Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

39
Unit 1 Relevant Electrostatics and Magnetostatics (Old and New) The whole of classical electrodynamics is encompassed by a set of coupled partial differential equations (at least in one form) bearing the name “Maxwell’s Equations”. On the basis of material already encountered and some new concepts, it is the purpose of this course to develop these equations and to make some elementary applications. To facilitate these intentions, we shall briefly review a little of the prerequisite elec- trostatic and magnetostatic material and, in so doing, revisit some of the necessary mathematics. It should be remembered that much of what now may be written as concise mathematical formalism was, in fact, often developed initially by painstaking experimental investigation. We shall seek to not stray too far from the physical sig- nificance and practicalities of the results “derived”. For the level and length of this course, it will not be possible to explore in detail the scope of the usefulness of the equations considered. In fact, in all but the simplest cases, solutions in closed form do not exist. It will quickly become apparent, especially in future studies based on this course, that numerical methods and approximation techniques are essential in most engineering applications of Maxwell’s equations. Still, there will be plenty of “interesting” problems which may be tackled with the arsenal of tools developed in this and earlier terms. 1

Transcript of Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Page 1: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Unit 1

Relevant Electrostatics andMagnetostatics (Old and New)

The whole of classical electrodynamics is encompassed by a set of coupled partial

differential equations (at least in one form) bearing the name “Maxwell’s Equations”.

On the basis of material already encountered and some new concepts, it is the purpose

of this course to develop these equations and to make some elementary applications.

To facilitate these intentions, we shall briefly review a little of the prerequisite elec-

trostatic and magnetostatic material and, in so doing, revisit some of the necessary

mathematics. It should be remembered that much of what now may be written as

concise mathematical formalism was, in fact, often developed initially by painstaking

experimental investigation. We shall seek to not stray too far from the physical sig-

nificance and practicalities of the results “derived”. For the level and length of this

course, it will not be possible to explore in detail the scope of the usefulness of the

equations considered. In fact, in all but the simplest cases, solutions in closed form

do not exist. It will quickly become apparent, especially in future studies based on

this course, that numerical methods and approximation techniques are essential in

most engineering applications of Maxwell’s equations. Still, there will be plenty of

“interesting” problems which may be tackled with the arsenal of tools developed in

this and earlier terms.

1

Page 2: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

A Few Preliminary Observations on Notation

(1) We shall use �� to represent the position vector in space for some observation point

(relative to an origin) and ��′ to indicate the position of sources (relative to an origin).

(2) Based on note (1), �� − ��′ will thus be the displacement vector from the source

to the observation position. This vector will assume various labels throughout the

course (eg., 𝑅, 𝑅12, etc.) as may be convenient.

Illustration:

(3) The text uses bold letters to represent vector quantities. We shall use vector signs

or ⇀ (eg., �� or⇀

𝐸≡ electric field intensity). We’ll use ˆ to indicate unit vectors (eg.,

�� ≡ unit vector in the 𝑥 direction).

Reference should be made to the text for various coordinate systems which may

be encountered. Also, the vector operators, divergence, gradient and curl (∇⋅, ∇ and

∇×) are found at the back of the text, while various vector operator identities are

located in Appendix A.3.

1.1 Electrostatics Review

For the sake of completeness and for future reference, we shall write down the following

previously encountered results.

Coulomb’s Law:

With reference to the accompanying diagram, the force, ��21, on charge 𝑄2 due to

charge 𝑄1 is given by

2

Page 3: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

��21 = −��12 =

= (1.1)

where we have assumed that the charges are in free space and 𝜖0 (=8.85×10−12 F/m)

is the permittivity of free space. The definition of unit vector is obvious.

Electric Field Intensity:

The electric field intensity, ��, is by definition the force per unit charge on a small

positive test charge, 𝑞𝑡, brought into the field and is in the direction of the force; i.e.

�� =𝐹

𝑞𝑡(1.2)

The unit is N/C or, more conveniently, volt/metre (V/M). For a volume charge

distribution with charge density 𝜌𝑣(��′),

��(��) = (1.3)

Note that 𝑑𝑣′ is the elemental volume at ��′. Clearly, the integral sums the effects of

all the charges of the form 𝜌𝑣(��′)𝑑𝑣′. In Cartesian coordinates, 𝑑𝑣′ = 𝑑𝑥′𝑑𝑦′𝑑𝑧′. This

integral can be “nasty” and is usually explicitly carried out only for simple cases.

(There are often better ways of finding ��, eg., Gauss’ Law).

Electric Flux Density and Gauss’ Law:

Gauss’ Law states that the electric flux, Ψ, measured in coulombs (C), passing

through any closed surface is equal to the total charge enclosed by that surface.

The electric flux density, ��, is measured in C/m2.

Ψ = 𝑄 =∮𝑆�� ⋅ 𝑑�� =

∫vol

𝜌𝑣𝑑𝑣 (1.4)

3

Page 4: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Note that

𝑑�� = 𝑑𝑆𝑆

where 𝑆 is the outward unit normal to the elemental surface 𝑑𝑆. For example, in

cartesian coordinates, 𝑑𝑆𝑧 = 𝑑𝑥𝑑𝑦.

We also note a so-called constitutive relation between �� and ��. In general, for an

isotropic space with permittivity 𝜖,

�� = 𝜖�� , (1.5)

and for free space, where 𝜖 = 𝜖0 it is

�� = 𝜖0�� .

Gauss’ Integral Theorem (or the Divergence Theorem):

This theorem states that

∮𝑆�� ⋅ 𝑑�� =

∫vol

∇ ⋅ ��𝑑𝑣 (1.6)

∇ ⋅ �� is the divergence of �� throughout the volume. Equation (1.6) is a general

statement for vector fields, not just ��-fields. Note that (1.4) and (1.6) give

∇ ⋅ �� = 𝜌𝑣 (1.7)

4

Page 5: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

which happens to be the “point form” of one of Maxwell’s equations alluded to earlier.

Aside: Recall

Equation (1.7) holds for time-varying fields also, and it will be used in this context

later in the course.

Work, 𝑊 , Done on a Charge, 𝑄, Moved in an ��–Field:

𝑊 = −𝑄∫ 𝐴

𝐵�� ⋅ 𝑑�� (1.8)

where 𝑑�� is a differential displacement vector along the path, 𝐶, from 𝐵 to 𝐴 as

illustrated.

Not surprisingly, only the component of �� along the path contributes to energy

expenditure or work done.

Potential Difference, 𝑉𝐴𝐵, and Potential at a Point and Potential Gradient:

From equation (1.8), we define the potential difference between points 𝐴 and 𝐵

as

𝑉𝐴𝐵 =𝑊

𝑄= −

∫ 𝐴

𝐵�� ⋅ 𝑑�� = 𝑉𝐴 − 𝑉𝐵 (1.9)

where 𝑉𝐴 ≡ potential at point 𝐴, 𝑉𝐵 ≡ potential at point 𝐵. The potential at a

point is the potential difference measured with respect to some arbitrarily chosen

zero reference point. For example, if 𝑉∞ = 0, then 𝑉𝐴 is the work done per unit

charge in moving the charge from ∞ to point 𝐴.

5

Page 6: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Conservative Field:

Note that, by definition, a conservative field is one which satifies the condition “

a closed line integral of the field is zero”. For example, the electrostatic field, ��, is

conservative so ∮𝐶�� ⋅ 𝑑�� = 0 (1.10)

Note that the path is closed. Equation (1.10) is not valid for time-varying fields –

more later!!

Also, recall, importantly,

�� = −∇𝑉 (1.11)

which provides a convenient way of finding the (vector) electric field, ��, if the (scalar)

potential field, 𝑉 , is known.

Continuity of Current:

Consider a current density, 𝐽 , measured in amperes per square metre (A/m2). The

principle of charge conservation leads to the point form of the equation of continuity

of current as

∇ ⋅ 𝐽 = −∂𝜌𝑣∂𝑡

(1.12)

Boundary Conditions:

(1) Perfect Conductor

Recall that within a perfect conductor no free charge and no electric field may

exist. Using this fact, it is easily argued from equation (1.10) that at the surface

between a perfect conductor and free space

𝐸𝑡 = 0 =⇒ 𝐷𝑡 = 0

and (1.13)

𝐷𝑁 = 𝜌𝑠 (or 𝐸𝑁 =𝜌𝑠𝜖0)

6

Page 7: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

where 𝑡 ≡ tangential, 𝑁 ≡ normal, and 𝜌𝑠 is a surface charge density (charge may

exist on the conductor surface).

(2) Perfect Dielectric

Now, in general, �� may be non-zero on both sides of the boundary. Again, arguing

from equation (1.10), it may be verified that for the case at hand

𝐸𝑡1 = 𝐸𝑡2

and (1.14)

𝐷𝑁1 = 𝐷𝑁2 (or 𝐸𝑁1 =𝜖2𝜖1

𝐸𝑁2)

with 𝜖1 and 𝜖2 being the material permittivities.

Poisson’s and Laplace’s Equations:

Finally, in the electrostatics portion of the previous course, the following important

results were derived from equations (1.7), (1.5), and (1.11):

(1) Poisson’s Equation:

∇ ⋅ ∇𝑉 = −𝜌𝑣𝜖

(1.15)

where 𝜌𝑣 is the usual volume charge density and 𝜖 is the permittivity of the region

under consideration.

(2) Laplace’s Equation:

A special case of (1.15) occurs in a region for which 𝜌𝑣 = 0. The result, known as

Laplace’s equation is clearly given by

∇ ⋅ ∇𝑉 = 0

commonly abbreviated as

∇2𝑉 = 0 (1.16)

At this point it would be a good idea to review a few relevant electrostatic

examples encountered in Engineering 5812.

7

Page 8: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

1.2 Magnetostatics Review

We have seen that static electric charges give rise to static electric fields. We have

also defined direct current (dc current) to be constituted of charges which flow with

a constant speed in one direction. These dc currents are a source of steady magnetic

fields. We shall have opportunity in the next course to see how time-varying magnetic

fields may be generated.

In this unit we will introduce the magnetic field intensity, �� , which is measured

in amperes per metre (A/m). There will be analogies observed between the equations

involving the electric field intensity, ��, and those involving ��. There is one significant

difference: while eletric fields may be generated by point charges, there are no known

point magnetic ‘charges’ – although to make calculations easier we sometimes assume

they exist (but NOT in this course)!

The Biot-Savart Law:

The Biot-Savart law is the magnetic field counterpart (sort of) to Coulomb’s law.

This law, which relates the magnetic field intensity, �� , in amperes per metre (A/m)

to a source current, 𝐼, may be written, with reference to the diagram which follows,

in differential form as

𝑑�� =𝐼𝑑��′ × ��

4𝜋𝑅2=

𝐼𝑑��′ × ��

4𝜋𝑅3

Illustration:

Clearly, from the cross product, the �� field element (𝑑��) produced by the current,

𝐼, flowing along the differential displacement, 𝑑��′, of the path, 𝐶, is perpendicular to

the plane containing 𝑑��′ and the unit vector �� from 𝑑��′ (i.e. the source region) to

the point of observation of the �� field. 𝑅 is simply the distance from 𝑑��′ to the place

8

Page 9: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

where �� is observed. The right hand rule for the cross product gives the orientation

of the contribution, 𝑑��, due to the current in element 𝑑��′.

A current element may not be isolated to check the law in its differential form, but

considering 𝐶 to be a closed path we may write an experimentally verifiable integral

form as

(1.17)

Equation (1.17) may also be written for surface current densities �� in A/m or

(volume) current density 𝐽 in A/m2 as

(1.18)

Illustration:

or

(1.19)

To emphasize and illustrate the importance of coordinate system representation of

these expressions consider (1.19):

Illustration:

9

Page 10: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

In view of the above illustration, (1.19) becomes

(1.20)

Again, note that the primed coordinates represent the source position and the un-

primed coordinates represent the field (observation) position. Comparing equation

(1.20) (i.e. the Biot-Savart law for the magnetic field) with equation (1.3) (i.e.

Coulomb’s law for the electric field), the similarities and differences are obvious: (1)

Both are inverse square laws, but (2) �� is in the direction of �� = (��− �� ′)/∣��− �� ′∣while �� is perpendicular to both the source vector, 𝐽 , and ��.

Ampere’s Circuital Law:

An analogy to Gauss’ law for electrostatics is Ampere’s circuital law for magne-

tostatics. Instead of “talking about” electric flux density and charge enclosed by a

surface, 𝑆, we consider a steady magnetic field intensity and a current, 𝐼𝑒, enclosed

by a contour, 𝐶. Ampere’s law states that the line integral of �� around any closed

path is exactly equal to the direct current, 𝐼𝑒, enclosed by the path. That is

∮𝐶�� ⋅ 𝑑�� = 𝐼𝑒 . (1.21)

Illustrations: (1) Current Filament:

10

Page 11: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

(2) “Thick” dc-Current-Carrying Wire:

Note, in general, that

𝐼𝑒 =∫𝑆𝐽 ⋅ 𝑑�� (1.22)

(Similar ideas exist for surface current density, ��).

If it is possible to choose a contour with the properties (1) �� is everywhere tan-

gential to the contour and (2) ∣��∣ is constant along the contour, equation (1.21)

(Ampere’s circuital law) may be used to find �� when 𝐼𝑒 is known. For this purpose,

the nature of �� may be observed from the Biot-Savart law (without actually doing

the integration) and symmetry arguments. For example, in the previous course, we

used symmetry and the Biot-Savart law to show that, for an infinitely long current

filament along the 𝑧-axis, the field had the form

�� = 𝐻(𝜌)𝜙

(that is, the �� field is a FUNCTION of 𝜌 and has only a 𝜙 COMPONENT) and that

a contour that led to a simple integration was a circle centred on the 𝑧-axis.

At this point it would be a good idea to review the magnetostatic examples

encountered in Engineering 5812.

11

Page 12: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Another Ampere’s Law Example:

Consider the situation depicted in the illustration associated with surface currents

in relation to equation (1.18). A similar example is addressed in the text on pages

222–223. Suppose that the surface current density is given by �� = 𝐾𝑥�� with 𝐾𝑥 > 0

in the 𝑧 = 0 plane. Let’s find the magnetic field intensity due to this source.

Illustration

(1) We’ll use the Biot-Savart law to establish the field directions: We may

think of the surface current as being composed of infinitely many current filaments

carrying current in the �� direction. From the Biot-Savart law, we know that each of

these filaments produces a magnetic field intensity which follows circular streamlines

with the filaments being at the centre of the circles. It may be easily argued, that

in this case, any two filaments which are symmetrical about a line parallel to and

half-way between them will produce fields whose 𝑧-components cancel, but whose 𝑦

components remain. In addition to this, there is no field component along the line of

the current (again the cross product in the Biot-Savart law ensures this) – i.e. there is

no field component in the ±�� direction. We have thus established that �� = 𝐾𝑥�� has

only components in the ±𝑦 directions, the negative 𝑦 direction being associated with

fields in the halfspace above the current surface and the positive direction associated

with fields below the surface – see illustration on the right.

(2) What about the coordinate dependencies? The current covers the entire

12

Page 13: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

𝑧 = 0 plane. Thus, at any point above or below the plane, the field will not depend

on 𝑥 or 𝑦. (In fact, we will see that it doesn’t depend on 𝑧 either, but we will have

to prove this.)

We are now ready to apply Ampere’s law. Since on 𝐻𝑦 components are present

let’s choose an integration contour (or path) whose ‘segments’ are either parallel to

or perpendicular to 𝑦. This contour is labelled above as rectangle 𝑎𝑏𝑐𝑑𝑎. Direct

application of ∮𝐶�� ⋅ 𝑑�� = 𝐼𝑒

gives

Similarly, if we use path 𝑎′𝑏′𝑐𝑑𝑎′ we get

Thus, at any point 𝑧 > 0 above the sheet, �� = 𝐻𝑦𝑦, and symmetry, we can im-

mediately see that at any point 𝑧 < 0 below the sheet, �� = −𝐻𝑦𝑦. Thus, from

∗,−𝐻𝑦 −𝐻𝑦 = 𝐾𝑥

and

which implies below the sheet �� = −𝐻𝑦𝑦 = .

Note that using �� as the outward pointing unit normal, �� = 12�� × �� for all 𝑧 ∕= 0.

What would happen to the field structure if another sheet with �� = −𝐾𝑥�� was

placed at position 𝑧 = 𝑧0?

13

Page 14: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

1.3 More Magnetostatics

1.3.1 More on Ampere’s Law

The Curl

The del operator, ∇, may also be “crossed” into a general vector, ��. The sym-

bolism is ∇ × �� and the operation is termed “the curl of ��”.

Definition: The curl of a vector function �� (i.e.��(𝑥, 𝑦, 𝑧) in cartesian coordinates) is

defined as

∇ × �� =

(��

∂𝑥+ 𝑦

∂𝑦+ 𝑧

∂𝑧

)× (𝐴𝑥��+ 𝐴𝑦𝑦 + 𝐴𝑧𝑧) .

Therefore, the vector which results is given by

∇ × �� = ��

(∂𝐴𝑧

∂𝑦− ∂𝐴𝑦

∂𝑧

)+ 𝑦

(∂𝐴𝑥

∂𝑧− ∂𝐴𝑧

∂𝑥

)+ 𝑧

(∂𝐴𝑦

∂𝑥− ∂𝐴𝑥

∂𝑦

)

=

∣∣∣∣∣∣∣∣∣

�� 𝑦 𝑧∂

∂𝑥

∂𝑦

∂𝑧𝐴𝑥 𝐴𝑦 𝐴𝑧

∣∣∣∣∣∣∣∣∣(1.23)

In expanding this determinant, the derivative nature of ∇ must be considered. The

determinant in equation (1.23) must be expanded from the top down so that we get

the derivatives shown in the second row. It might be noted that expressions such as

��× ∇ are defined only as differential operators and the normal rules which apply to

the cross products of ordinary vectors do NOT apply here. For example, in general,

��× ∇ ∕= −∇ × ��.

Notice that the curl of a vector is a vector! As with the previous vector operators

encountered, (1.23) will not have such a simple form in cylindrical and spherical

coordinates.

Illustration: �� = 𝑥��+ 𝑦𝑦 + 𝑧𝑧 .

The ∇ × ��, sometimes denoted as curl ��, is

∇ × �� =

14

Page 15: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Interpretation

Some of the finer mathematical details are not included in the following discussion.

However, the basic ideas required for interpretation of the curl are discussed. For

the sake of simplicity, consider the �� component of ∇ × ��, whose magnitude from

equation (1.23) is clearly given by

(∇ × ��

)⋅ �� =

∂𝐴𝑧

∂𝑦− ∂𝐴𝑦

∂𝑧.

With reference to the diagram below, which is a small (i.e. differential) rectangle on

the 𝑦-𝑧 plane, the definition of partial differentiation leads us to

∂𝐴𝑧

∂𝑦= lim

Δ𝑦→0

(𝐴𝑧 on b)− (𝐴𝑧 on d)

Δ𝑦

and

∂𝐴𝑦

∂𝑧= lim

Δ𝑧→0

(𝐴𝑦 on c)− (𝐴𝑦 on a)

Δ𝑧

Thus, getting a common denominator gives

(∇ × ��

)⋅�� = lim

Δ𝑦,Δ𝑧→0

(𝐴𝑦 on a)Δ𝑦 + (𝐴𝑧 on b)Δ𝑧 + (−𝐴𝑦 on c)Δ𝑦 − (𝐴𝑧 on d)Δ𝑧

Δ𝑦Δ𝑧

Using the limits we may legitimately write, on recalling from earlier discussions the

idea of the differential surface area, that

(∇ × ��

)⋅ �� = lim

Δ𝑆𝑥→0

∮𝑎𝑏𝑐𝑑

�� ⋅ 𝑑��Δ𝑆𝑥

(1.24)

15

Page 16: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

From this last expression, we see that the curl of �� in the �� direction may be

gotten by traversing the differential rectangle in the 𝑦-𝑧 plane as shown, following

the right-hand rule (i.e. fingers curled in direction of travel with thumb, in this case,

pointing in the 𝑥 direction). Again, it must be emphasized that the rectangle is of

differential proportions and it is assumed that 𝐴𝑦 and 𝐴𝑧 do not vary along a given di-

rection of travel – a more rigorous treatment of these ideas using higher order terms

in a Taylor series expansion would lead properly to the same interpretation. The

above discussion may be summarized as follows:

The “circulation” of the vector field per unit area around a differential area in the

𝑦-𝑧 plane is given by the �� component of ∇ × ��. A similar statement could be made

about the other components.

If we allow for all components, then equation (1.24) may be generalized as

(∇ × ��

)⋅ �� = lim

Δ𝑆→0

∮𝐶Δ𝑆

�� ⋅ 𝑑��Δ𝑆

(1.25)

where �� is the unit normal to the incremental surface, Δ𝑆, and 𝐶Δ𝑆 is the perimeter

of that surface.

To get an intuitive feeling for the curl, consider the following: If we have a paddle

wheel which we dip into some “force” field, the paddle wheel will spin differently

depending on its orientation in the field. If the wheel does not rotate at all, it means

that the curl is zero in the direction of the wheel’s axis. Larger angular “velocities”

of the wheel indicate larger values of the the curl. In order to find the direction of the

vector curl, we hunt around for the orientation that provides the greatest “torque”.

The direction of the curl is then along the axis of the wheel, as given by the right-hand

rule.

16

Page 17: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

CURL IN CYLINDRICAL COORDINATES

We state without proof that the curl in cylindrical coordinates is given by

∇ × �� =

(1

𝜌

∂𝐴𝑧

∂𝜙− ∂𝐴𝜙

∂𝑧

)𝜌+

(∂𝐴𝜌

∂𝑧− ∂𝐴𝑧

∂𝜌

)𝜙+

(1

𝜌

∂(𝜌𝐴𝜙)

∂𝜌− 1

𝜌

∂𝐴𝜌

∂𝜙

)𝑧

CURL IN SPHERICAL COORDINATES

We state without proof that the curl in spherical coordinates is given by

∇ × �� =1

𝑟 sin 𝜃

(∂(𝐴𝜙 sin 𝜃)

∂𝜃− ∂𝐴𝜃

∂𝜙

)𝑟 +

1

𝑟

(1

sin 𝜃

∂𝐴𝑟

∂𝜙− ∂(𝑟𝐴𝜙)

∂𝑟

)𝜃 +

1

𝑟

(∂(𝑟𝐴𝜃)

∂𝑟− ∂𝐴𝑟

∂𝜃

)𝜙

Stokes’ Theorem

Consider a surface in space which has been subdivided into incremental pieces as

shown.

Rearranging equation (1.25), for a particular surface increment we get

∮𝐶Δ𝑆

�� ⋅ 𝑑�� =(∇ × ��

)⋅(�� lim

Δ𝑆→0Δ𝑆

)(1.26)

If the line integral here is carried out for every such region in the whole surface, 𝑆,

there will be cancellation along all the cell sides except those which constitute the

outside boundary, 𝐶, of 𝑆. Thus, equation (1.26) becomes, on integrating over the

whole region, ∮𝐶�� ⋅ 𝑑�� =

∫𝑆

(∇ × ��

)⋅ 𝑑�� (1.27)

Here, we have used the fact that limΔ𝑆→0

Δ𝑆 = 𝑑𝑆 and ��𝑑𝑆 = 𝑑��. The very significant

result in equation (1.27) is known as Stoke’s Theorem. The following numerical

example will illustrate the geometry involved.

17

Page 18: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Example: A portion of a spherical surface is specified by 𝑟 = 4, 0 ≤ 𝜃 ≤ 0.1𝜋, 0 ≤𝜙 ≤ 0.3𝜋. The path segments forming the perimeter of this surface are given by

(1)𝑟 = 4, 0 ≤ 𝜃 ≤ 0.1𝜋, 𝜙 = 0, (2) 𝑟 = 4, 𝜃 = 0.1𝜋, 0 ≤ 𝜙 ≤ 0.3𝜋, and (3)

𝑟 = 4, 0 ≤ 𝜃 ≤ 0.1𝜋, 𝜙 = 0.3𝜋. We are given that the magnetic field intensity

is �� = 6𝑟 sin 𝜙 𝑟 + 18𝑟 sin 𝜃 cos𝜙 𝜙 and we desire to evaluate each side of Stokes’

theorem.

Solution:

18

Page 19: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Ampere’s Law and Stokes’ Theorem:

Equations (1.21), (1.22) and (1.27) together give us the means of writing Ampere’s

law in point form:

We had in (1.21) ∮𝐶�� ⋅ 𝑑�� = 𝐼𝑒 .

and, more generally, in (1.22) ∫𝑆𝐽 ⋅ 𝑑�� = 𝐼𝑒 .

In (1.27), if we replace �� by ��, the magnetic field intensity, we get another integral

form of Ampere’s law

Recall that the path 𝐶 surrounds the surface 𝑆. Comparing the righthand equality in

this equation with equation (1.22), we immediately have the point form of Ampere’s

law as

∇ × �� = 𝐽 (1.28)

[ASIDE: Recall that for electrostatics we had∮𝐶 �� ⋅𝑑�� = 0 (for any closed 𝐶) which,

from Stokes’ theorem, leads to ∇ × �� = 0 – this is definitely NOT TRUE for time-

varying fields. Likewise, we shall see that (1.28) will have to be modified to include

time-varying characteristics.]

1.3.2 A Magnetic Constitutive Relationship and MagneticFlux Density

For electric fields, we have referred to equation (1.5) as a constitutive relation. An-

other of the so-called constitutive relationships states that the magnetic flux density,

��, measured in webers per square metre (Wb/m2) or tesla (T) is related to the

magnetic field intensity, ��, in A/m via the expression

�� = 𝜇��

where 𝜇, in henrys per metre (H/m), is referred to as the magnetic permeability of

the medium in which �� exists. The weber may in fact be seen (when we get to time-

19

Page 20: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

varying fields) to be equivalent to the volt-second. The permeability is a “description”

of the magnetic nature of the medium (we will consider this in more detail in

Section 1.3.6). For free space, or a non-magnetic material,

�� = 𝜇0�� (1.29)

where 𝜇0 = 4𝜋 × 10−7 H/m. Note that “magnetic flux”, Φ, in webers is given by

analogy to equation (1.4) as

Φ =∮𝑆�� ⋅ 𝑑�� = 0 (Closed surface here.)

since no isolated magnetic “charge” (pole) has been found. On replacing �� with ��

in the divergence theorem given in equation (1.6), we get

∇ ⋅ �� = 0 (1.30)

If the surface is not closed,

We may now summarize the set of equations which describes all non-time-varying

electromagnetics:

Point Form:

∇ × �� = 0

∇ × �� = 𝐽

∇ ⋅ �� = 𝜌𝑣

∇ ⋅ �� = 0

Integral Form: ∮𝐶�� ⋅ 𝑑�� = 0∮

𝐶�� ⋅ 𝑑�� = 𝐼𝑒 =

∫𝑆𝐽 ⋅ 𝑑��∮

𝑆�� ⋅ 𝑑�� = 𝑄 =

∫vol

(∇ ⋅ ��

)𝑑𝑉 =

∫vol

𝜌𝑣𝑑𝑉∮𝑆�� ⋅ 𝑑�� = 0

20

Page 21: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

We note too that a vector field is uniquely specified by its divergence and curl.

Also, given any vector field ��

(1) ∇ ⋅ �� = 0 implies that the field is solenoidal – i.e., no source or sink in regions

where this is true;

(2) ∇ × �� = 0 implies that the field is irrotational.

1.3.3 Relevant Scalar and Vector Potentials:

Scalar Potential

We have already seen the scalar potential field, 𝑉 , associated with the electric

field and noted the simplicity with which �� may be found when 𝑉 is known – i.e.,

�� = −∇𝑉 (see equation (1.11)). In fact, using what we already know about a

conservative �� field (i.e., ∇ × �� = 0) and Gauss’ law in point form (∇ ⋅ �� = 𝜌𝑣)

it is easy to derive Poisson’s equation (see equation (1.15)). Equation (1.15) has a

solution given by equation (18), p. 92 of the text. As we have said, if 𝑉 can be

determined, �� follows immediately from (1.11).

In summary, then, starting with

(Conservative electrostatic field)

(Gauss’ law)

(Constitutive relationship)

and choosing �� = −∇𝑉 , we may arrive at Poisson’s equation, ∇2𝑉 = −𝜌𝑣𝜖0, from

which, as noted above,

𝑉 (��) =∫𝑣′

𝜌𝑣(��′)𝑑𝑣′

4𝜋𝜖0 ∣�� − ��′∣ (1.31)

21

Page 22: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Vector Potential

A similar idea can be constructed for finding the magnetic field, ��, from a so-

called vector potential which we shall label ��. Consider this time

∇ ⋅ �� = 0 (1.30) (Gauss’ Law – no magnetic poles)

∇ × �� = 𝐽 (1.28) (Ampere’s Law)

�� = 𝜇0�� (1.29) (Constitutive Relationship)

It may be recalled (see Appendix A.3 of text) that for any vector �� ,

(1.32)

Thus, we choose a “vector potential” �� defined by

�� = ∇ × �� (1.33)

which by (1.32) is a solution of (1.30). We know from (1.29) and (1.20) (the Biot-

Savart law) that �� depends on the (assumed known) source current density, 𝐽 . Thus,

if we are able to relate �� to 𝐽 , (1.33) will yield ��. [By the way, is �� unique?]

Using (1.29) in (1.33) and substituting into (1.28), we get

(1.34)

a second-order partial differential equation relating our chosen vector potential to

source current density, 𝐽 . We wish to obtain a solution for �� from (1.34) so that it

may be used in (1.33) to find �� or ��.

Again using Appendix A.3 of the text, we observe that

∇ × (∇ × ��) = ∇(∇ ⋅ ��)− ∇2��

where (in cartesian coordinates)

∇2�� =∂2��

∂𝑥2+

∂2��

∂𝑦2+

∂2��

∂𝑧2.

Using this identity in (1.34) yields

∇(∇ ⋅ ��)− ∇2�� = 𝜇0𝐽 . (1.35)

22

Page 23: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Let us now assume that, for the case of magnetostatics, �� is solenoidal, i.e.

∇ ⋅ �� = 0 (1.36)

so that �� is uniquely specified by (1.33) and (1.36). [We shall address this assumption

in the near future.] Based on this requirement, (1.35) reduces to

∇2�� = −𝜇0𝐽 (1.37)

Comparing (1.37) with Poisson’s equation gives us some insight as to why �� is called

a vector potential. The solution of (1.37) may be thus patterned after (1.31) and we

write it down directly as

��(��) =∫𝑣′

𝜇0𝐽(��′)𝑑𝑣′

4𝜋 ∣�� − ��′∣ . (1.38)

Similarly, in terms of surface current density, ��,

(1.39)

and for line currents, 𝐼,

(1.40)

Note that while �� is perpendicular to the plane defined by 𝐼𝑑��′ and the observation

point, �� is in the same direction as 𝐼. We note that (1.38)–(1.40), unlike the Biot-

Savart law, have no cross product. In general, this makes these integrals easier to

evaluate. As is seen by the following example, however, problems may arise when

dealing with idealized sources, for example, line currents which are infinitely long,

etc..

23

Page 24: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

The previous (approximately) 2 pages should be handwritten notes on the example

using a vector potential.

Ampere’s Law implies Biot-Savart

We have Ampere’s law in point form as

∇ × �� = 𝐽 (I)

and using the vector potential of (1.34) along with equations (1.29) and (1.33)

�� = ∇ ×∫𝑣′

𝐽(��′)𝑑𝑣′

4𝜋 ∣�� − ��′∣ . (II)

Note: ��(��) ≡ ��(𝑥, 𝑦, 𝑧) so, in cartesian coordinates, the curl in (II) is with respect

to the unprimed coordinates (𝑥, 𝑦, 𝑧).

From Appendix 3.A we have for any vector field �� and scalar 𝐶

∇ × (𝐶��) = ∇𝐶 × ��+ 𝐶∇ × ��. (III)

Identifying 𝐶 ≡ ∣�� − ��′∣ and �� ≡ 𝐽 (��′), and interchanging the order of integration and

differentiation (which is allowable for well-behaved integrands) in (II), (III) becomes

(IV)

because 𝐽(��′) is NOT a function of (𝑥, 𝑦, 𝑧) but of (𝑥′, 𝑦′, 𝑧′) (which means ∇𝑥,𝑦,𝑧 ×𝐽(��′) = 0). It is easily verified (DO THIS) that

where

1

∣�� − ��′∣ =

so that (IV) becomes

∇ ×{

1

∣�� − ��′∣𝐽(��′)

}=

=

26

Page 25: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Substituting this last expression into (II) gives

�� =

which is exactly the Biot-Savart law of equation (1.20). Working this procedure in

reverse would produce Ampere’s law from the Biot-Savart law.

Justification for Setting ∇ ⋅ �� = 0 – for Non-time-varying Currents

Preliminaries:

Recall from an earlier course the definition of the convolution of functions, 𝑓(𝑡)

and 𝑔(𝑡), for example:

𝑓(𝑡)𝑡∗ 𝑔(𝑡) =

∫𝑓(𝑡′)𝑔(𝑡− 𝑡′)𝑑𝑡′

=∫

𝑓(𝑡− 𝑡′)𝑔(𝑡′)𝑑𝑡′

= ℎ(𝑡), say.

Taking derivatives we find

𝑑

𝑑𝑡

{𝑓(𝑡)

𝑡∗ 𝑔(𝑡)}

=𝑑

𝑑𝑡

∫𝑓(𝑡′)𝑔(𝑡− 𝑡′)𝑑𝑡′

=∫

𝑓(𝑡′)𝑑

𝑑𝑡𝑔(𝑡− 𝑡′)𝑑𝑡′

= 𝑓(𝑡)𝑡∗ 𝑑𝑔(𝑡)

𝑑𝑡

OR

𝑑

𝑑𝑡

{𝑓(𝑡)

𝑡∗ 𝑔(𝑡)}

=𝑑

𝑑𝑡

∫𝑓(𝑡− 𝑡′)𝑔(𝑡′)𝑑𝑡′

=∫

𝑑

𝑑𝑡𝑓(𝑡− 𝑡′)𝑔(𝑡′)𝑑𝑡′

=𝑑𝑓(𝑡)

𝑑𝑡

𝑡∗ 𝑔(𝑡)

So, for convolutions,

𝑑

𝑑𝑡(𝑓 ∗ 𝑔) =

𝑑𝑓

𝑑𝑡∗ 𝑔 = 𝑓 ∗ 𝑑𝑔

𝑑𝑡=

𝑑ℎ

𝑑𝑡.

Finally we note that

∇ ⋅ �� =∂𝐴𝑥

∂𝑥+

∂𝐴𝑦

∂𝑦+

∂𝐴𝑧

∂𝑧.

27

Page 26: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Using these preliminary considerations we return to the main problem of showing

that ∇ ⋅ �� = 0 for non-time-varying currents.

With a view to using the above, we write (1.38)

explicitly as

where3𝑑∗ represents a 3-dimenensional spatial convolution. Now,

𝐴𝑥(𝑥, 𝑦, 𝑧) =

𝐴𝑦(𝑥, 𝑦, 𝑧) =

𝐴𝑧(𝑥, 𝑦, 𝑧) =

On the basis of the preliminaries, we may write the derivatives of the last expressions

as

so that

28

Page 27: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

But for non-time-varying currents, the equation of continuity gives

Specification of the vector potential in this way is called setting the gauge and ∇ ⋅ �� =

0, in particular, is the the so-called Coulomb gauge. Remember, for a complete

(unique) specification, both ∇ × �� and ∇ ⋅ �� must be established.

1.3.4 The Lorentz Force Equation and Related Concepts

Recall that a charge, 𝑄, in an electric field ��, experiences a force, ��𝑒, given as

��𝑒 = 𝑄�� (1.41)

Note that 𝐹𝑒 is in the same direction as �� and can therefore cause a linear acceleration

of the charge – i.e. the force can change the charge’s velocity, thus imparting or

removing kinetic energy. Another way of saying this is that the force is capable of

doing work on the charge. Recall that the acceleration vector, ��, by Newton’s second

law of motion (�� = 𝑚��, where 𝑚 is mass) is in the same direction as the force.

Next, a charge moving with a velocity �� in a magnetic field of flux density ��

experiences a force given by

��𝑚 = 𝑄(�� × ��). (1.42)

We note from the cross product that this force is perpendicular to both the field

vector and the velocity vector. Of course, a force of this nature may cause a change

in direction of �� but not in its magnitude (the analogy in mechanics is the centripetal

force on a mass moving with constant speed in a circular path). Thus, since �� causes

no change in ∣��∣, it’s neither capable of increasing nor decreasing the energy of the

charge involved.

Equations (1.41) and (1.42) hold for both steady and time-varying fields.

29

Page 28: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Combining the electric and magnetic field effects, we get from equations (1.41)

and (1.42)

𝐹 = 𝑄(�� + �� × ��). (1.43)

This is the famous LORENTZ FORCE EQUATION. It applies to any charge motion

in electric and magnetic fields (particle accelerators, magnetrons, CRT’s etc.).

Example: Determine the velocity relations and trajectory (path of motion) for a pos-

itive charge 𝑞 of mass 𝑚 moving in a magnetic field specified by �� = 𝐵0𝑧 if at time

𝑡 = 0, the charge is located at the origin and has a velocity given by �� = 𝑣0𝑦.

30

Page 29: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

1.3.5 Force and Torque

Force on Current Elements (Line, Surface and Volume):

Charges and Line Charges:

If we agree to drop the 𝑚 subscript from equation (1.42), but remember from the

context that 𝑄(�� × ��) is force due to the charge moving in a magnetic field, then a

differential force 𝑑�� on a differential charge 𝑑𝑄 is given by

𝑑�� = 𝑑𝑄�� × �� (1.44)

Recalling that for a constant line current 𝐼, 𝑑𝑄 = 𝐼𝑑𝑡 and that the velocity is �� =

𝑑��/𝑑𝑡, it is easy to see that equation (1.44) may be written for a differential line

current as

𝑑�� = 𝐼𝑑��× �� (1.45)

Illustration:

Integrating over a closed path (circuit) and assuming the current to be constant,

(1.45) leads to

�� =∮𝐶𝐼𝑑��× �� = −𝐼

∮𝐶�� × 𝑑�� (1.46)

SPECIAL CASE 1: Furthermore, if the field is uniform so that �� may also be

removed from the integral in (1.46) it becomes obvious that the total force on the

circuit is zero:

This does NOT mean that the torque is zero! (Think about force couples encountered

in mechanics).

SPECIAL CASE 2: Suppose next that we consider applying (1.46) to a portion of

33

Page 30: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

a straight conductor in a constant �� field – the path is no longer closed. We have

𝐹 = 𝐼∫𝐶𝑑��× �� = −𝐼�� ×

∫𝐶𝑑�� = −𝐼�� × ��

or

�� = 𝐼��× ��.

Thus, the magnitude of the force is given by the familiar expression

∣∣∣𝐹 ∣∣∣ = 𝐵𝐼𝐿 sin 𝜃 (1.47)

where 𝜃 is the angle between �� and the field vector. Of course, if equation (1.47) is

to be applied to an extended piece of current-carrying conductor, the conductor must

be straight and the field must be uniform.

Forces Experienced by Volume and Surface Current Densities:

In the prerequisite to this course, the (volume) current density was defined as

𝐽 = 𝜌𝑣 ��

and noting that 𝑑𝑄 = 𝜌𝑣𝑑𝑣 we have from (1.44) that 𝑑�� = 𝜌𝑣𝑑𝑣��× ��, which implies

𝑑�� = 𝐽 × ��𝑑𝑣 (1.48)

Thus, the total force on a finite volume is given by

�� =∫vol

𝐽 × ��𝑑𝑣 (1.49)

It may be similarly argued for surface current density, �� that

𝑑�� = �� × ��𝑑𝑆 (1.50)

Thus, the total force for a finite surface is given by

�� =∫𝑆�� × ��𝑑𝑆 (1.51)

Read the paragraphs on the Hall voltage in Section 9.2 of the text and know how

to explain the effect in terms of equation (1.44).

34

Page 31: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Torque on a Closed Circuit

While the force on a closed (dc) circuit in a uniform �� field is zero, the moment

of force or torque is not generally zero.

Definition: The torque, 𝑇 , or moment of force is defined with respect to the accom-

panying diagram as follows:

𝑇 = �� × 𝐹 (1.52)

(with reference to the origin in this illustration).

Note that 𝑇 is perpendicular to the plane containing �� and 𝐹 and points in the

direction determined by the usual right-hand rule for cross products. The magnitude

of 𝑇 is then

∣𝑇 ∣ = ∣��∣∣�� ∣ sin 𝜃 (1.53)

(i.e. 𝑇 = 𝑅𝐹 sin 𝜃).

Next consider a force couple, ��1 and ��2 so that ��1 + ��2 = 0 or 𝐹1 = −��2. This

couple is applied to some “object” in space as shown:

Illustration:

(��2 − ��1) is the “object” between points 𝑃2 and 𝑃1 of the force applications.

The net torque about an axis through the origin (⊥ to page) is given by

35

Page 32: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

(1.54)

which indicates that 𝑇 is independent of the choice of origin for the position vectors,

��1 and ��2. The individual moment arms have disappeared and the vector (��1 − ��2)

between the points of application has emerged.

Magnetic Torque on a Small Closed Loop (of Differential Area):

Consider, without loss of generality, a differential current loop in the 𝑥− 𝑦 plane

centered at the origin 𝑂 as shown.

Illustration:

If the dimension of the circuit is small and �� is continuous (but not necessarily

uniform), then over the small loop �� may be considered as some constant equal to

the value of the field at the origin (i.e., �� = ��0) – this is the zero-order effect.

For side 1,

Therefore, for side 1, the differential torque, 𝑑𝑇1, about an axis through 𝑂 is given

by

36

Page 33: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

Pages 27 and 28 should be done on looseleaf and end with equation (1.56).

SPECIAL CASE: Constant �� and 𝐼

If the applied �� field is uniform and the current is uniform in any planar loop,

we may integrate equation (1.56) to obtain

Using the magnetic dipole moment,

�� = 𝐼𝑆�� ,

with 𝑆 being the total surface area of any planar loop enclosed by a current, 𝐼, the

torque becomes

𝑇 = ��× �� . (1.57)

Illustration:

We see that 𝑇 tends to rotate the loop such that �� (= ��𝐼𝑆) and the induced field

��𝐼 tend toward the �� direction. That is, the induced field tends to line up with the

37

Page 34: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

impressed or applied field so as to maximize the total ��-field. We note that in the

magnitude sense, (1.58)

Therefore,

if �� ⊥ �� , 𝑇 → 𝑇max

if �� ∣∣ �� , 𝑇 → 0

The foregoing serves as a basis for modeling the nature of magnetic materials (and

classifying them as diamagnetic, paramagnetic, ferromagnetic, antiferromagnetic, fer-

rimagnetic or superparamagnetic). Please read the text, Section 9.5, pp 273-276 for

a description of the classes of magnetic materials.

1.3.6 Magnetic Materials and Boundary Conditions

Magnetization and Permeability

The motion of the bound charges (i.e. NOT free charge) inside a material due

to electrons orbiting and spinning and nucleii spinning may have associated with it

some sort of net bound current, 𝐼𝑏. We had earlier the differential magnetic dipole

moment, 𝑑�� = 𝐼𝑑�� where 𝑑�� is the vector differential area enclosed by the current,

𝐼. If we choose to rename the magnetic dipole moment as �� (rather than 𝑑��) for

the differential area,

�� = 𝐼𝑏𝑑�� (1.59)

due to the bound current. Consider a sample of material containing 𝑛 magnetic dipole

moments per unit volume (so that in a volume Δ𝑣, there are (𝑛Δ𝑣) dipoles).

The total dipole moment (a vector sum) may thus be written as

(1.60)

This could be zero, and often is if the 𝑚𝑖’s are randomly distributed. However, it

is possible that the total dipole moment is non-zero and this is especially true if the

material is subjected to a ��-field or has a history of being so subjected.

38

Page 35: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

By definition, the magnetization, �� , is defined as

�� = limΔ𝑣→0

��totalΔ𝑣

= limΔ𝑣→0

1

Δ𝑣

𝑛Δ𝑣∑𝑖=1

��𝑖 (1.61)

Thus, �� has units of A/m which is the same as the units of ��. Consequently, ��

may be regarded as a kind of �� field arising from the alignment of the ��𝑖’s within the

material when an external �� (or ��) field is applied. Thus, if this alignment occurs,

the total �� field appears larger, in general, than it would be if �� were not present.

[Remember, �� is due to the “free” current; �� is due to the “bound” current].

Now,and

�� = 𝜇0(�� + ��) (1.62)

where �� is the total magnetic flux density, �� arises from “external” or free current

and �� accounts for all of the atomic magnetic effects. This may be further emphasize

by writing

�� =��

𝜇0− �� . (1.63)

Special Case: Linear Isotropic Material

In general, magnetic materials are not linear – for example, a change in the ex-

ternal field impinging a magnetic material will not, in general, cause a proportional

change in the magnetic field within the material. However, for small flux densities

(low currents) it is possible that a material may be considered to be linear: In par-

ticular, we assume that the magnetization �� depends in a linear way on the applied

�� field and may thus be modeled as

�� = 𝜒𝑚�� (1.64)

where 𝜒𝑚 is dimensionless and is referred to as the magnetic susceptibility of the

material. Using (1.64), equation (1.62) may be cast as

�� = 𝜇0(�� + 𝜒𝑚��)

�� = 𝜇0(1 + 𝜒𝑚)��

39

Page 36: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

or

�� = 𝜇�� (1.65)

where 𝜇 (= 𝜇0(1 + 𝜒𝑚))is the permeability of the medium under consideration. We

define a dimesionless constant called the relative permeability, 𝜇𝑅, of a medium as

𝜇𝑅 = (1 + 𝜒𝑚) =𝜇

𝜇0

(1.66)

where, of course, 𝜇0 is the permeability of free space.

[NOTE: In anisotropic magnetic media, 𝜇 is, in geneal, a 3× 3 matrix.]

Example: Problem 19, page 301 of text.

40

Page 37: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

The Boundary Conditions for Magnetic Fields

Just as was done for the static electric field, we now wish to consider the situation

in which magnetic fields exist at a boundary between two media of differing mag-

netic properties. Boundaries represent mathematical discontinuities so that, at such

interfaces, the derivatives encountered in Maxwell’s equations do not strictly exist.

Thus, in addition to the equations which we already have, we should develop a set of

“boundary conditions” so that the former equations may be supplemented to handle

e-m fields at the media interfaces.

Consider the figure below, where the interface is basically defined by (homoge-

neous) regions of differing permeabilities.

Referring to the CLOSED cylindrical surface on the left and using Gauss’ law for the

magnetic field we write∮𝑆 �� ⋅𝑑�� = 0 where for the top base of the region, Δ�� = Δ𝑆��1

and for the bottom base, Δ�� = Δ𝑆��2. Thus, Gauss’ law may be written as

(1.67)

In the limit as the sides of the cylindrical surface tend toward 0 (i.e. as the boundary

is approached from “top” and “bottom” – see figure on the right above) it is clear

that the unit normals to the surface may be written as

��1 → ��

��2 → −��

41

Page 38: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

and (side contribution) → 0. Equation (1.67) becomes

��1 ⋅ ��− ��2 ⋅ �� = 0

or using the subscript 𝑁 to indicate “normal” components we have

𝐵𝑁1 = 𝐵𝑁2 (1.68)

This implies that the normal component of the �� field is continuous across the bound-

ary. Given that magnetic field constitutive relationship, it immediately follows that

𝐻𝑁2 =𝜇1

𝜇2

𝐻𝑁1 (1.69)

and it is seen that the normal component of the magnetic field intensity is discontin-

uous across the boundary.

Next, to investigate the tangential components of the �� fields at the boundary,

consider the figure below:

Let’s apply Ampere’s law to the closed path. If we allow for the possibility that

a uniform surface current density, ��, may exist on the boundary and that the com-

ponent of this density normal to the closed path is 𝐾, then

In the limit, as the ends tend toward 0,

42

Page 39: Unit 1 Relevant Electrostatics and Magnetostatics (Old and ...

or

𝐻𝑡1 −𝐻𝑡2 = 𝐾 (1.70)

Of course, if no surface current density exists on the boundary, we have the following

very important simplification:

𝐻𝑡1 −𝐻𝑡2 = 0 (1.71)

In terms of the unit normal, ��, to the surface, it is easy to see that equations (1.70)

and (1.71) may be written as

and

respectively.

Equation (1.71) implies that, if there are no surface currents, the tangential ��

fields are continuous at the boundary. This has very important consequences! – stay

tuned.

43