12
Vector calculus

In Chapter 8 we introduced the idea of a vector as a quantity with both magnitude and direction and we discussed vector algebra, particularly as applied to analytical geometry, and the differentiation and integration of vectors with respect to a scalar parameter. In this chapter we extend our discussion to include directional derivatives and integration over variables that are themselves vectors. This topic is called vector calculus or vector analysis. It plays a central role in many areas of physics, including fluid mechanics, electromagnetism and potential theory.

12.1 Scalar and vector fields

If scalars and vectors can be defined as continuous functions of position throughout a region of space, they are referred to as fields and the region of space in which they are defined is called a domain. An example of a scalar field would be the distribution of temperature T within a fluid. At each point the temperature is represented by a scalar field T(r) whose value depends on the position r at which it is measured. A useful concept when discussing scalar fields is that of an equipotential surface, that is, a surface joining points of equal value. This is somewhat analogous to the contour lines on a two-dimensional map, which join points of equal height. An example of a vector field is the distribution of velocity v(r) in a fluid. At every point r, the velocity is represented by a vector of definite magnitude and direction, both of which can change continuously throughout the domain. In this case, we can define flow lines such that the tangent to a flow line at any point gives the direction of the vector at that point. Flow lines cannot intersect. This is illustrated in Figure 12.1.

images

Figure 12.1 The motion of a fluid around a smooth solid. The coloured lines are the flow lines and the arrows show the direction of the vector field, in this case the velocity v(r).

In the rest of this section, we shall extend our discussion of differentiation to embrace scalar and vector fields. Since we are primarily interested in applications where these fields are physical quantities, we shall assume throughout this chapter that they and their first derivatives are single-valued, continuous and differentiable.

12.1.1 Gradient of a scalar field

If we consider the rate of change of a scalar field ψ(r) as r varies, this leads to a vector field called the gradient of ψ(r), written as . We will define and derive an expression for this by reference to Figure 12.2.

images

Figure 12.2 Diagram for the derivation of the gradient of a scalar field ψ(r). PR is normal to the equipotential surface ψR at the point of intersection P.

Consider the point P on an equipotential surface ψ = ψP of the scalar field ψ(r). Let R be a point on the normal to the surface through P that also lies on an equipotential surface ψR > ψP. If we take the two surfaces to be close, then they will be approximately parallel to each other and grad ψ(r) at P is defined as a vector in the direction PR of magnitude

numbered Display Equation

This definition is similar to the definition of a derivative; hence the name gradient. Now let PQ be the signed distance from P to the surface ψR measured in the positive x-direction, and let α be the angle between PR and the x-direction, as shown in Figure 12.2. Then as PR → 0, , and thus the component of in the x-direction is

(12.1) Unnumbered Display Equation

since the point Q is on the surface ψ = ψR. The right-hand side of (12.1) is . Similarly, the components of grad ψ(r) in the y and z directions are and , and hence

(12.2) Unnumbered Display Equation

Further, if we make a small displacement δr = (δx, δy, δz) from P in any direction, we have

(12.3) Unnumbered Display Equation

which using (12.2) is

(12.4a) Unnumbered Display Equation

Similarly the corresponding differential [cf. (7.10), (7.11)] is given by

(12.4b) Unnumbered Display Equation

It follows from (12.4a) and (12.4b) that the rate of change of the field with respect to an infinitesimal displacement depends on the direction of travel. For this reason, it is called the directional derivative. To find it, consider moving a distance ds in the direction specified by a unit vector , so that . Substituting this expression into (12.4b) and dividing by ds then gives

(12.5) Unnumbered Display Equation

as the directional derivative of ψ in the direction .

Another way of writing (12.2) and (12.5) is in terms of an object , called del, and defined in the Cartesian system by

(12.6) Unnumbered Display Equation

so that (12.2) becomes

(12.7) Unnumbered Display Equation

Del is called a vector operator, meaning that it acts on (i.e. operates on) a scalar field ψ to give a vector field . It is also an example of a differential operator in that it involves derivatives and, like all operators,1 it acts only on objects to its right.

The directional derivative (12.5) can also be rewritten using (12.7) and the definition of , when it becomes

(12.8) Unnumbered Display Equation

Since is, by definition, normal to the equipotential surface at P(r), (12.8) satisfies the requirement that if is along the direction of a tangent to the equipotential surface at P and it attains its maximum value of when and are in the same direction. The relation between these quantities is shown in Figure 12.3.

images

Figure 12.3 Gradient and directional derivative, where the vector u indicates the direction of defined in the text.

12.1.2 Div, grad and curl

Because del is a vector operator, it can form both scalar and vector products with vector fields. Thus if V is the vector field

numbered Display Equation

then the scalar product with del is

(12.9) Unnumbered Display Equation

This is called the divergence of V and is written . Therefore . The vector product of del with V is

(12.10) Unnumbered Display Equation

and is called the curl of V. Thus . The origin of these names will emerge later in the chapter. Note that in both cases del is to the left of V because it is an operator. Thus, for example, and do not have the same meaning, even though they are both scalar products of the same quantities. The former is the simple scalar given in (12.9), the latter is the scalar differential operator

numbered Display Equation

Various combinations of div, grad and curl can also be formed. For example, if ψ is a scalar field, then is a vector field. Hence, if we choose , we can take its divergence to give

(12.11a) Unnumbered Display Equation

where the scalar operator

(12.12) Unnumbered Display Equation

is called the Laplacian operator and is called the Laplacian of ψ. The Laplacian is an important operator in physical science and occurs very frequently, for example, in the wave equation

numbered Display Equation

where is the wave velocity. Similarly we note that, since the divergence of a vector field V is itself a scalar field, we can take its gradient to give

(12.11b) Unnumbered Display Equation

From this we see that grad div acts on a vector field V to give a vector field (12.11b) and is quite different from div grad, which acts on a scalar field ψ to give a scalar field (12.11a). This illustrates again that care is required with the order of factors when operators are involved. The Laplacian can also operate on a vector field V to give another vector field defined by

(12.13) Unnumbered Display Equation

The combination and grad div are two of only five valid combinations of pairs of div, grad and curl. The other three, together with important identities which they satisfy, are

(12.14a) Unnumbered Display Equation

(12.14b) Unnumbered Display Equation

(12.14c) Unnumbered Display Equation

For example, from (12.2) and (12.10) we have

numbered Display Equation

and the other two identities also follow from the definitions of div, grad and curl.

In addition, there are many other identities involving del and two or more scalar or vector fields. They can all be verified by using the previous formulas, taking to be a differential vector operator. Some useful identities involving two fields are given in Table 12.1, where a, b are arbitrary vector fields, and ψ and φ are arbitrary scalar fields.

Table 12.1 Some useful identities involving del

12.1.3 Orthogonal curvilinear co-ordinates

So far, we have defined div, grad and curl in Cartesian co-ordinates. However, in problems with spherical or cylindrical symmetry, it is much easier to work in spherical or cylindrical polar co-ordinates, which reflect the symmetry of the problem. As we saw in Section 11.3, these two co-ordinate systems are examples of orthogonal curvilinear co-ordinates ui(i = 1, 2, 3), which are such that distances dr are obtained from formulas of the type

(12.15) Unnumbered Display Equation

where the unit vectors ei are orthogonal. The scale factors hi are given by (11.30b), which for the special case of polar co-ordinates are [cf. (11.36) and (11.42)]

(12.16) Unnumbered Display Equation

Here we shall first give the forms of etc. in orthogonal curvilinear co-ordinates, and then obtain the corresponding expressions from them for the cases of spherical and cylindrical polar co-ordinates.

Consider firstly the gradient of a scalar ψ, that is, . Returning to Figure 12.2, we let PQ be in the direction of u1, rather than x as before. Then the component of in the direction of u1 (with u2 and u3 held fixed) is the directional derivative , where ds = hdu1, that is, the component of in the direction e1, is

numbered Display Equation

and similarly for the other directions. Thus,

(12.17) Unnumbered Display Equation

For example, for spherical polar co-ordinates

numbered Display Equation

and using (12.16), gives

numbered Display Equation

The derivations of the corresponding results for and using the technique above are more difficult. The derivations are much easier using results we will obtain in Sections 12.3 and 12.4, and will be given there. For the present we will just quote the results

(12.18) Unnumbered Display Equation

(12.19) Unnumbered Display Equation

and

(12.20) Unnumbered Display Equation

The expressions for given earlier for the special case of Cartesian co-ordinates are easily regained by setting (u1, u2, u3) = (x, y, z) and hx = hy = hz = 1 in (12.17) to (12.20), respectively. The corresponding results for spherical polar and cylindrical polar co-ordinates are similarly obtained using (12.16) for the scale factors and are shown in Table 12.2.

Table 12.2 Grad, div, curl and the Laplacian in polar co-ordinates

12.2 Line, surface, and volume integrals

In this section, we shall extend the discussion of line integrals given in Chapter 11 to embrace integrals of a vector field and use vector methods to define integrals of a vector field over a curved surface.

12.2.1 Line integrals

In Section 8.3.1, we saw that the running vector

(12.21) Unnumbered Display Equation

where a and b are fixed vectors, described a straight line passing through the point r = a in the direction as the scalar parameter s varied in the range − ∞ < s < ∞. More generally, any running vector r(s), where r is a differentiable function of s, will describe a curve in space and for any given s the differential

(12.22) Unnumbered Display Equation

is an infinitesimal vector directed along the tangent to the curve, as shown in Figure 12.4 for the point P corresponding to s = s0. For example, in Cartesian co-ordinates

numbered Display Equation

is the vector equation of the parabola y2 = 4ax lying in the plane z = 0, and the corresponding differential

numbered Display Equation

is an infinitesimal vector directed along the tangent to the curve, as shown in Figure 12.4 for the point P corresponding to s = s0.

images

Figure 12.4 Definition of a space curve for the line integral of a scalar product. The path is defined in terms of a parameter s, so that r = r(s).

Now suppose we have a vector field V(r). Then we can define two line integrals

(12.23) Unnumbered Display Equation

where C as usual denotes the path, or contour, of integration. The first of these is by far the most important in physics, and is the only one we shall discuss. One important example is the work done by a force field F(r). If F(r) is the force acting at the position r, then F(r) · dr is the work done by the force in moving from r to r + dr, and the integral

(12.24) Unnumbered Display Equation

is the work done in moving from the initial position ri to the final position rf along the path C. If the force returns to ri, then ri = rf and the integral is denoted

numbered Display Equation

where the circle on the integral sign emphasises that the path is a closed loop.

So far, we have not used any co-ordinates to define the integrals. In general, if we use a set of co-ordinates u1,  u2,  u3, such that

numbered Display Equation

and

numbered Display Equation

then,

(12.25) Unnumbered Display Equation

In Cartesian co-ordinates

numbered Display Equation

so that (12.25) becomes

(12.26a) Unnumbered Display Equation

Here x, y, z (and in general u1,  u2,  u3) are not independent variables along the path C, but are specified by a single parameter s, so that C = r(s) and (12.26a) becomes

(12.26b) Unnumbered Display Equation

using (12.22). In particular, s may be chosen to be one of the co-ordinates themselves, for example x, when (12.26a) may be used directly together with the relations y = y(x),  z = z(x) along the contour C.

At this point we note that (12.26) is identical with the line integral (11.13) discussed in Section 11.1.3, if the functions Q, R, P are replaced by functions Vx,  Vy,  Vz. Hence the methods and results discussed in Section 11.1.3 can be carried over, with a trivial relabeling, to the line integrals (12.26) as we shall illustrate in the next section.

12.2.2 Conservative fields and potentials

The result of a line integral of a vector between any two points will in general depend on the path taken between them. If, however, the line integral is independent of the path for any choice of end points within the field, the vector field V is said to be conservative. Conservative fields play an important role in physics, as we shall now see.

Suppose that

(12.27) Unnumbered Display Equation

Then, using the expression (12.10) for the curl in Cartesian co-ordinates, we see that (12.27) implies that

numbered Display Equation

which is precisely the condition [cf. (11.23b) and (7.19b)] that

(12.28) Unnumbered Display Equation

is an exact, or perfect, differential. From (12.28) we immediately see that

numbered Display Equation

i.e.

(12.29) Unnumbered Display Equation

and that

(12.30) Unnumbered Display Equation

where ψA and ψB are the values of ψ at the points A and B. Hence V is a conservative field and can be derived from a scalar field ψ, called a potential field, or just a potential. We note that ψ(r) is only defined up to a constant by (12.29) and (12.30). This is usually chosen by requiring that ψ has a given value ψ0 at a reference point r0, or sometimes that ψ(r) → 0 as |r | → ∞.

The above argument shows that (12.27) is a sufficient condition for V to be a conservative field. That it is also a necessary condition is seen by reversing the argument. If V(r) is a conservative field, then we can define a potential by

numbered Display Equation

since the integral is independent of the chosen path between the reference point r0 and the point r. This implies that

numbered Display Equation

so that and by (12.14a). Hence is not only a sufficient condition, but also a necessary condition for V to be a conservative field.

An important example of a conservative field is the gravitational field. In general, if F(r) is the force acting on a particle at a position r, it is usual to introduce the potential energy due to gravity such that

(12.31) Unnumbered Display Equation

that is, the force acts in the direction of maximally decreasing potential energy. The work W done when F moves a particle from A to B is

(12.32) Unnumbered Display Equation

so that the work done by the force equals the loss of potential energy.

Of course not all forces are conservative. If dissipative forces such as friction are involved, then energy will be lost in moving from A to B in a way that depends on the path and a potential cannot be defined.

12.2.3 Surface integrals

In three dimensions, a surface is defined by an equation of the form

(12.33) Unnumbered Display Equation

where f(x, y, z) is a given function and d is a constant.2 Simple examples are the equation of a plane [cf (8.45a)],

(12.34) Unnumbered Display Equation

which is an example of an open surface; and the equation of a sphere

(12.35) Unnumbered Display Equation

which is an example of a closed surface. In addition, since (12.33) defines an equipotential surface for the scalar field f, it follows from the discussion of in Section 12.1.1 (cf. Figure 12.2) that at any point on the surface

(12.36) Unnumbered Display Equation

are the two vectors normal to the surface.

We now introduce integrals of a vector field V(r) over a surface S as follows. Given a small surface element ds, we form a vector surface element

(12.37) Unnumbered Display Equation

where is a unit vector normal to the surface at the position of ds, so that the direction of varies continuously over S. Surface integrals can now be defined of the form

(12.38) Unnumbered Display Equation

In each case, the integral is a double integral over a surface S, which may be open or closed. If the surface is closed, is chosen to point outwards from the closed region. If the surface is open it must be two-sided, that is, it is only possible to get from one side to the other by crossing the curve bounding the surface. Figure 12.6a shows a two-sided surface, whereas Figure 12.6b shows a so-called Mobius strip, which is one-sided.

images

Figure 12.6 Examples of open surfaces that are (a) two-sided (b) one-sided and (c) the use of a projection of a two-sided surface onto a plane to define the direction of .

For open surfaces, one must choose the direction for . However, if a direction is associated with a boundary curve that surrounds the surface, as it is in some very important applications, then is chosen to be ‘right-handed’. To see what this means, let us suppose that the surface and its boundary curve were to be projected onto a plane. Then, as shown in Figure 12.6c, is chosen so that the direction of integration around the contour of integration corresponds to that of a right-hand screw pointing in the direction of .

Of the two integrals (12.37), the scalar integrals are by far the more important. Their evaluation is often facilitated by choosing an appropriate co-ordinate system. For example, if one is integrating over a planar surface that lies in the x–y plane, then ds = dx dy k and

numbered Display Equation

On the other hand, suppose S lies on surface of a sphere of radius a. Then if we take the origin to be at the centre, the equation of the sphere in spherical co-ordinates is r = a, and one sees from Figure 11.14 that

(12.39a) Unnumbered Display Equation

Similarly, θ = θ0 is the equation of a cone with its axis along the z-direction and, from Figure 11.14, one sees that in this case

(12.39b) Unnumbered Display Equation

More generally, given any set of orthogonal curvilinear co-ordinates (u1, u2, u3), keeping any one of them constant defines a surface with, for example,

(12.40) Unnumbered Display Equation

if u1 is constant. Hence if

numbered Display Equation

an integral over a surface on which u1 is constant reduces to

numbered Display Equation

with similar expressions if either u2 or u3 is constant.3 These are straightforward double integrals, which can be evaluated using the methods discussed in Section 11.2.

If the surface does not correspond to a constant value of a suitably chosen orthogonal curvilinear co-ordinate, the integral can be evaluated using the projection method. In this method, the surface is projected onto a plane and the integral evaluated using Cartesian co-ordinates. This is illustrated in Figure 12.7, which shows an element of surface ds projected onto an element dA in the xy plane. From this figure,

numbered Display Equation
images

Figure 12.7 Diagram used to illustrate the evaluation of surface integrals by the projection method.

If the surface S is given by f(x, y, z) = d, then evaluated at the point on the surface, and so

(12.41) Unnumbered Display Equation

This general formula can be used to convert an integral over a curved surface S to an integral over A in the xy plane, as illustrated in Example 12.11 below.

12.2.4 Volume integrals: moments of inertia

In Section 11.4, we considered volume integrals of the form

(12.42) Unnumbered Display Equation

in Cartesian co-ordinates, where the integral extends over the region Ω, and the abbreviated notation on the left-hand side is sometimes used for convenience in what follows. We can now also define similar integrals over a vector field, i.e.

(12.43) Unnumbered Display Equation

whose evaluation essentially involves evaluating three integrals of the form (12.42).

To illustrate this, let us consider a solid body with variable density ρ occupying a region of space Ω. Then since the mass of a volume element is , the total mass of the body is given by

(12.44) Unnumbered Display Equation

while the formula

numbered Display Equation

for the centre-of-mass of a system of point particles of masses mi at positions r, becomes

(12.45) Unnumbered Display Equation

for an arbitrary solid body. Similarly, the formula

numbered Display Equation

for the moment of inertia I, where roi is the perpendicular distance from the mass mi to the axis of rotation, becomes

(12.46) Unnumbered Display Equation

12.3 The divergence theorem

The divergence theorem4 states that, for any vector field V,

(12.49) Unnumbered Display Equation

for any surface S enclosing a region Ω. The quantity V · ds is called the flux of V through ds and the circle on the double integral is to emphasise that S is a closed surface, by analogy with closed paths in line integrals. This circle is sometimes omitted and (12.49) is also sometimes written in the abbreviated form

numbered Display Equation

already used in (11.49) for volume integrals. However, in whatever form it is written, the theorem states that the volume integral of the divergence of V is equal to the total flux out of the bounding surface S, since points out of a closed surface.

We shall derive the divergence theorem and two well-known identities resulting from it in Section 12.3.1 below. Before that, we point out that the divergence theorem is central to the physical interpretation of divergence. To see this, we apply (12.49) to the case when S encloses a small volume element that shrinks to a point as . In this limit, the variation of V in can be neglected, so the left-hand side of (12.49) becomes , implying

(12.50) Unnumbered Display Equation

In other words, at a point r is the flux per unit volume out of an infinitesimal volume surrounding r. For example, if V = ρ v, where ρ is the density and v is the velocity field of a fluid, the flux V · ds is the rate of flow of mass through the surface. Hence if is greater than zero, there is a net flow of mass away from r, so that either the density is decreasing at the point, or a source (i.e. a point where fluid is entering the system) is present. On the other hand, if there is no source or sink (i.e. a point where fluid is leaving the system) at r, and the density is constant, which is normally a good approximation for a liquid, then . In this latter case V is called a solenoidal field. Although we have chosen the example of a fluid, the same ideas may be applied to other situations, including the flow of electric current.

12.3.1 Proof of the divergence theorem and Green's identities

To derive the divergence theorem, we consider a segment through the region Ω lying parallel to the x-axis and with constant infinitesimal cross section dy dz, as shown in Figure 12.85. Further, let the unit vectors and be the outward normals on the surface elements 1 and 2, respectively, where the segment intersects the surface of the region Ω, so that and

numbered Display Equation
images

Figure 12.8 A segment through the region Ω lying parallel to the x-axis and with constant infinitesimal cross section dy dz, used in the derivation of the divergence theorem.

Then, since

numbered Display Equation

at fixed y, z, we have

(12.51) Unnumbered Display Equation

where the right-hand side is the net flux through the surface elements 1 and 2 from the x-component Vxi.

All that remains now is to add together the contributions from enough segments to cover the whole region Ω, so that (12.51) becomes

numbered Display Equation

The contributions from the y and z components of V can be calculated in a similar way, and adding all three components we obtain

numbered Display Equation

which is the divergence theorem.6

Finally, we use the divergence theorem to derive two other useful results as follows. Let φ and ψ be two scalar fields continuous and differentiable in some region Ω bounded by a closed surface S. Applying the divergence theorem to gives

(12.52a) Unnumbered Display Equation

This is known as Green's first identity. Similarly, interchanging φ and ψ gives

numbered Display Equation

Subtracting these two equations gives

(12.52b) Unnumbered Display Equation

which is Green's second identity.

*12.3.2 Divergence in orthogonal curvilinear co-ordinates

Having derived the divergence theorem (12.49) using Cartesian co-ordinates, then the corollary (12.50) follows, and can be regarded as an alternative definition of the divergence, independent of the co-ordinate system. In particular, it can be used to find the general expression (12.18) for the divergence in an arbitrary set of orthogonal curvilinear co-ordinates (u1, u2, u3).

To do this, we consider the region bounded by surfaces of constant ui and constant ui + δui as shown in Figure 12.9. The edges AB, AD and AA' are along the orthogonal co-ordinate axes, and so are of approximate length h1δu1, h2δu2 and h3δu3, where hi are the coefficients defined in (11.30b). We first calculate the contribution to the integral

numbered Display Equation

from the faces ABCD and A'B'C'D'. If V1, V2, V3 are the components of V along u1, u2, u3, then the contribution from the face ABCD is approximately

numbered Display Equation

evaluated at u3, while the contribution from A'B'C'D' is approximately

numbered Display Equation

evaluated at u3 + δu3, where terms of third order in δui have been neglected. Applying the Taylor series to h1h2V3 at fixed u3 and neglecting terms of order (δu3)2 gives

numbered Display Equation

so that the net contribution from these two faces is

numbered Display Equation
images

Figure 12.9 Construction to derive the divergence in orthogonal curvilinear co-ordinates.

Now the volume element is

numbered Display Equation

to the same order and so from (12.50) the contribution to the divergence is

numbered Display Equation

on taking the limit . Contributions from other pairs of faces may be found in a similar way and putting these together yields

numbered Display Equation

which is the required result (12.18). We leave it as an exercise for the reader to show that the corresponding result (12.19) for the Laplacian follows from combining this result with (12.17) for the gradient, and that the corresponding results for cylindrical and spherical spherical co-ordinates given in Table 12.2 follow on substituting the appropriate values for h1,  h2 and h3.

*12.3.3 Poisson's equation and Gauss' theorem

The electrostatic field E obeys the fundamental equation

(12.53) Unnumbered Display Equation

where ρ is the electric charge density and the constant ϵ0 is the electric permittivity of free space. This equation is called Poisson's equation. Since is the flux of E per unit volume away from the point at which it is evaluated, the interpretation of Poisson's equation is that the electric charge is the source of the electrostatic field.

If we now apply the divergence theorem (12.49) to the field E, and use (12.53), we immediately obtain

(12.54) Unnumbered Display Equation

This relation is called Gauss' theorem. It says that the electric flux through a closed surface S is equal to ϵ− 10 times the total charge enclosed by the surface.

Gauss' theorem is useful in that it allows the field due to a given charge distribution ρ(r) to be evaluated relatively easily in cases where there is a high degree of symmetry. For example, let us suppose that we have a charged sphere centred at the origin with radius R and total charge Q, and that the charge density within the sphere is also spherically symmetric, that is, ρ(r) = ρ(r). Then the resulting field must also be spherically symmetric, that is, it must be of the form

(12.55) Unnumbered Display Equation

so that E(r) points away from (or towards) the origin and its magnitude is the same in all directions. Hence if we choose the surface S to be a sphere of radius r > R, as shown in Figure 12.10, then E(r) is perpendicular to S and

numbered Display Equation

by Gauss' theorem. Consequently,

(12.56) Unnumbered Display Equation

which reduces to Coulomb's law

(12.57) Unnumbered Display Equation

for a point charge at the origin if we allow R → 0 at fixed Q.

images

Figure 12.10 The spherical surface S used to calculate the electric field due to a spherical charge distribution of radius R for r > R.

This analysis is easily generalised to other inverse square law forces. In particular, if g is the gravitational field, so that the force on a point particle of mass is F = mg, then g obeys the Poisson equation

(12.58) Unnumbered Display Equation

where ρ is the mass density and G is the gravitational constant. The result corresponding to (12.56) for a spherically symmetric sphere of total mass M is

numbered Display Equation

which reduces to

numbered Display Equation

when R → 0 at fixed M. This is basis of the approximation that the Earth may be treated as if all its mass were concentrated at its centre when calculating its gravitational field for r > R. However, the approximation is not exact, because the earth is flattened at the poles.

Finally, we note that both the electrostatic and gravitational fields are conservative, satisfying

numbered Display Equation

so that we can introduce scalar potentials φ and ψ by

(12.59) Unnumbered Display Equation

in accordance with the discussion of Section 12.2.2. For the electrostatic case, substituting (12.59) into (12.53) gives

(12.60) Unnumbered Display Equation

which is Poisson's equation for the electrostatic potential; and if one requires φ → 0 as r → ∞, one easily shows that the potential corresponding to (12.57) is the familiar Coulomb potential

(12.61) Unnumbered Display Equation

images

Figure 12.11 The infinitesimal cylinder used in Example 12.14, where S is the surface of the conductor and E = 0 within the conductor.

*12.3.4 The continuity equation

Let us consider a fluid of density ρ(r, t) with a velocity field v(r, t) at time t. Then if we consider a surface element ds, which may lie within the body of the fluid, the mass of liquid passing through ds in unit time is j · ds, where j = ρv is the current vector. If mass is conserved, the rate of change of the mass

numbered Display Equation

contained in a given region Ω must be balanced by the rate at which mass flows out through the surface S bounding Ω, i.e.

(12.62) Unnumbered Display Equation

Equation (12.62) is the statement of mass conservation in integral form. However, it is often more convenient to express it in differential, or local, form, that is, one that refers only to quantities at a single point in space. This can be achieved by using the divergence theorem on the right-hand side of (12.62) and taking the derivative inside the integral on the left-hand side to give

numbered Display Equation

Since this must hold for any region Ω, we must have

(12.63) Unnumbered Display Equation

at any point in space.

Equation (12.63) is called the equation of continuity and is the statement of mass conservation in differential, or local, form. Furthermore, any ρ(r, t) that satisfies a relation of the form (12.63), whatever the relation between the density ρ and the current j, is the density of a conserved quantity. This is because the argument can be reversed, that is, (12.62) follows from (12.63) using the divergence theorem. Then, if we let the surface S recede to infinity, we obtain

(12.64) Unnumbered Display Equation

where the integral extends over all space, provided ρ,  j → 0 sufficiently rapidly at infinity, as they usually do. Many examples of conserved quantities occur in physics, including electric charge and energy. However, the relation between the density ρ and the current j is not always as simple as j = ρv, as shown in Example 12.15.

12.4 Stokes' theorem

Given a closed contour C, spanned by a surface S, and a vector field V defined on S, then Stokes' theorem states that

(12.68) Unnumbered Display Equation

where the sense of the vector element ds is given by a right-handed screw rule with respect to the direction of integration around C.

The line integral on the right-hand side of (12.68) is called the circulation of V around the loop C. Thus the theorem states that the surface integral of is equal to the circulation of V around the bounding curve C. This is closely related to the interpretation of curl. To see this, we apply (12.68) to a loop C that encloses a small surface element , which shrinks to a point when ds → 0. In this limit, the variation of V and can be neglected on ds, so that the left-hand side of (12.68) becomes , implying

(12.69) Unnumbered Display Equation

In other words, at a point r is the circulation per unit area around the boundary of an infinitesimal surface ds containing the point r. For example, let us again consider a vector field V = ρv, where ρ is the density and v is the velocity field of a fluid. Then for a uniform flow pattern, such as that shown in Figure 12.12a, and V is said to be irrotational. On the other hand, at the centre of a vortex, like that shown in Figure 12.12b, clearly . It is also non-zero in a non-uniform parallel motion, as shown in Figure 12.12c, since the velocities on either side of a point are different. Essentially, when there is rotational motion in addition to, or opposed to, translational motion. A practical viewpoint is to consider what would happen if one inserted a small ‘paddle wheel’, which is free to rotate about its axis. In the flow pattern of Figure 12.12a, where , it would not rotate: the motion is irrotational. In Figures 12.12b and 12.12c, where , it would rotate.

images

Figure 12.12 Flow of a fluid. The coloured lines are the flow lines; the arrows show the direction of the vector field V. Their lengths show the relative magnitudes of V.

In the rest of this section we will first derive Stokes' theorem, and then consider some applications.

12.4.1 Proof of Stokes' theorem

We start by considering a closed curve C surrounding a plane surface S parallel to the x–y plane, so that z is constant. Then

numbered Display Equation

and

numbered Display Equation

But by Green's theorem in the plane (11.20), we have

numbered Display Equation

so that

(12.70) Unnumbered Display Equation

where , a unit vector in the z-direction. Furthermore, in the limit ds → 0, where the variation of over the surface can be neglected, we obtain

(12.71) Unnumbered Display Equation

As there is nothing special about the z-direction – we may choose it in any direction we like – it follows that (12.70) and (12.71) hold for any finite or infinitesimal planar surface, respectively, where is the normal defined in the usual sense. We will now use this result to derive Stokes' theorem.

Consider an open surface, which must be two-sided, divided into small regions ds, as shown in Figure 12.13a. As ds0, each element, irrespective of its shape, approaches ever more closely to an element of the plane tangential to the surface at the centre of the surface ds. Therefore, (12.71) implies (12.69), and (12.70) becomes

numbered Display Equation

as ds → 0, where the circulation is around the boundary of ds. If we sum over all ds,

numbered Display Equation

and from the enlarged section shown in Figure 12.13b it is clear that all interior contributions to the circulation will vanish, resulting in

numbered Display Equation
images

Figure 12.13 Construction to derive Stokes' theorem.

This is Stokes' theorem as required. It is worth emphasising that the right-hand side is an integral over any surface that is bounded by the curve C. Note also the direction of the circulation, which is ‘right-handed’ relative to the directions ds, as discussed in Section 12.2.3 [cf. Figure 12.6]. In the following subsections we will consider some applications of this theorem.

*12.4.2 Curl in curvilinear co-ordinates

Having derived Stokes' theorem (12.68) and its corollary (12.69), the latter can be regarded as an alternative definition of curl, independent of the co-ordinate system. Here we shall use it to obtain the expression for curl in an arbitrary system of orthogonal linear co-ordinates.7

To do this, we consider the infinitesimal surface element ds = ds e3 swept out when u1u1 + du1 and u2u2 + du2 at constant u3, as shown in Figure 12.14. Then from (12.69), we have

(12.72) Unnumbered Display Equation

where C is the contour ABCD shown in Figure 12.4. We now write

numbered Display Equation

where e1 and e2 are unit vectors along the directions AB and AD, respectively, and use the fact that in the limit that du1,  du2 tend to zero the corresponding lines may be approximated by straight lines, and ABCD may be approximated by a rectangle, since e1 and e2 are orthogonal. Hence the contribution of V1 to the line integral arises solely from the arcs AB and DC and is

numbered Display Equation
images

Figure 12.14 Construction to derive Stokes' theorem in curvilinear co-ordinates.

Similarly, the contribution from V2 is

numbered Display Equation

and ds = h1h2u1u2, so that on substituting into (12.72) we obtain

numbered Display Equation

This identical to the e3 component given in (12.20) and analogous results follows for the other components.

*12.4.3 Applications to electromagnetic fields

Finally we illustrate the use of Stokes' theorem by applying it to the behaviour of electric and magnetic fields, starting with the electric field E. In free space, this is determined by the fundamental equations

(12.73) Unnumbered Display Equation

where B is the magnetic field intensity, ρ is the charge density, and ϵ0 is the electric permittivity. Of these, (12.73a) was discussed in Section 12.3.3, where we saw that it expressed the fact that charge is the source of electric flux. However, in contrast to electrostatics, if there are time-dependent magnetic fields present, no longer vanishes. Hence E is not in general a conservative field and loop integrals of the form

numbered Display Equation

no longer vanish. Rather, by Stokes' theorem and (12.73), we have

(12.74) Unnumbered Display Equation

where S is any open surface spanning the loop C. This is Faraday's law of induction that states that the ‘emf’ ϵC induced around a loop C is equal to minus the rate of change of the magnetic flux through the loop. We also note that the argument can be reversed: if (12.74) holds, then Stokes' theorem gives

numbered Display Equation

which can only hold for an arbitrary open surface S if (12.74) is satisfied. Equation (12.73b) and (12.74) are the differential and integral forms of Faraday's law.

Equations (12.73a) and (12.73b) are the first two Maxwell's equations in free space. The remaining two are

(12.75) Unnumbered Display Equation

where j is the electric current density, μ0 is the magnetic permeability of free space, and the speed of light c = (μ0ϵ0)− 1/2. On comparing with (12.73a), we see that (12.75a) reflects the experimental observation that there are no free magnetic charges. The second equation (12.75b) indicates that non-zero magnetic fields can be generated by currents or time-dependent electric fields. In the absence of the latter, it becomes

(12.76) Unnumbered Display Equation

By Stokes' theorem

numbered Display Equation

giving

(12.77) Unnumbered Display Equation

where S is any surface spanning the loop C. This is called Ampère's law and it states that the line integral of B around a closed loop is equal to μ0 times the total current Iencl flowing through the loop. It enables the magnetic field to be calculated quickly in symmetrical situations, as we shall illustrate.

images

Figure 12.15 The circuit C used to derive (12.79), where the current I at the centre is directed out of the page.

Problems 12

  1.   12.1 A scalar field electrostatic potential is given by φ = x2y2 and the associated electric field E is given by . What is the magnitude and direction of E at (2, 1)? In what direction does φ increase most rapidly at the point (−3, 2) and what is the rate of change of φ at the point (1, 2) in the direction 3ij?

  2.   12.2 Given the scalar function ψ = x2y2z, find (a) at (1, 1, 1); (b) the derivative of ψ at (1, 1, 1) in the direction i − 2j + k; (c) the equation of the normal to the surface ψ = x2y2z = 0 at (1, 1, 1).

  3.   12.3 If A = 2xz2iyz j + 3xz3k and S = x2yz, find in Cartesian co-ordinates (a) curl A, (b) , (c) , (d) and (e) .

  4.   12.4 Given a scalar field ψ and a vector field V, show (a) that

    numbered Display Equation

    Hence show (b) that if , where α is a scalar field, then

    numbered Display Equation
  5.   12.5 Show, without explicitly writing out the components, that

    numbered Display Equation
  6.   12.6 If ψ = 2yz and , express

    numbered Display Equation

    in spherical polar co-ordinates.

  7.   12.7 Directly evaluate the line integral

    numbered Display Equation

    where

    numbered Display Equation

    around the circle (x2 + y2) = a2 in the x–y plane. Verify your result using Green's theorem in the plane.

  8.   12.8 A force field is

    numbered Display Equation

    Find the work done in moving a particle round a closed curve from the origin to the point (x, y, z) = (0, 0, 2π) along the path

    numbered Display Equation

    and then back to the origin along the z-axis.

  9.   12.9 Find the work done by a force F given by

    numbered Display Equation

    when moving a particle clockwise along a semicircle of unit radius in the x–y plane from x = −1 to x = 1 with y ≥ 0.

  10.  12.10 Find the work done by a force F = (x2 + y2)j when moving between the points A(x = a,  y = 0,   z = 0) and B(x = 0,  y = a,  z = 0) along a path C, where C is (a) along the x-axis to the origin, then along the y-axis to B, and (b) along the arc of the circle x2 + y2 = a2,   z = 0, in the positive quadrant.

  11.  12.11 A force moves around a closed loop starting at the origin along the curve to (2, 1), then parallel to the x axis to (0, 1) and finally returning to the origin along the y axis. Use Green's theorem in the plane to calculate the work done by the force.

  12.  12.12 Show that

    numbered Display Equation

    is a conservative field and find a scalar potential φ, such that .

  13.  12.13 A force field

    numbered Display Equation

    where a, b, c, are constants. For what values of a, b, c, is F a conservative field? Find the scalar potential in this case.

  14.  12.14 Let be that part of the surface of the cylinder

    numbered Display Equation

    for which x > 0,   y > 0. What is the value of the surface integral

    numbered Display Equation

    if A = 6yi + (2x + z)jxk and S is that part of that lies on the curved surface of the cylinder?

  15.  12.15 Evaluate the integral

    numbered Display Equation

    where and S is the surface of a unit cube 0 ≤ x ≤ 1,  0 ≤ y ≤ 1, 0 ≤ z ≤ 1 , without using the divergence theorem. The vector s is defined in the outward direction from each face of the cube.

  16.  12.16 Evaluate the integral

    numbered Display Equation

    over the curved surface of a hemisphere of radius a with its centre at the origin and base in the x–y plane, where ds = |ds | and σ is a constant.

  17.  12.17 A sphere of uniform density ρ has mass M and radius a. Calculate its moment of inertia about (a) a tangent to the sphere and (b) an axis through the centre of the sphere.

  18.  12.18 A cylinder of uniform density ρ0 has mass M, radius a and length d. Calculate its moment of inertia about an axis that lies in the plane of the base of the cylinder and passes through the centre of the base.

  19.  12.19 Use the divergence theorem to evaluate

    numbered Display Equation

    where F = 4xz iy2 j + yx k and S is the surface of the cube bounded by

    numbered Display Equation
  20.  12.20 Scalar fields φi(i = 1, 2, …) are solutions of the equations

    numbered Display Equation

    where the γi are constants, within a region Ω, subject to the boundary conditions φi = 0 on the closed surface S enclosing Ω. Show that

    numbered Display Equation

    if ij γi ≠ γj.

  21.  12.21 Prove the identity

    numbered Display Equation

    for any scalar field ψ. A scalar field ψ satisfies the conditions ψ = 0 on S and in Ω, where S is the closed surface surrounding the region Ω. Show that ψ = 0 in Ω.

  22. *12.22 State Gauss' theorem for the gravitational field. A homogeneous spherical shell has mass M, inner radius a and outer radius b > a. Find an expression for the gravitational field due to the shell for (a) r > b, (b) r < a and (c) a < r < b, where r = |r |. Finally, calculate the potential at any point with r < a assuming the potential goes to zero as r → ∞.

  23. *12.23
    1. Prove the relation
      numbered Display Equation
      where φ is a scalar field and E is a vector field.
    2. Let ρ(r) be an electric charge density, which vanishes outside a finite region Ω1 enclosing the origin, with total charge

      numbered Display Equation

      Write down an approximate value for the electrostatic field E and potential φ on a sphere centred at the origin with radius R, assuming that R is very large compared to the dimensions of Ω1.

    3. Show that, in the same approximation,

      numbered Display Equation

      where Ω2 is the interior of the sphere of radius R, and find the value of the constant c. [Hint: use Poisson's equation .]

  24. *12.24 In a homogeneous continuous medium, Maxwell's equations take the form

    numbered Display Equation

    with , and the constants ϵ and μ are the permittivity and permeability of the medium.

    1. Show that Maxwell's equations imply that ρ is the density of a conserved charge.
    2. In a conductor, the current obeys Ohm's law j = σE, where σ is the conductivity. Find the charge density ρ as a function of time if ρ(r, t = 0) = ρ0(r).
  25.  12.25 Verify Stokes' theorem for the vector

    numbered Display Equation

    where S is the surface of the hemisphere

    numbered Display Equation

    and C is the boundary of S.

  26.  12.26 Use Stokes' theorem (12.68) to prove the relation

    numbered Display Equation

    where φ is a scalar fields and S is an open surface bounded by a closed curve C. [Hint: apply Stokes' theorem to the vector field V = φ c, where c is a constant vector.]

  27.  12.27 A force field F = y2i + x2j acts on a particle. Write down a line integral corresponding to the work done by the field when the particle moves once round the circle x2 + y2 = a2,  z = 0 in the anticlockwise direction. Evaluate this integral (a) directly and (b) by converting it to a surface integral. (c) Is the force field F conservative?

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset