three dimensions) for
which u = constant.
Thus the properties of grad u may be summarized as follows:

1. Grad u is, at any point, normal to the line (in two dimensions) or surface (in three di-
mensions) for which u is constant.
2. Grad u has direction in which u changes most rapidly, and its magnitude is the directional
derivative of u.

Once we know grad u, the rate of change of u in an arbitrary direction n is given by (the
directional derivative)

du
n grad u = (5.112)
dn
Let us illustrate the preceding points with the help of an example. Suppose u stands for a
potential function V{x, y, z). Consider two contour lines of constant potential, as shown in
Fig. 5.19(a), such that Vx{x, y, z) = C1 and V2{x, y, z) = C2. Consider a change in V{x, y, z) due
to displacement dr along a contour line. This gives
dV = grad u dr = dr (5.113)
Since on a contour line V = constant, dV = 0, and V V dr = 0, if dr is along the line of con-
stant V. Since neither W n o r dr are zero, the vector Wmust be normal to dr. This is the gen-
eral result discussed above; that is, at every point in space, W i s perpendicular to the constant
potential surface (or energy surface) passing through that point.

Figure 5.19 (a) For displacement dr


along the contour line, dV = 0; hence
VV dx = 0, which is possible if VV is
perpendicular to dr. (b) W i s in the di-
rection in which V increases most
rapidly, and dr points in the direction of
increasing potential energy.
Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 167

As shown in Fig. 5.19(b), suppose dr points in the direction of increasing potential energy.
Thus, for change dV > 0, VV dr = 0, which is possible only if W is in the direction in which
V increases most rapidly. Furthermore, if F = W , then F is also normal to the constant en-
ergy surface everywhere, except that it points from a higher to a lower potential energy. Also,
the closer the energy surfaces, the larger will be the gradient; that is, the force is larger where
the potential energy is changing rapidly, VV = (dV/dr)max.
Remember that u(x, y, z) is a scalar function and hence defines a scalar field. Examples are
temperature and pressure changes in certain volumes of matter. These regions of space may be
defined by means of a temperature gradient VT(x, y, z) or pressure gradient VP(x, y, z). Every
point in space may be described by means of one value of temperature if we are dealing with a
temperature scalar field or one value of pressure if dealing with a pressure scalar field.

Divergence of a Vector (V A = div A)

As stated earlier, when the gradient operator performs a scalar product with a vector point func-
tion A, it results in divergence A. That is, in Cartesian coordinates,

div A = V A = (l ~ + j ^ - + (Axi + Ay] + A z k)


ox ay
dAx dAv dA
X
1 +
(5.114)
dx dy dz
Examples of vector point functions are an electric field vector E(x, y, z) and a velocity vector
v(x, v, z).The diveregences of such vector point functions describe respective vector fields.
Sometimes it is convenient to write (x,, x2, x3) as coordinates instead of (x, y, z), in which case
we may write

dAx dAy dA2 dA, dA2


div A s V A s= - l -
dx dy dz dx, dx-, dx-,

(5.115)

To fully appreciate the meaning of the divergence of a vector quantity, we consider an


example of fluid flow. Consider a region of fluid flow and place a rectangular volume element
dV = dx dy dz located at the origin O(x, y, z) with its faces normal to the three axes, as shown
in Fig. 5.20. Let vector A represent the volume of the flow crossing per unit area per second or
the mass of the fluid crossing per unit area (or the momentum carried per unit volume). With-
out being specific, let A represent the rate of fluid flow crossing per unit area, the area being nor-
mal to A. We consider the volume element to be small enough so that the flow rate A is constant
over each face. Let Ax, Ay, and Az be the three components of vector A. If the flow rate is Ax(x)
at x, then at (x + dx) the flow rate Ax(x + dx) will be, expanding in a Taylor's series,

Ax(x + dx) = Ax(x) dx (5.116)


dx
168 Vector Analysis, Vector Operators, and Transformations Chap. 5

dx

(x + dx)

Figure 5.20 For calculating the divergence of a vector quantity,

where we have ignored the higher terms; thus

Rate of fluid flowing in at x = Ix = Ax dy dz (5.117)


i ^
Rate of fluid flow out at x + dx = Or = I A, + ^dx\dy:dz (5.118)

The net outward flow parallel to the X-axis through volume element-rfV = dxdy dz is

O r - L = \A, ' dx\dydz-


' Axdydz
=
* - (5.119a)
OJC } -OLK

with similar expressions for the other two sets of parallel faces:

(5.119b)

(5.119c)
dz
Thus the net outward flow rate out of a small rectangular volume element dV = dx dy dz is ob-
tained by summing these three equations; that is,

Net flow rate = dV= (V A)SV (5.120)


dx + dy + dz
Equation (5.120) states that the div A is the net outward flow rate (volume or mass) per unit vol-
ume. But the net flow is also equal to the component of A normal to the surface area dS; that is,
Net flow rate = n A dS (5.121)
Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 169

where n is the unit vector normal to the surface pointing outward. From Eqs. (5.120) and
(5.121), summing over all the volume and the surface, and replacing by integrations (through-
out the whole volume and the entire surface area), we get

[ f [ V A J V = f f n -AdS (5.122)
v s
which is the mathematical statement of Gauss's theorem or the divergence theorem.

Gauss's Theorem or the Divergence Theorem. The divergence of a vector field multi-
plied by a volume is equal to the net flow of that vector field across the surface bounding
that volume.

That is, since n A = the component of A normal to S, Eq. (5.122) states that

/Total amount of V A\ _ /total outward flux\


\inside volume V / \ through surface S /
Suppose A = v, the velocity of the moving fluid at any point; then Eq. (5.122) takes the form

(5.123)

which states that.

Total volume of the fluid volume of the fluid


begin produced within flowing across S
volume Vper second per second

It is assumed that the fluid is incompressible. Also, V v is positive at the source from which
the fluid is flowing out, while V v is negative at the sink into which the fluid is flowing. On the
other hand, if

A = pv = mass flowing through a unit area per second


= momentum per unit volume

Gauss's theorem takes the form

n -(pv)dS (5.124)
V 5

which for an incompressible fluid may be written as

P( \ f V- vdV= p[ I" n \dS (5.125)


170 Vector Analysis, Vector Operators, and Transformations Chap. 5

Again depending on the sign of V v, there will be a source or sink. (A numerical example il-
lustrating the preceding material is given next.) An incompressible fluid must flow out of a given
volume element as rapidly as it flows in; that is, there are no sources or sinks and div A = 0, and
the flow field is said to be solenoidal.

y Example 5.3
Consider a box of sides 8 cm by 6 cm by 4 cm, as shown in Fig. Ex. 5.3. Fluid enters the bottom of the
box and leaves through the left, right, and top surfaces. No fluid flows out of the front or back surfaces.
Calculate the net outward flow through the box for the velocities shown.

V
3
fl
3'

Az = 4 cm ,
/ /
7
>
/
\ /
\ /
>
\ 30
) w
A4 A2 fl2
,y = 6 cm
n /
X
/> <
/
/ An
/ /
/ 1
/ 1 1

.,J
Ax = 8 cm
f

n
l = -j Ax = 8 cm = 50j cm/s
n2 = +1 Ay = 6 cm V
2 = (20.8T + 12j) cm/s
"3 = +J Az = 4 cm V3 = 20j cm/s
n4 = i AxA>> = 48 cm 2 v4 = ( - 15i + 26]) cm/s
n
5 = + k AJCAZ = 32 cm 2

"6 = -k AJCAZ = 24 cm 2

"l = 50 cm/s
A y = AxAyAz V2 = 24 cm/s
= ( 8 X 6 X 4 ) cm 3 v3 = 20 cm/s
D4 = 30 cm/s

Figure Ex. 5.3


Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 171

Solution
Using the divergence theorem as applied to an incompressiblefluidflowfrom Eq. (5.123),

n v dS (i)

We can solve the problem either by using the left or right side of this equation. We shall do both to show
them to be equal. Also note that only normal components of the velocity contribute to the netflow.We
start with the right side of Eq. (i) which may be written as

ffn (ii)

where A, represents the areas of the different faces of the box. Note that the fluid isflowingin only from
surface 1.

The unit vectors i, j , and k are 1 o' 0"


0 j := 1 k: = 0
Taking into consideration the 0 0 1
direction of fluid flow, the unit
normal vectors perpendicular nl :=-j n2: = i n3 :=j n4:=-i n5 : = k n6 :=-k
to the six surfaces are:
vl : = 50-j- v3: = 2 0 T
sec sec

I vl I =5Ocm-sec~' |v3| =20-cm" sec '


The components of the velocities
of the fluid flow, in and out, are
v4 :=(-]
v2: =
sec

-1
v2 =24.013'cnrsec |v4| =30.017'cnvsec
Al, A2, A3, and A4 are
the areas of the four surfaces that x:=8-cm y: = 6-cm z :=4-cm
are needed to calculate the flow
rate. A1 :=x . z A2 :=yz A3 :=xz A4 :=y-z

V is the volume of the box. A1


= 32>cm A2 = 24* cm 2 A3 = 32- cm 2 A4 = 24- cm 2
AI is the rate of the fluid flowing
V:=xyz AI : = nl-vl-Al
into the box. Note that the
negative sign indicates the fluid is
V = 192-cm3 AI=-1.6-1O3 -cmJ-sec
flowing in, while the positive sign
means fluid is flowing out.

The amount of fluid flowing out is given by the flowing relation

AO :=n2- (20.8-i + 12-j)- -A2-i-n3- 20-j- -A3 + n4- (- 15-i + 26-j)- -A4
172 Vector Analysis, Vector Operators, and Transformations Chap. 5

or by using the simpler relation:


AO : =

Thus the outgoing fluid rate AO AO = 1.499-103 -cm 3 'sec '


and the net flow rate AN from the
box are as shown. The negative AN: = AI + AO AN =-100.8-cm -sec
sign means that the box works as
a sink, that is, an amount of fluid is
absorbed.

Alternative approach

The same result can be obtained by using the left side of Eq. (i) rewritten,

Avl Av2 Av3


A-vdxdydz= AV
Ax Ay Az

Using the data given above, Ax : = 8-cm Ay : = 6-cm Az :=4cm


we see that this gives the
AV: = Ax-AyAz AV = 192-cm3
same results.
i . ( v 2 - v41 -i
1
' =4.475-sec ' - =-5'sec- l
Ay
Ax

Ax Ay

Fnet=-100.8-cm3-sec '

EXERCISE 5.3: Repeat the calculations if in addition there is afluidflowingin from the front surface
at a speed of 10 cm/s normal to the surface and also leaving from the back surface with a speed of 10 cm/s,
making an angle of 45 with the normal.

Curl of a Vector (V x A = Curl A)

We now take the vector or cross product of the gradient operator with a vector, resulting in a
vector called the curl A or rot A (meaning the rotation of a vector field). Thus (in Cartesian
coordinates)

_ (* d * d * d \ ~ /v >v
curl A = V x A = i + Jj + k x (Axri + AA
yi + Az It)
dx dy dz) '

dAy dAI
, (5.126)
dz dz dx dx dy
Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 173

which may also be written in short form as


i J k
a d d
curl A = V x A = (5.127)
dx dy dz

To understand the physical signficance of the curl of a vector quantity, let us consider a
fluid flow described by the velocity vector v. Place a small paddle wheel in the fluid. In addi-
tion to being carried in the fluid, the wheel will tend to rotate in the regions where curl A # 0.
A fluid that has a nonvanishing curl is said to have a vortex field; a fluid that everywhere has a
vanishing curl has an irrotational field.
Consider a vector field that has a nonvanishing curl. Let this field be represented by a
velocity vector v at any point in the field. Suppose the component vy increases with z, while the
component vz increases with y, as shown in Fig. 5.21. Both dvy/dz and dvjdy are positive; but
rivjdz results in a negative curling (clockwise rotation) about the X-axis, whereas dvjdy results
in a positive curling (counterclockwise rotation) about the X-axis. Thus the X component of
curl v is
dvz dvy
(curl v) = (5.128)
dy dz

We can assign similar meaning to the other two components.


The geometrical meaning of curl A may also be given by means of Stokes' theorem; that is,

= ff A dr (5.129)

Zi

v
- y

t dy
h

Figure 5.21 For calculating the curl


of a velocity vector v.
174 Vector Analysis, Vector Operators, and Transformations Chap. 5

Stokes' Theorem. The line integral of a vector field along a closed path is equal to the
surface integral over an area bounded by the path.

In Eq. (5.129), C is a curve bounding the surface S in space, n is a unit vector normal to
S, and dr is taken along the path C. Thus the integral on the right is taken around the path, that
is, the boundary enclosing any surface that appears on the left. For a small surface AS, we may
write Stokes' theorem as

n (V x A)AS = Ar (5.130)

as used in the numerical example.


Finally, we may point out that

v2
V v = div grad (5.131)
L d
11 _i_
d d yj d /j d /v d
i i i + j + k
J dy + K dz dx dy dz
\ dx
>
d7 i d2' \
I"2 2 1 dy2 + dz2, 1
[dx '
must be a scalar, and V2, called the Laplacian operator, is such a scalar operator. On the other
hand, the cross product of V with itself is zero by definition because the two vectors are paral-
lel to each other; that is,
V x V = 0 (5.132)

y Example 5.4

Consider a 8 cm by 8 cm path ABCD located in a velocity field shown in Fig. Ex. 5.4. Evaluate the left
and right side of Stokes' theorem; that is,

x v) iidS = [v dr (i)

The integration on the right is along the path of the surface used in the left side. We may write Eq. (i) as

(V X v) nAS =
2." Ar (ii)
Let us consider the left side first.

9Vy\ IdVy
k(o - *) (iii)
\dy dz) +-J {dx dy 1
+
\ dy)
*\dz
or, in an approximate form, we may write
_ f. dvx f. Av,
Vx v= -k (iv)
dy = - k Ay
Sec. 5.8 Vector Differential Operators: Gradient, Divergence, and Curl 175

v = 4 cm/s

8 cm

Figure Ex. 5.4

L j . k are unit vectors while the i:=l j:=l k:=l


. elocity is in the x-direction only.
Ay:=8-cm Ax: = 8-cm vy:=0 vz: = 0
The surface is in the x-y plane.
Thus the resulting change in the . .cm
vxl : = 2 O vx2 :=4
elocity is Av. sec
S is the surface area.
Kl is equal to the product of curl Av : = v x 2 - vxl Av =-16 - cm1 sec
?f v and the surface area S, and
nence the left side of Eq. (ii) as S : = (8-cm)-(8-cm) S=64-cm z
jiven in Eq. (iii).
Av
K.2 is the right side of Eq. (ii). K2 : = vxl-Ax- vx2-Ay
Ay

Since Kl = K2, this proves the Kl =128-cm 2 -sec


Stokes' theorem.

Alternative treatment Ax: = 8-cm Ay:=8-cm


k :=
We use the same data as above. AS : = Ax-Ay AS=64-cm

From Eq.(iv) curl of v : = - curl_of_v =


Ay
From left side of Eq.(ii) -l
curl_of_v-k-AS = 128*cm *sec

i:=0..3 Ar:=8-cm

cm
sec
From the right side of Eq.(ii),
V " v.-Ar =128-cm 2 -sec ' cm
(while going around the surface) 0
sec
i=0
cm
Alternative treatment -4
sec
Second alternative treatment is cm
0
vfx : = 4- vix: = 20-- sec
176 Vector Analysis, Vector Operators, and Transformations Chap. 5

Ay:=8-cm As: = 64-cm

vix-vfx . -i vix vfx , 2 -


=2-sec -As = 128-cm -sec
Ay Ay

5.9 COORDINATE TRANSFORMATIONS

The results of the application of any physical law to a given system must be independent of the
coordinate system and the location of the origin of the coordinate system as well. Vectors have
this special feature and hence are frequently used in various situations. Thus it becomes rele-
vant to know the procedure by which vectors transfer from one coordinate system to another,
that is, to investigate the properties of such transformations. Futhermore, it is convenient and
useful to describe these vectors as well as their transformations in matrix notation.
We start with the description of a scalar in different coordinate systems. Suppose mass M
is placed at point P and its coordinates are (x, y) in the XY system and (x\ y') in the X' Y' sys-
tem, as shown in Fig. 5.22. The coordinates of the mass are different in the two coordinate sys-
tems, but the mass remains constant; that is,

M(x,y) = M(x',y') = constant (5.133)


Quantities that are invariant under a coordinate transformation are called scalars.
Let us now investigate the procedure for coordinate transformation of vectors. Consider a
point P that has coordinates (xit x2, x3) with respect to the XlX2X3 coordinate system and
(x[, x2, x3) with respect to the X[X'2X3 coordinate system, as shown in Fig. 5.23. Note that, for
convenience, we are using xlt x2, x3 instead of x, y, z.
For simplicity's sake, let us find the relationship between (x[, x'2) and {xh x2). Referring to
Fig. 5.24, x\ is given by

x\ = Oa = Ob + be + ca
= xx cos 6 + ce sin 6 + cP sin 6

= x, cos 6 + (ce + cP) sin 6

Y,
r M
y
p( x,y)
\ x', /)

y'\ i ^
x

\ Figure 5.22 Coordinates of a scalar


o mass M at point P.
Sec. 5.9 Coordinate Transformations 177

/>(*!, X2,
(x\, X2,

Figure 5.23 Coordinates of point P in


two different coordinate systems
rotated with respect to each other are
Oi, x2, x3) and (x[, x2, x'3).

= Xj cos 6 + eP sin d

= x{ cos 6 + x2 sin 6

[it
x, = x, cos 6 + x2 cos ( (5.134)

Similarly, x'2 = x, sin 6 + x2 cos 6

[it \
or x2 = X] cosl 1- 6\ + x2 cos (5.135)

P{xvx2)

Figure 5.24 To calculate (x[,x'2) in


terms of (xu x2).
178 Vector Analysis, Vector Operators, and Transformations Chap. 5

Thus we have been able to express x\ and x'2 in terms of xx, x2 and the cosines of angle 6. The
notation can be simplified by using directional cosines. Let ax be the cosine of the angle be-
tween the Xj-axis and X r axis or between the unit vectors xj and xx; that is,
ax = cos(Xj,X,) = cos(x[, x,) = x[ x{ = cos 6
Similarly,
(IT \
a2 = cos(XJ,X2) = cos(x2, x 2 ) = x 2 x 2 = cosl 61 = sin 6

(v \
/5l = cos(X2,X,) = cos(x 2 ,x,) = x 2 Xj = cosl + 6\ = - s i n 6

j82 = cos(X2, X2) = cos(x2, x 2 ) = x 2 x 2 = cos 6 (5.136)

Thus Eqs. (5.134) and (5.135) may be written as


x'x = axxx + a2x2 (5.137)

x2 = (5.138)

We may extend these equations to a three-dimensional case.


If we deal with a point P in space with coordinates (xh x2, x3) in the X,X2X3 system and
(x[, x2, x'3) in the X'XX'2X'3 system, then
x[ = axxx + a2x2 + a3x3 (5.139)
x2 = p2x2 + p3x3 (5.140)

(5.141)

where yx is the cosine of the angle between x 3 and x,. y2, y3, a3, and (33 have similar meanings.
The reverse transformation, that is, xx, x2, x3 in terms of x[, x'2, x'3, may be written as

xx = (5.142)
x2 = a2x[ + f32x'2 (5.143)

x3 = a3x[ y3x3 (5.144)

where ax is the cosine of the angle between the X r axis and Xj-axis, and the other directional
cosines have similar meaning.
The transformation equations, Eqs. (5.142) to (5.144), may be written in a much neater
and more compact form by using the following notation. Let A,-, be the cosine of the angle be-
tween the X;-axis and X^-axis; that is, the directional cosine Ay is
A^ s cos(X;,X7) = x; Xj (5.145)
Sec. 5.9 Coordinate Transformations 179

and A;7 = c o s ^ , X[) = x;. x,' (5.146)

The coordinates x[, x'2, x'3 may be expressed in terms of xx, x2, x3 as

X^ =
AJJXJ ~r A12^2 13*^3 (5.147)

x2 - X22x2 + A 23 x 3 (5.148)

^3 31*^1 32^2 (5.149)

while the reverse transformation is


X
\ = (5.150)
r = A v ' -I- \ r' - (5.151)
A2 A|2Aj 1 A22A2

x = A x' + A JC' +A JE' (5.152)

Using summation notation, these transformations may be written as


3
r' ^ A v 1 1 9 3 (5.153)
*i / / A^jXp I 1, /,, J

3
r = V A r' 1 = 1 9 3 (5.154)

where A,-,- may be thought of as elements of a 3 by 3 square matrix A defined as

An A,
A = A21 A23 (5.155)
31 A33
\
The matrix A. is called a transformation matrix or a rotation matrix and determines the proper-
ties of the coordinates of a point under transformation.
According to Eq. (5.155), we need nine quantities A,y to cause the coordinate transforma-
tion of a point. But looking further into the properties of A., we find that not all the quantities A,y
are independent. To understand this, we look at two geometrical relations. In Fig. 5.25, the line
OP makes angles 0,, 02, and 03 with the Xr, X2-, and X3-axes, respectively. Hence directional
cosines of the straight line OP are cos 0,, cos 02, and cos 03. As discussed in Section 5.6, the
sum of the squares of the direction cosines of any line is equal to unity; that is (after replacing
a. /3, and yby 6U 62, and 63, respectively)
cos 2 + cos 2 02 + cos 2 03 = 1 (5.156)
Now with reference to Fig. 5.26, if a line OP makes angles 0,, 02, 03 and line OQ makes angles
0[, 02, 03 with the axes ZjX2X3, the cosine of the angle between these lines is given by
cos 0 = cos 0, cos 0J + cos 02 cos 02 + cos 03 cos 03 (5.157)
180 Vector Analysis, Vector Operators, and Transformations Chap. 5

Figure 5.25 The angles 0h 02, 03 that


OP makes with the three axes are used
for calculating directional cosines.

Let us now consider a set of axes XYX2X3. Each of these, when rotated through an angle 6,
results in a new set of axes X\X'2X'y Let us describe the Xj-axis in the XXX2X3 system. Its direc-
tion cosines are A n , A12, A13, while for the X2-axis in the XXX2X3. system, they are A2i, A22, A23.
Since X[ is perpendicular to X'2, the angle 9 is 772; when we apply Eq. (5.157), we get

77
AnA21 + A12A22 + A]3A23 cos 6 = cos 0 (5.158)

which may be written as

(5.159)

In general, applying this to the other two combinations, we get


3
Hv i t = o, if i * j (5.160)
k=l

Figure 5.26 For calculating 6 be-


tween two lines OP and OQ.
Sec. 5.9 Coordinate Transformations 181

Similarly, if we apply Eq. (5.156) to the three axes X[, X'2, X'3 separately described in theX1X2XJ
system, we get

(5.161)

A31 A32 A33


33

"A'hich may be written in compact form as


3
if/=7 (5.162)
k=\

Equations (5.160) and (5.162) are called orthogonality conditions and apply to any set of coor-
jinate systems in which the coordinate axes are mutually perpendicular; that is, the systems are
nhogonal. Equations (5.160) and (5.162) may be combined into one as

(5.163)
k=\

i here Sy is called the Kronecker delta, which has the properties

1, iii=j
(5.164)
0, if i + j
Thus Eq. (5.163) results in six relations between the directional cosines, thereby reducing the
".umber of independent quantities Ay in the matrix A. to only three.
The transformation matrix X described here can be used to describe two different but
.iosely related transformation:

1. Coordinate transformation: In this case, point P is fixed, while the base vectors are trans-
formed (say from Xl X2 to X\X'2), causing the coordinates of point P to change, as shown
in Fig. 5.27(a). This is the interpretation we have explained here.
2. Point transformation: The alternative is to keep the coordinates (or base vectors) fixed and
let point P rotate to point P', as shown in Fig. 5.27(b), always keeping the distance from
the origin constant.

Let us reconsider Fig. 5.27(a) and (b). In Fig. 5.27(a), axes Xx and X2 are fixed and are the
reference axes, while axes X[ and X2 are obtained by a rotation through an angle 6. Thus the co-
ordinates of point P (x[, x'2) in the rotated coordinate system are given in terms of the coordi-
nates (xu x2) in the fixed coordinate system as
= JC, cos 8 + x2 sin 8 (5.165a)
= JC, sin 6 + x, cos 0 (5.165b)
In this case the transformation acts on the axes and is called a coordinate transformation. The
>ame result can be obtained if we keep the axes fixed but rotate point P through an angle 8 (in
182 Vector Analysis, Vector Operators, and Transformations Chap. 5

(a) (b)

Figure 5.27 (a) Coordinate transformation and (b) point transformation.

the direction opposite to that in which the axes were rotated) to P'. The coordinates of P' are
again given by Eqs. (5.165). This type of transformation, which acts on a point, is called a point
transformation. The two types of transformations are completely equivalent.
Finally, a set of quantities A{{Ah A2, A3) in an unprimed system may be transformed to a
primed system by means of a transformation matrix A, resulting in [see Eqs. (5.153)]

Al = 2 KjAj (5.166)
j

The quantities that obey such transformation rules are called vectors; that is, A,{A\, A2, A3) =
A is a vector quantity.
We have considered a transformation matrix A that is a 3 by 3 square matrix; that is, the
number of rows is equal to the number of columns. The matrix may not always be a square. For
example, the coordinates Qt,, x2, x3) of a point may be represented by a column matrix x as

(5.167)

or a row matrix
x = (x. (5.168)
A common practice is to use the column matrix given by Eq. (5.167) for representing a vector,
and we shall use this convention. Thus the coordinates xt(xx, x2, x3) and x[{x\, x'2, x'3) of point P
with respect to the two reference coordinates, the XXX2X3 and X\, X2, X'3 systems, respectively,
may be expressed in matrix representation. Thus the transformation equations given by
Eq. (5.153)

A= i = 1, 2, 3 (5.153)
Problems 183

may be written in matrix notation as


x' = Ax (5.169)
This is equivalent to
x[ 1*1\
\
4 =
/
A21
A31
i A12
A22
A32
A]3
A23
A33
uu / (5.170)

which is the same thing as


xi A,!

(5.171)
X
3 =
A31^ + + A33X3

Note that the multiplication in Eqs. (5.170) or (5.171) is possible only if (1) x and x' are column
matrices, and (2) the number of columns in A. must be equal to the number of rows in x. In gen-
eral, if we want to multiply matrix A with matrix B, the resulting matrix C is given by
C = AB (5.172)
where the number of columns in matrix A must be equal to the number of rows in B. Any ele-
ment Ctj of matrix C is given by

(5.173)

In general, matrix multiplication is not commutative; that is,


(5.174)

PROBLEMS
5.1. Prove the following inequalities:
(a) |A + B| * |A B (b) |A B| =s |A||B (c) A x B N A B
5.2 Find the resultant of three forces F b F2, and F 3 in terms of their magnitudes Fu F2, and F 3 and an-
gles 6U 62, and 03 between each pair of forces. Also find an expression for the angle a between the
resultant force F and the component force F].
5.3. Given the vectors A = (4, - 2 , 6) and B = (1, 3, -4), calculate (a) A + B, (b) A - B, (c) 3A +
2B, (d) 3A - 2B, (e) A, B, A2 + B2, (f) A B, (g) the angle between A and B, (h) A x B, (i) the
component of B in the direction of A, and (j) the directional cosines of A and B.
5.4. Given two vectors A = 2i + 3j + 4k and B = - 2 i - 3j - 4k, calculate (a) A + B, (b) A - B,
(c) 3A + 2B, (d) 3A - 2B, (e) A, B, A2 + B2, (f) A B, (g) the angle between A and B, (h) A x B,
(i) the component of B in the direction of A, and (j) the directional cosines of A and B.
5.5. Find the cosine of the angle between vectors A = 2i + 3 j + 2k and B = 2i - j + 2k\
5.6. Prove that the diagonals of an equilateral parallelogram are perpendicular.
5.7. Find a unit vector n that is perpendicular to vectors A = i + 2j + 3k* and B = 2i j + 2k.
5.8. A = 2! + cj + k is perpendicular to B = 1 + j + 2k. What is the value of c?
184 Vector Analysis, Vector Operators, and Transformations Chap. 5

5.9. Show that vectors A = 1 - 2j - k and B = (M + 8j - 10k" are perpendicular to each other.
5.10. For what values of c will the following two vectors be perpendicular to each other: A = ci + 2cj
- 4k" and B = 2i - cj + 2cft?
5.11. Show that the vector r = 2i + 5j lies in a plane perpendicular to the OZ-axis.
5.12. Vectors A and B represent the adjacent sides of a parallelogram. Show that the area of a parallel-
ogram is equal to |A X B .
5.13. A = i + j and B = i + j + k are vectors that represent diagonals along the face and through a
cube, respectively. Calculate the angle between these two vectors.
5.14. Let n be a unit vector in some fixed direction and A an arbitrary vector. Show that A = (A n )n
+ (n x A) x n.
5.15. For vectors A, B, and C as given, calculate the following quantities:

A = i + 2j + 3k, B = 3t - 2] + k, C =I +J - k

(a) A + B + C (b) A - B + C (c) A - B - C (d) (A + B) C


(e) A (B + C) (f) A (B x C) (g) (A X B) C (h) A x B x C
5.16. Using the fundamental definition of vector diffentiation, prove the results stated in Eqs. (5.82),
(5.83), and (5.85).
5.17. For vectors A and B as given, calculate the following:
A = ae-k'\ + bt] + k and B = (c sin wr) i + (d cos

t ^ d A dB dk dB d
(a) (b) B B
dt dt dt dt dt dt
d d
(c) 0 (d) (A x B)
dt
dt
5.18. Show that a triple scalar product can be written in determinant form as follows:

Ay
(AxB)-C= By
c
5.19. Prove the following cyclic relation for a triple scalar product:
A (B X C) = B (C X A) = C (A x B) = (A x B) C
5.20. Prove the following identity:
A X B X C = (A C)B - (A B)C = B(A C) - C(A B)
5.21. Using the properties of the del operator, V, prove the following vector identity: grad(i;) =
u grad v + v grad u.
5.22. Using the properties of the del operator, V, prove the following vector identity: curl (curl A) =
grad (div A) - V2A.
5.23. Calculate grad S, where S = 1/r3 and r = (x1 + y2 + z2)1'2.
5.24. Find the gradient of the following functions:
(a)/=x + y + z (b) f=xy + xz + yz (c) / = xy2 + yx2 + xyz
5.25. The potential that represents an inverse-square force is V(r) = klr, where r2 = x2 + y2 + z2. Using
the definition F = W, calculate the components of this force.
Problems 185

5.26. For Problem 5.25, find a unit vector that points in the direction of the maximum increase in V at
the position r = i + 2j + k.
5.27. For Problem 5.23, find a unit vector that points in the direction of the maximum increase in S at
the position r = 1 + 2j + k.
5.28. Calculate the divergence of r, div r = V r, where:
(a) r = xl + y] (b) r = xi + yj + zfi
(c) r = yl - x] - z (d) r = 4x1 + 2j + 4yic
5.29. Calculate the divergence of the following vector fields:
(a) r = A + y2] (b) r = x\ - y2j - (c) r = xyi yzj + zxfi
5.30. The gravitational force between two masses may be written as

Calculate the divergence of F.


5.31. Fluid flowing perpendicular to the six surfaces of a cube of side 10 cm has the magnitude vx =
10 cm/s, v_x = 20 cm/s, u^ = 20 cm/s, v_y = 10 cm/s, vz = 50 cm/s, and i>_z = 50 cm/s. Cal-
culate the net flud flowing out of the box. Is it a source or sink?
5.32. A cube of side 5 cm is placed in a fluid. For the following velocity values, calculate the average di-
vergence of the flow. Is there a source or sink within this box? The velocities are given in cm/s.

Front vF = 30l + 60] -30fc


Right VR = 20i - 20] + 20k"
Left vL = - lOOl + 200] + 300
Top vT = 5 1 + 5 0 ] + 10k"
Bottom vBl, = o
Back V
B S, = 250
Also draw a diagram showing these velocities.
5.33. A box of sides 15 cm by 10 cm by 10 cm is placed influid.Theflowrates are as shown in Fig. P5.33.
Calculate the net flow. Is there a source or sink within the box?

75 cm/s

100 cm/s
Figure P5.33
186 Vector Analysis, Vector Operators, and Transformations Chap. 5

5.34. Evaluate the curl of the following vector fields:


(a) r = xl + yj (b) r = yi + xj + zk
(c) r = A + / I (d) r = A - fl - ^k
(e) r = xy\ + yz\ + zxfi (f) Axi - 2j + 4yk
5.35. Evaluate V X F for the following forces:
(a) F = Fx\ + Fy] + Fzk
(b)F = (4abyz2 - l0bx2y2)\ + (9abxz2 - 6to 3 j)j + Sabxyzk
(c) Fx = 6abyz3 - 20bx3y2
3
Fy = 6abxz -
Fz = ISabxyz2
5.36 Using Stokes' theorem, calculate the average value of the curl of the fluid for a rectangular path
20 cm by 10 cm, as shown in Fig. P5.36.

Y i c
- 4 cm/s

O <D >- 8 cm/s

12 cm/s

Figure P5.36

5.37 Repeat Problem 5.36 if the velocities (in cm/s) at points A, B, C, and D are

vA = 10t + 5j, vB = 5l + 10J


v c = 51 + 10], vD = l(ff + 5]

5.38. Using Stokes' theorem, calculate the average value of the curl of thefluidfor a square of side 10 cm,
as shown in Fig. P5.38.

40 cm/s

50 cm/s

Figure P5.38

5.39. Prove the following relation from trigonometric considerations: cos2 a + cos2 /3 + cos2 y = 1.
5.40. A plane formed by the XY-axes is rotated through an angle of 30 about the Z-axis. Find the matrix
of rotation.
Problems 187

5.41. A cube is rotated about the Z-axis through 60. Find the rotation matrix for the coordinate
transformation.
5.42. Consider X. to be a two-dimensional transformation matrix. Prove by direct expansion that | \ ] 2 = 1.
5.43. Find the transformation matrix that will cause the rotation of the rectangular coordinates XY sys-
tem through an angle of 120 about the Z-axis.
5.44. Find the transformation matrix that will cause the rotation of the rectangular coordinates XY sys-
tem through an angle of 120 about an axis that makes equal angles with the original three rectan-
gular coordinate axes.

SUGGESTIONS FOR FURTHER READING


ARFKEN, G., Mathematical Methods for Physicists. New York: Academic Press, Inc., 1968.
ARTHUR, W., and FENSTER, S. K., Mechanics, Chapter 1. New York: Holt, Rinehart and Winston, Inc.,
1969.
BECKER, R. A., Introduction to Theoretical Mechanics, Chapter 1. New York: McGraw-Hill Book Co, 1954.
DAVIS, A. DOUGLAS, Classical Mechanics, Chapter 4. New York: Academic Press, Inc., 1986.
DAVIS, H. F., Introduction to Vector Analysis. Needham Heights, Mass.: Allyn & Bacon, 1961.
EISENMAN, R. L., Matrix Vector Analysis. New York: McGraw-Hill Book Co., 1963.
FOWLES, G. R., Analytical Mechanics, Chapters 1 and 2. New York: Holt, Rinehart and Winston, Inc., 1962.
HAUSER, W., Introduction to the Principles of Mechanics, Chapter 1. Reading, Mass.: Addison-Wesley
Publishing Co., 1965.
KJTTEL, C , KNIGHT, W. D., and RUDERMAN, M. A., Mechanics, Berkeley Physics Course, Volume 1, Chap-
ter 2. New York: McGraw-Hill Book Co., 1965.
KLEPPNER, D., and KOLENKOW, R. J., An Introduction to Mechanics, Chapters 1 and 5. New York:
McGraw-Hill Book Co., 1973.
LINDGREN, B. W., Vector Calculus. New York: Macmillan, Inc., 1964.
MARION, J. B., Classical Dynamics, 2nd ed., Chapter 1. New York: Academic Press, Inc., 1970.
MARION, J. B., Principles of Vector Analysis. New York: Academic Press, Inc., 1965.
ROSSBERG, K., Analytical Mechanics, Chapter 1. New York: John Wiley & Sons, Inc., 1983.
SCHEY, H. M., Div, Grad, Curl and All That. New York: W. W. Norton and Co., Inc., 1973.
SVMON, K. R., Mechanics, 3rd ed., Chapter 3. Reading, Mass.: Addison-Wesley Publishing Co., 1971.
TAYLOR, E. F., Introductory Mechanics, Chapter 4. New York: John Wiley & Sons, Inc., 1963.
WYLIE, C. R., Jr., Advanced Engineering Mathematics. New York: McGraw-Hill Book Co., 1960.

Das könnte Ihnen auch gefallen