appendix E

Vector and Tensor Analysis

1 INTRODUCTION

The equations of applied mathematics express relations between quantities that are capable of measurement in terms of certain defined units. The simplest types of physical quantities are completely defined by a certain simple number. Examples of such quantities are mass, temperature, length. Such quantities are called scalars.

There are, however, other physical quantities that are not scalars. Such quantities as the displacement of a point, the velocity of a particle, a mechanical force require three numbers to specify them completely. The three numbers that are required to specify the quantities are scalars and could be, for example, the components of the displacement of the particle with respect to an arbitrary cartesian coordinate reference frame.

We could carry out the various mathematical operations with these scalar quantities, but we would be neglecting the fact that from a physical point of view a displacement, for example, is one entity, and also we are introducing a foreign element into the question, that is, the coordinate system.

Accordingly it has been found convenient to introduce a mathematical discipline that enables us to study quantities of this type without recourse to a definite coordinate system.

It is only when we come to evaluate formulas numerically that it will be necessary to introduce a definite coordinate system. The mathematical technique that enables us to do this is vector analysis.

2 THE CONCEPT OF A VECTOR

A physical quantity possessing both magnitude (length) and direction is called a vector. Typical examples are force, velocity, acceleration, momentum. It is customary to represent vectors by letters in boldface type and scalars in lightface italics.

A vector may be indicated graphically by an arrow drawn between two points. It is thus evident that rectilinear displacements of a point and all physical quantities that can be represented by such displacements in the same manner that a scalar can be represented by the points of a straight line are vectors.

3 ADDITION AND SUBTRACTION OF VECTORS; MULTIPLICATION OF A VECTOR BY A SCALAR

A vector having been defined as an entity that behaves in the same manner as the rectilinear displacement of a point, vector addition is reduced to a composition of linear displacements.

Consider two vectors A and B, as shown in Fig. 3.1. The vector C, which is obtained by moving a point along A and then along B, is called the resultant, or sum, of the vectors A and B, and we write

images

From the nature of the definition of vector addition it is apparent that

images

and that therefore vector addition is commutative. If the vectors A and B are situated as shown in Fig. 3.2, then the resultant vector C is obtained by completing the parallelogram formed by the two vectors.

images

Fig. 3.1

images

Fig. 3.2

In general, the sum of the vectors A and B is obtained by placing the origin of B at the terminus of A. Then the vector C extending from the origin of A to the terminus of B is defined as the sum, or resultant, of A and B.

To add several vectors A, B, and C, first find the sum of A and B and the sum of A + B and C. It is seen that the result is the same as finding the sum of B and C first and then the sum of A and B + C. Thus vector addition is associative.

images

To subtract B from A, add –B to A, as shown in Fig. 3.3.

EQUALITY OF VECTORS

Two vectors are said to be equal if they have the same magnitude and the same direction.

MULTIPLICATION OF A VECTOR BY A SCALAR

By the product of a vector a by a scalar n we understand a vector whose magnitude is equal to the magnitude of the product of the magnitudes of a and n and has the same direction as a or the opposite direction, depending on whether the scalar n is positive or negative. We thus write

images

to denote this new vector.

UNIT VECTORS

A vector having unit magnitude is called a unit vector. The most common unit vectors are those which have the directions of a right-handed cartesian coordinate system as shown in Fig. 3.4. The vector i is a unit vector having the x direction of the coordinate system; the unit vector j has the y direction; and the unit vector k has the z direction.

THE COMPONENTS OF A VECTOR

The components of a vector A are any vectors whose sum is A. The components most frequently used are those parallel to the axes x, y, and z. These

images

Fig. 3.3

images

Fig. 3.4

are called the rectangular components of the vector. If Ax, Ay, and Az are the projections of A on the axes x, y, z, respectively, we may write

images

If a vector A is given in magnitude and direction, then its components along the three axes of a cartesian reference frame are given by

images

where |A| denotes the magnitude of the vector. If, conversely, the three components of the vector are assigned, the vector A is uniquely specified as the diagonal of the rectangular parallelepiped whose edges are the vectors iAx, jAy and kAz. Its magnitude is

images

and its direction is given by the three cosines which can be found from (3.5) and (3.6).

4 THE SCALAR PRODUCT OF TWO VECTORS

The scalar, or dot, product of two vectors is defined as a scalar quantity equal in magnitude to the product of the magnitudes of the two given vectors and the cosine of the angle between them. The scalar product of the two vectors A and B is thus given by the equation

images

The cosine of the angle between the directions of the two vectors becomes +1 when the directions are the same, –1 when they are opposite, and 0 when they are perpendicular.

It is clear from the definition of the dot product that the commutative law of multiplication holds, that is,

images

The fact that the distributive law of multiplication holds can be seen with the aid of Fig. 4.1. We have from the figure

images

From the fundamental unit vectors we can form the following scalar products:

images

images

Fig. 4.1

As a consequence of these results, we obtain

images

We may also write

images

5 THE VECTOR PRODUCT OF TWO VECTORS

The vector, or cross, product of two vectors is defined to be a vector perpendicular to the plane of the two given vectors in the sense of advance of a right-handed screw rotated from the first to the second of the given vectors through the smaller angle between their positive directions.

The meaning of this definition is made clear in Fig. 5.1. The magnitude of this vector is equal to the product of the magnitudes of the two given vectors times the sine of the angle between them.

The vector, or cross, product is denoted by A × B. As is clear from the definition and the figure, the commutative law does not hold for this type of multiplication; instead we have

images

Also we have

images

It follows from the definition that if the vector product of two vectors vanishes, the vectors are parallel provided neither vector is zero.

images

Fig. 5.1

VECTOR REPRESENTATION OF SURFACES

Let us consider a plane surface such as shown in Fig. 5.2. Since this surface has a magnitude represented by its area and a direction specified by its normal, it is a vector quantity. A certain ambiguity exists as to the positive sense of the normal. In order to remove this ambiguity, the following conventions are adopted:

If the surface is part of a closed surface, the outward-drawn normal is taken as positive.

If the surface is not part of a closed surface, the positive sense in describing the periphery is connected with the positive direction of the normal by the rule that a right-handed screw rotated in the plane of the surface in the positive sense of describing the periphery advances along the positive normal. For example, in Fig. 5.2, if the periphery of the surface is described in the sense ABC, the positive sense of the normal, and therefore the direction of the vector S representing the surface, is upward.

If a surface is not plane, it may be divided into a number of elementary surfaces each of which is plane to any desired degree of approximation. In this case the vector representative of the entire surface is the sum of the vectors representing its elements. Two surfaces, considered as vectors, are equal if the representative vectors are equal. Therefore two plane surfaces are equal if they have equal areas and are normal in the same direction even if they have different shapes.

A curved surface may be replaced by a plane surface perpendicular to its representative vector having an area equal to the magnitude of this vector. The vector representing a closed surface is zero because the projection of the entire surface on any plane is zero; since as much of the projected area is negative as positive, it therefore follows that the vector representing the entire surface has zero components along the three axes x, y, z, and consequently it equals zero.

THE DISTRIBUTIVE LAW OF VECTOR MULTIPLICATION

To prove that the distributive law of multiplication holds for the vector product, consider the prism of Fig. 5.3. The edges of this prism are the vectors P, Q,

images

Fig. 5.2

images

Fig. 5.3

P + Q, and R. The vectors representing the faces of the closed prism are

images

Therefore the representative vector for the entire polyhedral surface is

images

or

images

By the definition of the vector product we have the following relations for the unit vectors:

images

If we write the vectors P and Q in terms of their rectangular components,

images

and realize that the distributive law holds for a vector product, we obtain

images

in view of the properties of the unit vectors expressed by (5.5). This expression can be represented in a compact fashion by the determinant

images

6 MULTIPLE PRODUCTS

There are several types of multiple products used in vector analysis, and in this section the most important types will be considered.

a THE PRODUCT OF A VECTOR AND THE SCALAR PRODUCT OF TWO OTHER VECTORS, A(B·C)

Here B·C is a scalar so that A(B·C) is a vector parallel to A and is of course an entirely different vector from (A·B)C.

b SCALAR PRODUCT OF A VECTOR AND THE VECTOR PRODUCT OF TWO OTHER VECTORS

Consider A·B × C. In this case we have the important relation

images

The proof of this relation follows from the fact that each of these expressions represents the volume of the parallelepiped whose edges are A, B, C. Furthermore, all three expressions give this volume with the positive sign provided the vectors A, B, C, in that order, form a right-handed system.

c VECTOR PRODUCT OF A VECTOR AND THE VECTOR PRODUCT OF TWO OTHER VECTORS

Consider the product

images

By the definition of a cross product q is perpendicular to the vector a and also to the vector b × c. Accordingly the vector q lies in the plane of b and c and may be expressed in the form

images

where u and υ are scalar multipliers.

Now we have

images

or

images

Hence

images

Now let

images

where n is some scalar. To find the magnitude of n, consider a set of cartesian coordinate axes oriented so that the vectors a, b, and c have the following components in terms of these axes:

images

In this case we have

images

images

Hence, comparing (6.9) and (6.10), it is seen that the constant n is equal to 1.

7 DIFFERENTIATION OF A VECTOR WITH RESPECT TO TIME

The differential coefficient or derivative of a vector A with respect to a scalar variable t, say, the time, is defined as a limit by the equation

images

Since division by a scalar does not alter vectorial properties, the derivative of a vector with respect to a scalar variable is itself a vector.

As an example of differentiation of a vector with respect to a scalar, consider a particle p moving along a curve c, as shown in Fig. 7.1. Let the position of the particle with respect to the origin of a cartesian reference frame be denoted by r. In that case the velocity of the particle is given by

images

where ds is a differential of arc measured along the curve, as shown in Fig. 7.2. Now

images

images

Fig. 7.1

images

Fig. 7.2

where T is a unit vector tangent to the curve defining the path of the particle. Hence we have

images

The quantity ds/dt is the speed of the particle. The acceleration of the particle is defined as the time derivative of the velocity. We thus have

images

Now

images

Consider Fig. 7.3.

We may write

images

images

images

Substituting these expressions into (7.6), we obtain

images

Substituting this into (7.5), we finally obtain

images

We thus see that the acceleration of the particle consists of two terms. The first term depends upon the rate of change of speed of the particle and is directed along the tangent to the particle; the second term is the centripetal

images

Fig. 7.3

acceleration of the particle and depends on the radius of curvature ρ of the particle and the square of the speed. This acceleration is directed in a direction normal to the curve and toward the center of curvature.

Since the derivatives of vectors with respect to a scalar variable are deduced by a limiting process from subtraction of vectors and division by scalars, which are operations subject to the rules of ordinary algebra, it follows that the rules of the differential calculus can be extended at once to the differentiation of a sum of vectors,

images

or the product of a scalar and a vector,

images

And, similarly,

images

images

APPLICATION TO MECHANICS—CORIOLIS’ ACCELERATION

An interesting application of vector differentiation of importance to mechanics will now be discussed. Consider a small insect on a rotating plane sheet of cardboard. The cardboard will be assumed to be rotating about a fixed point O with a uniform angular evlocity w. It will be supposed that the insect may move with respect to the cardboard. In order to describe the velocity and acceleration of the moving insect, let a cartesian coordinate system x, y be drawn on the rotating cardboard with the center of rotation as the origin. Let the position vector of the insect be

images

where i and j are unit vectors in the x and y directions that rotate with the cardboard. The velocity of the insect v relative to fixed space is

images

Now, since i and j are unit vectors rotating with angular velocity w, it is easy to show by the reasoning of Eqs. (7.6) and (7.10) that

images

Hence (7.17) may be written in the form

images

where the dots indicate differentiation with respect to the time in the usual manner. The acceleration a of the insect is given by

images

Equation (7.20) indicates that the acceleration of the insect consists of three separate terms of the form

images

where

images

a1 is the acceleration of the insect with respect to the rotating coordinate system. If the angular velocity w were zero and the card were stationary, this would be the only acceleration. a2 is the so-called Coriolis’ acceleration. It depends on the angular velocity w of the rotating card and on the velocity of the particle relative to the rotating coordinate axis. This relative velocity vr is given by

images

If the scalar product of the relative velocity vr and the Coriolis’ acceleration a2 is taken, the following result is obtained:

images

This indicates that the Coriolis’ acceleration is perpendicular to the relative velocity.

The acceleration a3 is the centripetal acceleration. Its magnitude is proportional to the square of the angular velocity w of the frame of reference and to the distance of the particle from the center of rotation. It is directed radially toward the center of rotation. Equation (7.20) may be generalized, and the acceleration of a particle in a rotating reference plane such as the earth can be computed. The Coriolis’ acceleration causes deviation of falling bodies and projectiles and explains the motion of Foucault’s pendulum.

8 THE GRADIENT

Let ϕ(x,y,z) be a scalar function of position in space, that is, of the coordinates x, y, z. If the coordinates x, y, z are increased by dx, dy, dz, respectively, we have

images

If we denote by dr the vector representing the displacement specified by dx, dy, dz, then

images

In vector analysis a certain vector differential operator ∇ (read “del”) defined by

images

plays a very prominent role. The gradient of a scalar function ϕ(x,y,z) is defined by

images

Operating with ∇ on the scalar function ϕ(x,y,z), we get

images

This is just the expression (8.4) defined as the gradient of ϕ. Now, from (8.1) and (8.2), we have

images

The equation

images

represents a certain surface, and as we change the value of the constant, we obtain a family of surfaces.

Consider the surfaces of Fig. 8.1. If dn denotes the distance along the normal from the point P to the surface S2, we may write

images

where n is the unit normal to the surface S1 at P. We have

images

images

Fig. 8.1

and, in particular, if dr lies in the surface S1 we have

images

showing that the vector ∇ϕ is normal to the surface ϕ = const. Since the vector dr is arbitrary, we have from (8.9)

images

Hence ∇ϕ is a vector whose magnitude is equal to the maximum rate of change of ϕ with respect to the space variables and has the direction of that change.

9 THE DIVERGENCE AND GAUSS’S THEOREM

The scalar product of the vector operator ∇ and a vector A gives a scalar that is called the divergence of A; that is,

images

This quantity has an important application in hydrodynamics.

Consider a fluid of density ρ(x,y,z,t) and velocity v = v(x,y,z,t), and let

images

If S is the representative vector of the area of a plane surface, then V·S is the mass of fluid flowing through the surface S in a unit time.

Consider a small fixed rectangular parallelepiped of dimensions dx, dy, dz, as shown in Fig. 9.1. The mass of fluid flowing in through face F1 per unit time is

images

and that flowing out through face F2 is

images

Hence the net increase of mass of fluid inside the parallelepiped per unit time is

images

images

Fig. 9.1

Considering the net increase of mass of fluid per unit time entering through the other two pairs of faces, we obtain

images

as the total increase in mass in the parallelepiped per unit time.

But by the principle of conservation of matter, this must be equal to the time rate of increase of density multiplied by the volume of the parallelepiped. Hence

images

or

images

This is known in hydrodynamics as the equation of continuity. If the fluid is incompressible, then

images

The name divergence originated in this interpretation of ∇·V. For since – ∇·V represents the excess of the inward over the outward flow, or the convergence of the fluid, so ∇·V represents the excess of the outward over the inward flow, or the divergence of the fluid.

SURFACE INTEGRAL

Consider a surface s as shown in Fig. 9.2. Divide the surface into the representative vectors ds1, ds2, . . . , etc. Let Vi be the value of the vector function of position Vi(x,y,z) at dsi. Then

images

images

Fig. 9.2

is known as the surface integral of V over the surface s. The sign of the integral depends on which face of the surface is taken as positive. If the surface is closed, the outward normal is taken as positive.

Since

images

we have

images

The surface integral of a vector V is called the flux of V throughout the surface.

GAUSS’S THEOREM

This is one of the most important theorems of vector analysis. It states that the volume integral of the divergence of a vector field A taken over any volume V is equal to the surface integral of A taken over the closed surface surrounding the volume V; that is,

images

To prove Gauss’s theorem, let us expand the left-hand side of Eq. (9.13). We then have

images

Let us consider the first integral on the right. Integrate with respect to x, that is, along a strip of cross section dydz extending from P1 to P2 of Fig. 9.3. We thus obtain

images

images

Fig. 9.3

Here (x1, y, z) are the coordinates of P1 and (x2y,z) are the coordinates of P2. Now at P1 we have

images

At P2 we have

images

Therefore

images

where the surface integral on the right is evaluated over the entire surface. In the same manner we obtain

images

images

If we now add (9.18), (9.19), and (9.20), we obtain Gauss’s theorem:

images

GREEN’S THEOREM

By the use of Gauss’s theorem we are able to make some important transformations. Consider

images

that is, let the vector field A be the product of a scalar function u and the gradient of another scalar function w. Consider

images

If we place this value of ∇·A in the left-hand side of Gauss’s theorem (9.21), we obtain

images

This transformation is referred to as the first form of Green’s theorem.

In (9.24) if we interchange the functions u and w, we obtain

images

If we now subtract Eq. (9.25) from Eq. (9.24), we obtain

images

This transformation is referred to as the second form of Green’s theorem. The transformations (9.24) and (9.26) are of extreme importance in the fields of electrodynamics and hydrodynamics.

10 THE CURL OF A VECTOR FIELD AND STOKES’S THEOREM THE LINE INTEGRAL

Let A be a vector field in space and AB (Fig. 10.1) a curve described in the sense A to B. Let the continuous curve AB be subdivided into infinitesimal vector elements dI1, dI2,. . . , dIn. Take the scalar products A1·dI1,

images

Fig. 10.1

A2·dI2, . . . , An·dIn, where A1, A2, . . . , An are the values that the vector field A takes at the junction points of the vectors dI1, dI2, . . . , dIn. The sum of these scalar products, that is,

images

summed up along the entire length of the curve, is known as the line integral of A along the curve AB. It is obvious that the line integral from B to A is the negative of that from A to B.

In terms of cartesian components we can write

images

If F represents the force on a moving particle, then the line integral of F over the path described by the particle is the work done by the force.

Let A be the gradient ∇ϕ of a scalar function of position; that is,

images

Then

images

But we have

images

the total differential of ϕ. We thus have

images

where ϕB and ϕA are the values of ϕ at the points B and A, respectively. It follows from this that the line integral of the gradient of any scalar function of position ϕ around a closed curve vanishes, because if the curve is closed, the points A and B are coincident and the line integral is equal to ϕAϕA, which is zero. The line integral around a closed curve is denoted by an integral sign with a circle around it, as follows:

images

Let us suppose that the line integral of A vanishes about every closed path in space. If we denote the path of integration by a subscript under the integral sign (Fig. 10.2),

images

images

Fig. 10.2

and therefore

images

This shows that the line integral of A from A to B is independent of the path followed. It is apparent, therefore, that it can depend only upon the initial point A and the final point B of the path, that is,

images

Now if we take the two points A and B very close together, we have

images

or

images

As this is true for all directions, the vector A – ∇ϕ can have no component in any direction, and hence it must vanish. Therefore

images

That is, if the line integral of A vanishes about every closed path, A must be the gradient of some scalar function ϕ.

THE CURL OF A VECTOR FIELD

If A is a vector field, the curl, or rotation, of A is defined as the vector function of space obtained by taking the vector product of the operator ∇ and A. That is,

curl A = ∇ × A

images

This may be written conveniently in the following determinantal form:

images

If A = ∇ϕ,

images

It thus follows that if A is the gradient of a scalar, the curl of A vanishes.

LINE INTEGRAL IN A PLANE

To show the connection between the line integral and the curl of a vector field, let us compute the line integral of a vector field A around an infinitesimal rectangle of side Δx and Δy lying in the xy plane, as shown in Fig. 10.3. That is, we shall compute dI around this rectangle. We may write down the various contributions to this integral as follows:

images

where we have made use of the fact that Δx and Δy are infinitesimals. Adding the various contributions, we obtain

images

In view of (10.13), this may be written in the form

images

where ∇× Az is the z component of the curl of A and dsxy is the area of the rectangle ABCD.

Consider now a closed curve in the xy plane, as shown in Fig. 10.4. Divide the space inside C by a network of lines joining a network of infinitesimal rectangles.

images

Fig. 10.3

images

Fig. 10.4

If we take the sum of the line integrals around the various meshes, we obtain

images

Now it is easily seen that the contributions to the line integrals of adjoining meshes neutralize each other because they are traversed in opposite directions; the only contributions which are not neutralized are those on the periphery of the surface. Hence

images

where the line integral on the right is taken along the boundary curve C in the positive sense.

Now the sum on the right of (10.18) reduces to the following integral:

images

Hence, substituting this into (10.18), we obtain

images

That is, the line integral of a vector field A about the contour C of a plane surface S is equal to the surface integral of the normal component of the curl of A to the surface throughout the surface s.

Consider now the triangular surface of Fig. 10.5. It is easy to see that

images

images

Fig. 10.5

since the contribution along the lines OA, OB, and OC cancel each other. However, as a consequence of (10.21), we have

images

If we add these equations and make use of (10.22), we obtain

images

Now we may write

images

for the projections of the representative surface vector s of the plane ABC to the yz, xz, and xy planes. Using this notation, (10.24) becomes

images

Consider now the open surface of Fig. 10.6. We can regard the surface s of this open surface as being made up of an infinite number of elementary triangular surfaces. If we label r the typical triangle, we have from (10.26)

images

Now the sum of the line integrals of the elementary triangles reduces to a line integration about the periphery C of the closed surface S since the

images

Fig. 10.6

line integrations along adjacent triangles are described in opposite senses and hence cancel. We therefore have

images

In the limit the summation on the right of (10.27) reduces to an integral, and we have

images

As a consequence of (10.19), we then have

images

This relation is known as Stokes’s theorem. It states that the surface integral of the curl of a vector field A taken over any surface S is equal to the line integral of A around the periphery of the surface.

If the surface to which Stokes’s theorem is applied is a closed surface, the length of the periphery is zero and then

images

By the use of Stokes’s theorem we see that, if ∇ × A = 0 everywhere, A is the gradient of a scalar function, because if ∇ × A = 0, then the line integral of A around any closed curve vanishes. This is just the condition that A should be the gradient of a scalar function.

11 SUCCESSIVE APPLICATIONS OF THE OPERATOR ∇

It frequently happens in various applications of vector analysis that we must operate successively with the operator ∇. For example, since the curl of a vector field A, ∇ × A, is a vector field B, it is possible to take the curl of B, that is,

images

If we expand this equation in terms of the cartesian coordinates of A.we obtain

images

In the same manner the following vector identities may be established by expanding V and the other vectors concerned in terms of their components:

images

images

images

images

images

images

images

In the above set of equations the symbol (A ∇)B stands for the vector

images

The above formulas are very useful in applications of vector analysis to various branches of engineering and physics.

12 ORTHOGONAL CURVILINEAR COORDINATES

Many calculations in applied mathematics can be simplified by choosing instead of a cartesian coordinate system another kind of system that takes advantage of the relations of symmetry involved in the particular problem under consideration.

Let these new coordinates be denoted by u1, u2, u3. These are defined by specifying the cartesian coordinates x, y, z as functions of u1, u2, u3, as follows:

images

We shall confine ourselves to the case when the three families of surfaces u1 = const, u2 = const, u3 = const are orthogonal to one another. In this case the line element ds is given by

images

where h1, h2, h3 may be functions of u1, u2, u3. We shall also adopt the convention that the new coordinate system shall be right-handed like the old.

Consider now the infinitesimal parallelepiped whose diagonal is the line element ds and whose faces coincide with the planes u1 or u2 or u3 = const (see Fig. 12.1).

The lengths of its edges are h1 du1, h2 du2, h3du3, and its volume is h1 h2 h3 du1du2du3 Furthermore let ϕ(u1, u2, u3) be a scalar function and A be a vector field with components A1, A2, A3 in the three directions in which the coordinates u1 u2, u3 increase. The u1 component of the gradient of ϕ we can compute at once since by definition

images

We also have similar relations for the directions 2 and 3.

In order to calculate the divergence of a vector field A, we use Gauss’s theorem:

images

The contribution to the integral images through the area OBHC, taken in the direction of the outward normal, is —A1h2h3du2d3, while that through the area AFGJ is

images

From these and the corresponding expressions for the other two pairs of surfaces, we have by (12.4)

images

images

Fig. 12.1

We thus obtain

images

If A = ∇ϕ,

images

The components of the curl of A may be found by Stokes’s theorem,

images

For example, the component 1 of the curl of A is obtained by applying Stokes’s theorem to the surface OBHC. We calculate

images

By Stokes’s theorem this equals the 1 component of the curl of A, (∇ × A)1, multiplied by the area of the face OBHC. That is,

images

Hence

images

By a cyclic change of the indices we obtain

images

images

If we introduce unit vectors i1, i2, i3 along the directions 1, 2, and 3, we may write symbolically

images

In the case of cartesian coordinates we have u1 = x, u2= y, u3 = z, h1= h2= h3 = 1, and i1 = i, i2 = j, j3 = k. In this case (12.14) reduces to (10.14). We shall now apply these general formulas to two special cases that are particularly important in applications.

a CYLINDRICAL COORDINATES

The position of a point in space may be determined by the cylindrical coordinate system of Fig. 12.2. In this case we have

images

images

We have, therefore, in this case,

images

By (12.3) we obtain

images

By (12.6) we have

images

images

Fig. 12.2

The Laplacian operator as given by (12.7) gives

images

We obtain the components of the curl of A by (12.11), (12.12), and (12.13):

images

images

images

b SPHERICAL POLAR COORDINATES

Another very important coordinate system is that of spherical polar coordinates, as shown in Fig. 12.3. In this case we have

images

We have therefore

images

images

images

images

Fig. 12.3

images

images

13 APPLICATION TO HYDRODYNAMICS

Consider a region of space containing a fluid of density ρ(x,y,z,t). Let V be the volume inside an arbitrary closed surface s located in this region. Let Q(t) be the mass of fluid inside the volume Fat any instant. Then

images

If υ denotes the velocity of a typical particle of the fluid, the rate at which the mass of fluid inside V is increasing is

images

Now, if we differentiate (13.1) with respect to time and equate the result to (13.2), we have

images

But, by Gauss’s theorem, we have

images

Substituting this into (13.3) and transposing, we obtain

images

Now since the integrand is continuous and the volume V is arbitrary, we conclude that

images

This is the basic equation of hydrodynamics and is known as the equation of continuity.

If the fluid is incompressible, then ρ is a constant and we have

images

If the flow is irrotational, then we have

images

and we know from Sec. 10 that there exists a scalar function ϕ such that

images

If the fluid is incompressible, then from (13.7) we have

images

and hence ϕ, called the velocity potential, satisfies the equation

images

On a fixed boundary of the fluid, the velocity has no normal component, and if ∂/∂n denotes differentiation with respect to the normal direction of the boundary, we must have

images

as a consequence of (13.9) .

Let us denote any general property of a particle of the fluid, such as its pressure, density, etc., by the function H(x,y,z,t). Then by ∂H/∂t is meant the variation of H at a particular point in space as a function of the time t. If we take the total differential of H(x,y,z,t), we obtain

images

images

The quantity dH/dt is the variation of H when we fix our attention on the same particle of fluid.

The velocity of the particle is given by

images

By definition we have

images

Hence

images

images

EULER’S EQUATION OF MOTION

To obtain the equation of motion of a frictionless fluid, we consider the forces acting on an element of fluid whose volume is dx dy dz and whose mass is ρ dx dy dz. As a consequence of the pressure of the fluid, p, there will be a force in the x direction on the element of fluid under consideration, of magnitude

images

Let us also consider the action of an external force F acting on unit mass of fluid. The external force acting on the element under consideration in the x direction is Fxρ dx dy dz. The acceleration of the element in the x direction is images Hence, by Newton’s second law of motion, we have

images

By this reasoning, and considering the y and z directions, we obtain Euler’s equations of motion:

images

These three scalar equations may be combined into the single vector equation

images

As a consequence of Eq. (13.18), we have

images

Now by the vector identity (11.6), we have

images

Hence, by the use of the two equations above, we may write Euler’s equation of motion in the following form:

images

If the external force is conservative, it has a potential V and we have

images

If the motion of the fluid is irrotational, ∇ × v = 0 and v = ∇ϕ. In this case Eq. (13.25) becomes

images

or

images

where

images

Now if dr denotes an arbitrary path in the fluid at any instant, we have

images

Let us form the scalar product of ∇w and dr We then obtain

images

as a consequence of Eq. (13.28). Hence

images

where β(t) is an arbitrary function of time, since we have integrated along an arbitrary path in the fluid at any instant. That is, we have

images

Equation (13.32) is known as Bernoulli’s equation.

In the special case where p is constant and the motion is steady so that ∂ϕ/ϕt = 0, this equation takes the form

images

where β is now a constant. This equation states that, per unit mass of fluid, the sum of the kinetic energy, the potential energy V, and the pressure energy p/ρ has a constant value β for all points of the fluid.

Usually in aerodynamics the variations in V are so small that they can be neglected so that we can write

images

where p0 is the pressure when the fluid is at rest. It can be seen from this that in this case the pressure diminishes as the velocity increases in the form

images

Airplanes are equipped with instruments for measuring p and p0 so that the velocity of the machine relative to the air can be determined by the equation

images

14 THE EQUATION OF HEAT FLOW IN SOLIDS

Consider a region inside a solid body such as a large block of metal. Let a closed surface s be situated inside this region. Let the volume inside the surface s be denoted by V.

In suitable units the amount of heat H inside the volume V of this body is given by

images

where u = u(x,y,z,t) = temperature of body

c = specific heat of body

ρ = density of body

It is an empirical fact that the rate of flow of heat into the volume V may be expressed in the following form:

images

where k is a constant known as the thermal conductivity and ∂u/∂n is the derivative of the temperature with respect to the outward drawn normal to the surface s.

If we differentiate (14.1) with respect to t and equate the result to (14.2), we obtain

images

If dn is a vector drawn in the direction of the outward drawn normal to the surface s, we have

images

where du is the total derivative of the temperature and represents the change in temperature as one moves through the distance dn. If we divide both sides of (14.4) by the differential dn, we obtain

images

where n represents a unit vector in the normal direction. We then have

images

If we let

images

we have, in view of (14.3) and (14.6),

images

where we have used Gauss’s theorem to transform the last integral. Transposing, we may write

images

Since the integral is continuous and the volume V is arbitrary, the integrand must vanish and we obtain

images

or

images

If k is a constant, we have

images

or

images

where

images

Equation (14.13) was derived by Fourier in 1822 and is sometimes called the heat-flow, or diffusion, equation.

15 THE GRAVITATIONAL POTENTIAL

Consider two particles at Q and P of masses m1 and m2, respectively. Then, according to the Newtonian law of gravitation, there is a force of attraction between them given in magnitude by the equation

images

where k is a constant depending upon the units and r is the distance between the particles as shown in Fig. 15.1.

If we choose the unit of mass as that of a particle which, placed at a unit distance from one of equal mass, attracts it with unit force, Eq. (15.1) becomes

images

If we denote the vector QP by r, we may express the force per unit mass at P due to the attracting particle at Q by

images

F is called the intensity of force or the gravitational field of force at the point P, and we note that it may be expressed as the gradient of the scalar m/r. It follows therefore that F satisfies the equation

images

and is therefore a conservative field of force.

Let us suppose that the particle m is stationary at Q and that another of unit mass moves under the attraction of the former from infinity up to P

images

Fig. 15.1

along any path. The work done by the force of attraction during an infinitesimal displacement dr of the unit mass is dr. The total work done by the force while the unit particle moves from infinity up to P is

images

This is independent of the path by which the particle comes to P and is called the potential at P due to the particle of mass at Q. Let us denote it by V. We then have

images

while the intensity of force at P due to it is

images

That is, the intensity at any point is equal to the gradient of the potential.

If we now suppose that there are n particles of masses m1, m2, . . ,mn, relative to which P has position vectors r1, . . . , rn, respectively, then the force of attraction per unit mass at P due to the system is the vector sum of the intensities due to each, that is,

images

Now, by the same argument as before, the work done by the attracting forces on a particle of unit mass while it moves from infinity up to P is

images

Therefore the potential at P due to a system of particles is the sum of the potentials due to each. The potential V is a scalar function of position and, except at the points Qs where the masses are situated, it satisfies

images

which is Laplace’s equation. Since F = ∇V, this means that at points excluding matter we must have

images

CONTINUOUS DISTRIBUTION OF MATTER

Let us now suppose that the attracting matter forms a continuous body filling the space bounded by the closed surface s. We divide the body into an infinite number of elements of mass ρdV, where ρ is the density of the body and dV is the element of volume of the body, as shown in Fig. 15.2.

If P is outside the body, the sum in Eq. (15.9) passes into the integral

images

where r is the distance from the element of volume dV to the point P. gives the point P. This potential at P due to the entire body.

If the point P is inside the body, the integrand (15.12) becomes infinite. In this case we define the potential in the following manner: Surround the point P by a closed spherical surface s0, and consider the potential due to the matter in the space between s0 and s. The integrand is finite everywhere, since P is outside the region. Now let the surface s0 decrease indefinitely, converging to the point P as a limit.

We then consider

images

Since the volume of s0 is of the same order as r3, where r is the radius of the sphere s0, while the integrand becomes infinite like 1/r if ρ is finite, the value of the above integral tends to a definite limit, which is called the potential at P due to the whole body.

POISSON’S EQUATION

We have seen that the potential function satisfies ∇2V = 0, or Laplace’s equation, in the region outside matter. Let us now consider the equation satisfied by the potential in a region inside matter. Consider a point P inside a body of density ρ. Surround the point P by a sphere s0 of radius a, as shown in Fig. 15.3.

The gravitational field at the point P is the vector sum of the gravitational field F0 produced by the matter outside s0 and Fi produced by the matter inside s0. That is, we write

images

images

Fig. 15.2

images

Fig. 15.3

Now let us calculate ∇·FP, that is,

images

But since F0 is produced by matter outside s0, we have from (15.11)

images

Hence we have

images

Consider the sphere of radius a; its mass is therefore equal to

images

Now as a tends to zero, the intensity at the surface of this sphere is equal to

images

in magnitude and in the direction of the inwardly drawn normal to the surface. By Gauss’s theorem we then have

images

Hence

images

But

images

Hence by (15.21) the equation satisfied by the potential in a region containing matter is

images

This relation is known as Poisson’s equation. In view of (15.13) we see that we may take

images

as a solution of Poisson’s equation.

GAUSS’S LAW OF GRAVITATION

As a consequence of Poisson’s equation we may prove that the surface integral of the gravitational force F over a closed surface s drawn in the field is equal to –4π times the total mass enclosed by the surface.

To establish this, apply Gauss’s theorem to Eq. (15.22); we then have

images

But ∫ ∫ ∫ ρ dv is the total mass enclosed by the surface s. This theorem has many applications in potential theory and in the theory of electrostatics.

16 MAXWELL’S EQUATIONS

The study of electrodynamics affords one of the most important applications of vector analysis. According to the modern point of view, by an electromagnetic field is understood the domain of the five vectors E, B, D, H, and J. These vectors satisfy the differential equations

images

images

images

images

and in a homogeneous isotropic medium we have the additional relations

images

images

images

The above set of differential equations are the fundamental equations of Maxwell written in a rationalized mks system of units. In this system of units we have

E = electric intensity, volts per m

B = magnetic induction, webers per m

D = electric displacement, coulombs per m2

H = magnetic intensity, amp per m

J = current density, amp per m2

σ = electric conductivity, 1/ohm-m

K= KrK0 = electric inductive capacity of medium

μ = μrμ0 = magnetic inductive capacity of medium

Kr = dielectric constant

μr = permeability

K0 = 8.854 × 1012 farad per m

μ0 = 4π× 10–7 = 1.257 × 10-6 henry per m

ρ = charge density, coulombs per m3

c = images= 2.998 × 108 m per sec

The solution of electrodynamic problems depends on the solution of these equations in special cases. For example, we may discuss briefly the following special cases.

ELECTROSTATICS AND MAGNETO STATICS

In the case that the field vectors are independent of the time and in a region where J = 0, we see that all the terms involving partial derivatives with respect to time vanish and the electric vectors and magnetic vectors become independent of each other. We then have

images

images

images

and

images

images

Since the curls of E and H vanish, we see that both these fields may be derived from potential functions. It is conventional to write

images

images

where VE and VM are the electric and magnetic scalar potentials, respectively.

In view of (16.9) and (16.12) we see that these potentials satisfy the equations

images

and

images

That is, the electric potential satisfies Poisson’s equation, and the magnetic potential satisfies Laplace’s equation. If the charge density ρ vanishes, then both VE and VM satisfy Laplace’s equation and the solution of electrostatic and magnetostatic problems reduces to the solution of this equation subject to the proper boundary conditions.

17 THE WAVE EQUATION

Let us eliminate the magnetic-intensity vector H from the Maxwell field equations. To do this, let us take the curl of both sides of Eq. (16.1); we then have

images

in view of the fact that B = μH and that the operators × and ∂/∂t commute.

Equation (16.2) may be written in the form

images

Substituting this into (17.1) and making use of the identity

images

we obtain

images

Now, in view of (16.4), we have

images

and it may be shown that in free space or in a conducting medium ρ is independent of the field distribution and may be taken to be equal to zero. Accordingly, Eq. (17.4) becomes

images

As a consequence of (16.7) we also have

images

Eliminating E from the Maxwell equations, we also obtain

images

It is thus seen that the three vectors H, E, and J satisfy equations of the same form.

In free space we have the conductivity σ = 0 and μr = Kr = 1; we then obtain

images

and

images

If we let

images

We may write (17.9) in the form

images

This equation is the general wave equation in vector form and is discussed in Chap. 12. It is there shown that this equation governs the propagation of various entities, such as electric waves, displacements of tightly stretched strings, deflections of membranes, etc. It will be shown that such disturbances are propagated with a velocity equal to c.

Equations (17.9) and (17.10) are taken as the starting point in the theory of electromagnetic waves. In this case, c is the velocity of light in free space and is approximately 3 × 108 m per sec.

18 THE SKIN-EFFECT, OR DIFFUSION, EQUATION

In metals σ is of the order of 107(l/ohm-m) and imagesr ≈ 1, and it may be seen that the term involving the second derivative in (17.7) may be neglected for frequencies of the order of 1010 cycles per sec and lower. In such a case this equation reduces to

images

We thus see that the current density in metals satisfies an equation of the same form as is satisfied by the equation of heat flow in solids as given in Sec. 14. In the electrical-engineering literature this equation is called the skin-effect equation. The solution of this equation in various special cases is considered in Chap. 11.

19 TENSORS (QUALITATIVE INTRODUCTION)

In many physical problems vector notation is not sufficiently general to express the relations between the quantities involved. For example, in electrostatics, if a dielectric medium is isotropic, the relation between the electric-displacement vector D and the electric-intensity vector E is given by the vector equation

images

where the scalar K is the electric inductive capacity of the medium (see Sec. 16). If, however, the dielectric medium is not isotropic, the electric-displacement vector D does not have the same direction as the vector E, so that the scalar K must be replaced by a more general quantity to effect the required change in direction and magnitude. Such a quantity is called a tensor.

A similar situation arises in the formulation of elastic problems. If F is the stress vector in an isotropic medium and X is the strain vector, then the relation

images

where k is a scalar, exists. If the medium is not isotropic, F and X are no longer in the same direction and Eq. (19.2) takes on a more general form in which k is a tensor.

After this qualitative introduction, the simpler analytical properties of tensors will be discussed in the following sections.

20 COORDINATE TRANSFORMATIONS

The fundamental definitions involved in the calculus of tensors or tensor analysis are intimately connected with the subject of coordinate transformations. In this section some of the fundamental properties of these transformations will be discussed.

LINEAR TRANSFORMATIONS AND THE SUMMATION CONVENTION

Consider a set of variables (u1,u2,u3) related to another set of variables (w1,w2,w3). (The superscripts denote different variables and not the powers of the variables involved.) Let the relation between the two sets of variables be of the form

images

where the coefficients images are constants. In this case the variables (u1,u2,u3) are related to the variables (w1,w2,w3) by a linear transformation. This transformation may be expressed by the equation

images

For convenience, in tensor analysis, the following summation convention has been adopted:

The repetition of an index in a term once as a subscript and once as a superscript will denote a summation with respect to that index over its range. In (20.2) the range of n is from 1 to 3, so that, by the above summation convention, (20.2) may be written in the form

images

The determinant images is the determinant of the linear transformation (20.1). If C≠0, the transformation will have an inverse. Let images be the cofactor of images in the determinant C, divided by C. Then, by Chap. 2, Eq. (4.8), we have

images

The symbol images is called a Kronecker delta and has the values indicated in (20.4). In order to solve the system of Eqs. (20.1), multiply both sides by b\ and obtain

images

Therefore the solution of the system (20.3) is

images

and therefore the transformation (20.1) is reversible.

SUCCESSIVE LINEAR TRANSFORMATIONS

Consider the two linear transformations

images

and

images

where all the indices range from 1 to n. If (20.7) is substituted into (20.8), the result is

images

where

images

is the transformation matrix of the linear transformation (20.9). These linear transformations may also be expressed by means of the matrix notation of Chap. 3 in the following form:

images

Therefore

images

where the matrices (x), (y), (z) are column matrices of order n and the matrices [a], [b], and [c] are square matrices of order n.

FUNCTIONAL TRANSFORMATIONS, CURVILINEAR COORDINATES

Consider a three-dimensional Euclidean space. The position of a point P in this space can be specified by its coordinates referred to an orthogonal cartesian system of axes. Let (y1,y2,y3) denote such a system of coordinates. Let

images

be a transformation of coordinates from the rectangular cartesian coordinates (y1,y2,y3) to some general coordinates (x1,x2,x3), not necessarily rectangular cartesian coordinates. For example, (x1,x2, x3) may be spherical coordinates. The inverse of the transformation of coordinates (20.13) is the transformation that takes one from the coordinates (x1,x2,x3) to the rectangular cartesian coordinates (y1,y2,y3). Let

images

be the inverse of the coordinate transformation (20.13). When this transformation exists, it is seen that to each set of values yr there is a unique set of values of xr, and vice versa. Hence the variables xr determine a point in space uniquely and they are called curvilinear coordinates.

SPHERICAL POLAR COORDINATES

The spherical polar coordinates of Sec. 12 may be taken as a special example of the general transformations (20.13) and (20.14). In this notation, consider the spherical polar coordinates of Fig. 20.1. In this figure y1, y2, y3 are the cartesian coordinates of a point P, and x1, x2, x3 are the spherical polar coordinates of P. x1 is the radius vector from the origin to the point P, and

images

Fig. 20.1

x2 and x3 are the angles shown in Fig. 20.1. In this special case the transformation (20.13) takes the form

images

The inverse transformation (20.14) in this case takes the following form:

images

21 SCALARS, CONTRAVARIANT VECTORS, AND COVARIANT VECTORS

The fundamental types of tensors will now be defined with respect to functional transformations from one set of curvilinear coordinates to another. The discussion of tensors with respect to general functional transformations follows the same fundamental principles and may be carried out by considering a space of n dimensions.

a SCALARS

Let a Euclidean space be specified by the curvilinear coordinates (x1,x2,x3). Let new curvilinear coordinates images be introduced by the transformation

images

Let a quantity have the value N in the old variables xr and the value images in the new variables images. Then, if the quantity has the same value in both sets of variables so that

images

the quantity N is called a scalar, or an invariant, or a tensor of order zero.

b CONTRAVARIANT VECTORS

Let three quantities have the values (A1,A2,A3) when expressed in terms of the coordinates (x1,x2,x3), and let these quantities have the values images when expressed in the coordinates images If

images

then the quantities (A1,A2,A3) are said to be the components of a contra-variant vector or a contravariant tensor of the first rank with respect to the transformation (21.1).

Taking differentials on both sides of (21.1),

images

Since the components of an ordinary vector in three-dimensional space transform as the differentials of the coordinates, it is seen that the components of an ordinary vector are actually the components of a contravariant tensor of rank 1.

If three quantities have the values (A1,A2,A3) in the system of coordinates (x1,x2,x3) and the values images in the system of coordinates images and if

images

then (A1A2,A3) are the components of a covariant vector or a covariant tensor of rank 1.

TENSORS OF HIGHER RANKS

In the above three classes of quantities, scalars or invariants, contravariant and covariant vectors, there exist components in any two coordinate systems, and the components in any two coordinate systems are related by the defining laws of transformation. There are other quantities whose components in any two coordinate systems are related by characteristic laws of transformation. These laws of transformation are involved in the definitions of tensors of higher ranks. There are three varieties of second-rank tensors:

images

images

images

The A’s with the bars denote components in the coordinates images and those without the bars denote components in the coordinates (x1,x2,x3).

Tensors of higher rank are defined by similar laws; for example, a mixed tensor of rank 4 is

images

A very useful mixed tensor of the second rank is the Kronecker delta images This is defined by the equation

images

To determine whether images is a mixed tensor of the second rank, we investigate whether it satisfies the definition (21.8).

images

But

images

since images and images are independent variables. Hence

images

We thus see that images has the same components in all coordinate systems. The relation

images

is a very useful one.

22 ADDITION, MULTIPLICATION, AND CONTRACTION OF TENSORS a ADDITION

It can be easily shown that the sum or difference of two or more tensors of the same rank and type is a tensor of the same rank and type. For example, if

images

it follows from (22.1) that Cmn is a contravariant tensor of the second rank.

b THE OUTER PRODUCT

If the contravariant tensor Am is multiplied by the covariant tensor Bn, the product is the following mixed tensor of the second rank:

images

It is easy to see that images transforms like (21.8). The product (22.2) is called the outer product. It may be obtained with tensors of any rank or type. For example, the tensor images may be multiplied by the tensor Bpq to obtain the following outer product:

images

c CONTRACTION

Consider the mixed tensor images and let the index m be set equal to n so that m = n. The law of transformation (21.8) now gives the result

images

and hence images is a scalar. This process of summing over a pair of contra-variant and covariant indices is called contraction. As another example, consider the mixed tensor images and let m = q. Let Bnp = images The law of transformation now gives

images

Therefore images is a covariant tensor of rank 2. The process of contraction always reduces the rank of a mixed tensor by 2.

d THE INNER PRODUCT

If two tensors are multiplied together and then contracted, the result is called the inner product. For example, the inner product is

images

The inner product of Amn and Brpq is

images

The product Am Bm is equivalent to the scalar product of two vectors in rectangular coordinates. In tensor analysis the length L of the tensor Am or Am is defined by the equation

images

The scalar product of the two vectors Am and Bm is

images

where θ is the angle between the vectors Am and Bm. Hence

images

If Am and Bm are orthogonal, then

images

23 ASSOCIATED TENSORS

a THE QUOTIENT LAW OF TENSORS

It can be shown by applying the law of transformation of tensors that, if the following relation exists,

images

where Cr is known to be a tensor and Bst is a known arbitrary tensor, then A(r,s,t) is a tensor of the rank

images

This law is of importance in recognizing tensors.

b THE METRIC TENSOR

Let the cartesian coordinates of the point P be yr. The element of length dS of such a coordinate system is given by

images

Since the cartesian coordinates are related to the general coordinates xi by the functional relations

images

we have

images

It therefore can be seen that

images

where

images

This equation shows that gmn is symmetric, and since dS is an invariant for arbitrary values of the contravariant vector dxm, it follows from the quotient law that gmn is a double covariant tensor. The tensor gmn is called the fundamental, or metric, tensor. Let g denote the determinant |gmn| and Gmn denote the cofactor of gmn in g. We now define the quantity gmn by the equation

images

Now, from the theory of determinants [see Chap. 2] we have

images

Hence by the quotient law it can be seen that gmn is a double contravariant tensor of rank 2.

Two vectors related by the following equations,

images

and

images

are called associated vectors. In the literature of tensor analysis it is sometimes said that Am and Am are the same vector, Am being the contravariant components and Am the covariant ones. Besides the fundamental, or metric, tensor gmn and the second fundamental tensor gmn there is the third associated tensor, images defined by the equation

images

This relation follows from (23.9).

Associated tensors of any rank may be obtained in the same manner as (23.10) and (23.11). For example,

images

If we are given a contravariant vector Ar, we can form an invariant A by means of the equation

images

A is called the magnitude of the vector A. In the same manner the magnitude B of the covariant vector Bm is defined by the equation

images

A unit vector is one whose magnitude is unity. If (23.6) is divided by images we have

images

From this it can be concluded that dxm/dS is a unit vector. By the use of (22.9) and (23.11) it can be shown that the angle θ between the vectors Am and Bm may be expressed in the form

images

24 DIFFERENTIATION OF AN INVARIANT

Let ϕ be an invariant function of a parameter t; then, in the new variables images we have images and it can be shown that

images

so that dϕ/dt is also an invariant. If ϕ is an invariant function of xr, we have

images

This shows that ∂ϕ/∂xr is a covariant vector since it transforms like one.

25 DIFFERENTIATION OF TENSORS: THE CHRISTOFFEL SYMBOLS

In Sec. 24 it has been shown that the derivative of a scalar point function is a covariant vector. However, the derivative of a covariant vector is not a tensor, for if

images

the derivative of Ām with respect to images is

images

The presence of the second derivative shows that ∂Am/∂xn does not transform like a tensor.

In order to define “derivatives” of tensors that preserve the proper tensor character, it is necessary to introduce the following notation:

images

and

images

The quantity [mn,p] is called the Christoffel symbol of the first kind, and images is called the Christoffel symbol of the second kind.

a THE INTRINSIC DERIVATIVE OF A VECTOR

It can be shown that the quantity

images

is a contravariant vector. This vector is defined to be the intrinsic derivative of the vector Ar with respect to the parameter t.

The quantity

images

can be shown to be a covariant vector and is defined to be the intrinsic derivative of Ar with respect to the parameter t.

b THE COVARIANT DERIVATIVE OF A VECTOR

The covariant derivative of the contravariant vector Ar with respect to the coordinate xs is defined by the equation

images

This quantity can be shown to be a mixed tensor contravariant in r and covariant in s. The covariant derivative of the covariant vector Ar with respect to the coordinate xs is defined by the equation

images

This quantity can be shown to be a tensor covariant in both r and s.

26 INTRINSIC AND COVARIANT DERIVATIVES OF TENSORS OF HIGHER ORDER

The intrinsic derivative of the tensor images with respect to the parameter t is defined by the equation

images

It can be shown that this quantity is a tensor of the type images

The covariant derivative of the tensor images with respect to the variable xn is defined by the equation

images

This is a tensor with one covariant index more than images The definitions of the intrinsic and covariant derivatives are quite general. Terms of the forms (25.5) and (25.7) appear in the derivatives of a mixed tensor for each contravariant index, and terms of the forms (25.6) and (25.8) appear in the derivatives of mixed tensors for each covariant index.

If the coordinate system is a rectangular cartesian one, the gmn quantities are all constants and the Christoffel systems are identically zero. In such case the intrinsic derivatives are identical to the ordinary derivatives, and the covariant derivatives are ordinary partial derivatives.

The fundamental, or metric, tensor grs has components that are constants in cartesian systems, and therefore

images

in such a system. Since grs is a tensor, it follows that, in every coordinate system, the following equation must be satisfied:

images

Therefore the covariant derivative of the metric tensor grs is zero. It can also be shown that the covariant derivative of the associated tensor grs is zero.

27 APPLICATION OF TENSOR ANALYSIS TO THE DYNAMICS OF A PARTICLE

In the brief introduction to the basic ideas and definitions of tensor analysis given in this appendix, space does not permit the discussion of numerous applications to applied mathematics. These applications will be found in the treatises on tensor analysis listed in the references at the end of this chapter. For purposes of illustration, however, a brief discussion of the application of tensor analysis to the dynamics of a particle will be given in this section.

Consider a particle of mass M that is moving in space. Let the position of this particle be given by a system of curvilinear coordinates xr. As the time t varies, the particle will describe a certain curve in space, called the trajectory of the particle. The equations of this curve may be written in the form

images

If we transform to another curvilinear coordinate system images in accordance with the transformation

images

then the trajectory of the particle may be obtained in terms of t by substituting (27.1) into (27.2).

Consider the quantity

images

In the new coordinates images the corresponding quantities are

images

This equation expresses the fact that the υr are the components of a contra-variant vector. In dynamics this vector is called the velocity of the particle. The velocity vector υr is a function of the time t. If the intrinsic derivative with respect to t is taken, the following result is obtained:

images

where images are the Chistoffel symbols of the coordinate system xr..

If the coordinates are rectangular cartesian, the Christoffel symbols are identical to zero and the vector ar becomes

images

These are the components of acceleration along the three coordinate axes; (27.5) is called the acceleration vector of the particle. If the mass of the particle M remains constant, it is obviously a quantity independent of the coordinate system and the time and is, therefore, an invariant.

Newton’s second law of motion may be expressed in the form

images

where the contravariant vector Fr is called the force vector. This vector completely specifies the magnitude and direction of the force acting upon the particle. In cartesian coordinates (27.7) takes the well-known form

images

and F1, F2, F3 are the components of the force along the three axes of coordinates. In curvilinear coordinates this equation takes the form

images

WORK AND ENERGY: LAGRANGE’S EQUATIONS

If the axes images are rectangular cartesian, a force whose components in these coordinates are images and whose point of application is moved through a small displacement images does an amount of work given by

images

In orthogonal cartesian coordinates the associated vector images has exactly the same components as images and the expression for δ W becomes

images

From the relation connecting images and images it can be seen that δW can also be written in the form

images

Since δW is an invariant, it can be concluded from the above equation that the Fr are the components of a covariant vector. Since the components of Fr in an orthogonal cartesian coordinate system are those of the force vector, it follows that Fr is the covariant force vector in the coordinates xr.

The covariant and contravariant components of this vector are related to each other by the following formulas:

images

If the expression Frdxr is a perfect differential, then the force is said to be conservative and a function V may be defined by the equation

images

The function V is called the force potential. From (27.14) it follows that

images

The kinetic energy T of a particle is defined as images, where υ is the magnitude of the velocity. Therefore

images

where gmn is the metric tensor of the curvilinear coordinates xm.

LAGRANGE’S EQUATIONS

If T is differentiated partially with respect to imagesr, the result is

images

Therefore

images

Also, if T is differentiated partially with respect to xr, the result is

images

From (27.18) and (27.19) it can be deduced that

images

where [mn,r] is the first Christoffel symbol. Therefore

images

As a consequence of (27.5) it can be seen that the right-hand member of (27.21) is equal to Mgrsas = Mar. Therefore (27.21) may be written in the form

images

These are Lagrange’s equations. If the force system is conservative, these equations may be written in the form

images

PROBLEMS

1. The point of application of a force F = 5, 10, 15 lb is displaced from the point (1,0,3) to the point (3,–1,–6). Find the work done by the force.

2. Find the scalar product of two diagonals of a unit cube. What is the angle between them?

3. A force F acts at a distance r from the origin. Show that the torque L about any axis through the origin is L = (r × F) a when a is a unit vector in the direction of the axis.

4. Show that the lines joining the midpoints of the opposite sides of a quadrilateral bisect each other.

5. Show that the bisectors of the angles of a triangle meet at a point.

6. What is the cosine of the angle between the vectors A = 4i + 6j + 2k and B = i – 2j + 3k ?

7. Let r be the radius vector from the origin to any point and a constant vector. Find the gradient of the scalar product of a and r.

8. A plane central field A is defined by A = rF(r). Determine F(r) so that the field may be irrotational and solenoidal.

9. A central field A in space is defined by A = rF(r). Determine F(r) so that the field may be irrotational and solenoidal.

10. If r is a unit vector of variable direction, the position vector of a moving point may be written r = rr. Find by vector methods the components of the acceleration F parallel and perpendicular to the radius vector of a particle moving in the xy plane.

11. Prove that the vector ∇ϕ is perpendicular to the surface ϕ(x,y,z) = const.

12. Find ∇u if u = In(x2 + y2 + z2).

13. Show that if r is the position vector of any point of a closed surface s, then

images

where V is the volume bounded by s.

14. Show that a vector field A is uniquely determined within a region V bounded by a surface S when its divergence and curl are given throughout V and the normal component of the curl is given on S.

15. With every point of a curve in space there is associated a unit vector t the direction of which is that of the velocity of a point describing the curve. This vector therefore has the direction of the tangent. Prove that t·dt/ds = 0, where s is the length of the arc of the curve measured from a fixed point on it. What is the geometrical meaning of dt/ds ?

16. The expression for (ds)2 is (ds)2 = (u2 + v2)[(du)2 + (dv)2] + u2v2()2 in parabolic coordinates. What is the form assumed by Laplace’s equation in these coordinates?

17. Write the heat-flow equation in cylindrical and spherical coordinates.

18. What form does the general wave equation take in cylindrical coordinates?

19. Write Maxwell’s equations in cylindrical coordinates.

20. Show that the potential due to a solid sphere of mass M and uniform density at an external point at a distance r from the center is M/r and, hence, that the gravitational intensity is M/r2 toward the center.

21. Show that the gravitational field inside a homogeneous spherical shell is zero.

22. Find the potential inside a solid uniform sphere and the gravitational force inside the sphere.

23. Find the covariant components of the acceleration vector in spherical coordinates.

24. Show that the velocity vector is given by (1/M)grs(∂T/∂imagess).

25. If ϕ is an invariant function of xr and imagesr, show that the partial derivative of images with respect to imagesr is a covariant vector.

26. If a particle is moving with uniform velocity in a straight line prove that δυr/δt = 0.

27. Find the contravariant components of the acceleration vector in cylindrical coordinates.

REFERENCES

1901. Gibbs, J. Willard, and Edwin Bidwell Wilson: “Vector Analysis,” Yale University Press, New Haven, Conn.

1911. Coffin, Joseph George: “Vector Analysis,” John Wiley & Sons, Inc., New York.

1928. Weatherburn, C. E.: “Advanced Vector Analysis with Application to Mathematical Physics,” George Bell & Sons, Ltd., London.

1928. Weatherburn, C. E.: “Elementary Vector Analysis with Applications to Geometry and Physics,” George Bell & Sons, Ltd., London.

1931. McConnel, A. J.: “Applications of the Absolute Differential Calculus,” Blackie & Son, Ltd., Glasgow.

1931. Wills, A. P.: “Vector Analysis with an Introduction to Tensor Analysis,” Prentice-Hall, Inc., Englewood Cliffs, N.J.

1932. Gans, Richard: “Vector Analysis with Applications to Physics,” Blackie & Son, Ltd., Glasgow.

1946. Brillouin, L.: “Les Tenseurs en mécanique et en élasticité,” Dover Publications, Inc., New York.

1951. Sokolnikoff, I. S.: “Tensor Analysis,” John Wiley & Sons, Inc., New York.

This product is commonly called the scalar triple product and is denoted by (ABC).

See J. L. Synge and B. A. Griffith, “Principles of Mechanics,” 1st ed., Chap. 13, McGraw-Hill Book Company, New York, 1942.

For a more rigorous derivation of Poisson’s integral, refer to O. D. Kellogg, “Foundations of Potential Theory,” Springer-Verlag OHG, Berlin, 1929.

See A. J. McConnel, “Applications of the Absolute Differential Calculus,” pp. 143–150, Blackie & Son, Ltd., Glasgow, 1931.