HEAD PREVIOUS

Chapter 1
Maxwell's Equations and Electromagnetic Fields

1.1  Introduction

1.1.1  Maxwell's Equations (1865)

The governing equations of electromagnetism

∇. E= ρ

ϵ0
      
∇. B = 0
(Coulomb/Gausss Law)
      
(1.1)
∇∧E = −B

∂t
      
∇∧B = μ0 j + 1

c2
E

∂t
(Faradays Law)
      
(Amperes Law)

E        electric field, describes the force felt by a (stationary) charge q: F = q E
B        magnetic field, describes the force felt by a current
       i.e. a moving charge (velocity v): F = q vB
Thus the Lorentz Force (on charge q) is

F = q ( E+ vB)
(1.2)

figure/rho_j.gif
Figure 1.1: Charge density is local charge per unit volume. Current density is current per unit area.
ρ        electric charge density (Coulombs/m3). Total charge Q = ∫V ρd3x
j        electric current density (Coulombs/s/m2)
       Current crossing area element dA is j . dA Coulomb/s = Amps.
[Note. Sometimes the first Maxwell equation is called Gauss's rather than Coulomb's Law.]

1.1.2  Historical Note

Much scientific controversy in 2nd half of 19th century concerned question of whether E, B were `real' physical quantities of science or else mere mathematical conveniences for expressing the forces that charges exert on one another. English science (Faraday, Maxwell) emphasized the fields; German mostly the act-at-a-distance. Since  ∼  1900 this question has been regarded as settled in favor of the fields. And modern physics, if anything, tends to regard the field as more fundamental than the particle.

1.1.3  Auxiliary Fields and Electromagnetic Media

Electromagnetic texts often discuss two additional "auxiliary" fields D the "electric displacement" and H the "magnetic intensity" which account for dielectric and magnetic properties of materials. These fields are not fundamental and introduce unnecessary complication and possible confusion for most of our topics. Therefore we will avoid them as much as possible. For the vacuum, ϵ0 E = D and B = μ0 H.

1.1.4  Units

Historically there were two (or more!) different systems of units, one defining the quantity of charge in terms of the force between two stationary charges (the "Electrostatic" units) and one defining it in terms of forces between (chargeless) currents (the "Electromagnetic" system). Electrostatic units are based on Coulomb's law ∇. E = ρ/ϵ0 and electromagnetic units on the (steady-state version of) Ampere's law ∇∧B0 j. The quantities 1/ϵ0 and μ0 are therefore fundamentally calibration factors that determine the size of the unit charge. Choosing one or other of them to be 4π amounts to choosing electrostatic or electromagnetic units. However, with the unification of electromagnetism, and the subsequent realization that the speed of light is a fundamental constant, it became clear that the units of electromagnetism ought to be defined in terms of only one of these laws and the speed of light. Therefore the "System Internationale" SI (or sometimes MKSA) units adopts the electromagnetic definition because it can be measured most easily, but with a different μ0, as follows. "One Ampere is that current which, when flowing in two infinitesimal parallel wires 1m apart produces a force of 2 ×10−7 Newtons per meter of their length." An Amp is one Coulomb per second. So this defines the unit of charge. We will show later that this definition amounts to defining
μ0 = 4π×10−7        (Henry/meter)
(1.3)
and that because the ratio of electromagnetic to electrostatic units is c2
ϵ0 = 1

c2 μ0
= 8.85 ×10−12       (Farad/meter)
(1.4)
μ0 is called the "permeability of free space". ϵ0 is called the "permittivity of free space". [See J.D. Jackson 3rd Ed, Appendix for a detailed discussion.]

1.2  Vector Calculus and Notation

Electromagnetic quantities include vector fields E,B etc. and so EM draws heavily on vector calculus. ∇ is shorthand for a vector operator (gradient)
∇ϕ =
∂ϕ

∂x
, ∂ϕ

∂y
, ∂ϕ

∂z

= ∂ϕ

∂xi
    (suffix notation)
(1.5)
giving a vector gradient from a scalar field ϕ. ∇ can also operate on vector fields by scalar (.) or vector (∧) multiplication.

1.2.1  Divergence


∇. E= ∂Ex

∂x
+ ∂Ey

∂y
+ ∂Ez

∂z
= ∂Ei

∂xi
(1.6)

1.2.2  Curl


∇∧E=
∂Ez

∂y
∂Ey

∂z
, ∂Ex

∂z
∂Ez

∂x
, ∂Ey

∂x
∂Ex

∂y

= ϵijk ∂Ek

∂xj
(1.7)

1.2.3  Volume Integration





V 
    d3x
(1.8)
d3x is shorthand for dxdydz = dV, the volume element.
figure/elements.gif
Figure 1.2: Elements for surface and line integrals.

1.2.4  Surface Integration





S 
v. dA    .
(1.9)
The surface element dA or often dS is a vector normal to the element.

1.2.5  Line (Contour) Integration





C 
v. dl    .
(1.10)
Line element dl.

1.2.6  The Meaning of divergence: ∇.

figure/volume_element.gif
Figure 1.3: Cartesian volume element.
Consider a volume element. Evaluate the total flux of a vector field v out across the element's surface. It is the sum v.dA over the six faces of the cuboid
[vx (x + dx/2) − vx (x−dx/2)] dydz+
[vy (y + dy/2) − vy (y−dy/2) ] dzdx+
[vz (z + dz/2) − vz (z−dz/2)] dxdy
=
(1.11)
dxdydz
dvx

dx
+ dvy

dy
+ dvz

dz

=
d3x ∇. v= dV ∇. v
So for this elemental volume:



dS 
v. dA=


dV 
∇. v d3x
(1.12)
figure/add_vol_elem.gif
Figure 1.4: Adjacent faces cancel out in the sum of divergence from many elements.
But any arbitrary finite volume can be considered to be the sum of many small cuboidal elements. Adjacent internal face contribution cancel out hence only the external surface contributions remain, so



S 
v. dA=


V 
∇. v d3x
(1.13)
for any volume V with surface S, and arbitrary vector field v. This is Gauss's Theorem.

1.2.7  The Meaning of Curl: ∇∧

figure/curl.gif
Figure 1.5: Rectangular surface element with axes chosen such that the normal is in the z-direction.
Consider an arbitrary rectangular surface element Choose axes such that normal is in z-direction and edges along x and y. Arbitrary vector field v(x). Evaluate the contour integral of v, clockwise around dC (relative to the direction z), round boundary of element, regarding the center as x.

(⎜)



dC 
v. dl
=
v(x,y−dy/2). dx + v(x+dx/2,y) . dy
(1.14)
+
v(x,y+dy/2) . (−dx)+ v(x−dx/2,y) . (−dy)
=
−dvxdx + dvydy =
∂vy

∂x
∂vx

∂y

dxdy
=
(∇∧v)z dA = (∇∧v) . dA    .
(1.15)
So integral v.d l around element is equal to the curl scalar-product area element.
figure/stokes.gif
Figure 1.6: Arbitrary surface may be divided into the sum of many rectangular elements. Adjacent edge integral contributions cancel.
Apply to arbitrary surface; divide surface up into many elements dA. All internal edge integrals cancel. Hence

(⎜)



C 
v.dl=


S 
( ∇∧v) . dA
(1.16)
This is Stokes' Theorem.

1.3  Electrostatics and Gauss' Theorem

Gauss's theorem is the key to understanding electrostatics in terms of Coulomb's Law ∇.E = ρ/ ϵ0.

1.3.1  Point Charge q

Apply Gauss's Theorem to a sphere surrounding q
figure/charge_sphere.gif
Figure 1.7: Spherical volume, V over which we perform an integral of Coulomb's law to deduce E.




S 
E. dA=


V 
∇. E d3x =


V 
ρ

ϵ0
 d3x = q

ϵ0
    .
(1.17)
But by spherical symmetry E must be in radial direction and Er has magnitude constant over the sphere. Hence ∫S E. dA = ∫S ErdA = ErS dA = Er 4πr2. Thus
Er4πr2 = q

ϵ0
       or
(1.18)

Er = q

4πϵ0 r2
       i.e.       E = q

4πϵ0
r

r3
    .
(1.19)
Consequently, force on a second charge at distance r is
F = q1q2

4πϵ0r2
(1.20)
Coulomb's inverse-square-law of electrostatic force.

1.3.2  Spherically Symmetric Charge ρ(r)

Notice that point-charge derivation depended only on symmetry. So for a distributed charge-density that is symmetric argument works just the same i.e.
E=
^
r
 
Er        Er = q

4πϵ0r2
(1.21)

where now     q =


V 
ρ d3x =
r

0 
ρ(r) 4πr2dr .
(1.22)
Electric field due to a spherically symmetric charge density is equal to that of a point charge of magnitude equal to the total charge within the radius, placed at the spherical center.

1.3.3  Arbitrary Charge Distribution

If there is no specific symmetry Gauss's Theorem still applies:
figure/gauss.gif
Figure 1.8: Arbitrary volume for Gauss's Theorem.




S 
E. dA=


V 
∇. E d3x =


V 
ρ

ϵ0
 d3x = q

ϵ0
(1.23)
q is the total charge (integral of charge density) over the volume. ∫S E. dA is the total flux of electric field across the surface S.

1.3.4  Intuitive Picture

figure/eline_start.gif
Figure 1.9: Intuitive picture of charges and field-lines.
Each (+ve) charge is the origination point of an electric-field-line. [Each −ve charge is termination of a field line]. The total charge in volume V determines the number of field-lines that start in V. Field lines only start/end on charges (Coulomb's Law) so all must escape from the volume, crossing surface S (somewhere). [Field lines that start and end in V contribute neither to ∫E. dA nor to q, because of cancellation]. ∫S E. dA can be thought of as counting the "number of field lines" crossing the surface. [Of course it is an arbitrary choice how big we consider the charge is that gives rise to one field-line.] Intuitive view of electric field "intensity": Strength of E is proportional to the number of field-lines per unit area. (Fig 1.10).
figure/intensity.gif
Figure 1.10: Spacing of field-lines is inversely proportional to field-strength.
All these intuitive views are conceptually helpful but are not formally necessary. Electromagnetism is considered completely described by Maxwell's equations without need for these pictures.

1.3.5  Electric Potential (for static problems [(∂)/(∂t)] → 0)

In the static situation there is no induction and Faraday's law becomes ∇∧E = 0. By the way, this equation could also be derived from the inverse-square-law by noting that
figure/irrotation.gif
Figure 1.11: Each element contributes an irrotational component to E. Therefore the total E is irrotational.

∇∧ r

r3
= 0
(1.24)
so by the linearity of the ∇∧ operator the sum (integral) of all Electric field contributions from any charge distribution is curl-free "irrotational":
∇∧
ρ(r′)

4πϵ0
rr

|rr′|3
 d3r′ = 0
(1.25)
[This shows that the spherical symmetry argument only works in the absence of induction · B would define a preferred direction; asymmetric!] For any vector field E, ∇∧E = 0 is a necessary and sufficient condition that E can be written as the gradient of a scalar E = − ∇ϕ.

Necessary


( ∇∧∇ϕ)z =

∂x
∂ϕ

∂y

∂y
∂ϕ

∂x
= 0       (et sim x, y)     .
(1.26)
Curl of a gradient is zero.

Sufficient (prove by construction)

figure/potential_curl.gif
Figure 1.12: Two different paths from 0 to x construct a closed contour when one is reversed.
Apply Stokes' theorem to a closed contour consisting of any 2 paths between points 0 and x.

(⎜)



C 
E. dl=


x

0 
E. dl

Path 2
 



x

0 
E. dl

Path 1
 
=


S 
∇∧E. dS

= 0
by hypothesis. 
(1.27)
So ∇∧E = 0 ⇒ ∫oxE. dl is independent of chosen path, i.e. it defines a unique3 quantity. Call it −ϕ(x). Consider ∇ϕ defined as the limit of δϕ between adjacent points.
− ∇ϕ = ∇

x

 
E. dl
= E       .       
(1.28)
Many electrostatic problems are most easily solved in terms of the electric potential ϕ because it is a scalar (so easier). Governing equation:
∇. E
=
− ∇. ∇ϕ = − ∇2 ϕ = ρ

ϵ0
(1.29)
2
=
2

∂x2
+ 2

∂y2
+ 2

∂x2
       is the "Laplacian" operator
(1.30)
2 ϕ = −ρ

ϵ0
       "Poissons Equation".
(1.31)

1.3.6  Potential of a Point Charge [General Potential Solution]

One can show by direct differentiation that
1

r
= − r

r3
       .
(1.32)
So by our previous expression E = [q/(4 πϵ0)] [(r)/(r3)] we can identify
ϕ = q

4 πϵ0
1

r
(1.33)
as the potential of a charge q (at the origin x =0).

1.3.7  Green Function for the Laplacian

For a linear differential operator, L, mathematicians define something called "Green's function" symbolically by the equation
LG(x, x′) = δ(xx′)       .
(1.34)
If we can solve this equation in general, then solutions to
Lϕ = ρ(x)
(1.35)
can be constructed for arbitrary ρ as
ϕ(x) =
G(x, x′) ρ(x′)  d3x′
(1.36)
because of the (defining) property of the δ-function

f(x′) δ(xx′)  d3x′ = f(x)        .
(1.37)
When L is the Laplacian, ∇2, the Green function is

G(x,x′) = −1

4π|xx′|
       .
(1.38)

This fact may be derived directly from the solution for the potential for a point charge. Indeed, a point charge is exactly the delta-function situation whose solution is the Green function. In other words, the charge density for a point charge of magnitude q at position x′ is
ρ(x) = q δ(xx′)        ;
(1.39)
so the point-charge potential, namely,
ϕ = q

4 πϵ0
1

|xx′|
      ,
(1.40)
is the solution of the equation:
2 ϕ = −   q

ϵ0
δ(xx′)       .
(1.41)
Consequently, solution of Poisson's equation can be written as the integral of the Green function:

ϕ(x) =

−ρ( x′)

ϵ0


−1

4π|xx′|

 d3 x′ =
ρ(x′)

4 πϵ0
 d3x′

|xx′|
    .
(1.42)

Informally, the smooth charge distribution ρ can be approximated as the sum (→ ∫) of many point charges ρ(x′) d3x′, and the potential is the sum of their contributions.

1.3.8  Boundary Conditions

Strictly speaking, our solution of Poisson's equation is not unique. We can always add to ϕ a solution of the homogeneous (Laplace) equation ∇2ϕ = 0. The solution only becomes unique when boundary conditions are specified. The solution
ϕ(x) =
ρ( x′)

4 πϵ0
 d3x′

|xx′|
(1.43)
is correct when the boundary conditions are that
ϕ→ 0        as        | x| → ∞    :
(1.44)
no applied external field.
In practice most interesting electrostatic calculations involve specific boundaries. A big fraction of the work is solving Laplace's equation with appropriate boundary conditions. These are frequently the specification of ϕ on (conducting) surfaces. The charge density on the conductors is rarely specified initially.

1.3.9  Parallel Plate Capacitor

figure/pp_cap.gif
Figure 1.13: The parallel-plate capacitor.
Idealize as 1-dimensional by ignoring the edge effects. 1-d Laplace equation in the vacuum gap (where ρ = 0) is
d2ϕ

dz2
= 0
(1.45)
Solution ϕ = a − Ez        a,E const.
Hence electric field is E = E z.
Notice how this arises purely from the translational invariance of ϕ( [d/dx] = [d/dy] = 0 ).
Choose z=0 as one plate of capacitor. Other at z = d.
Choose ϕ(z=0) = 0: reference potential, making a = 0.
Potential of other plate: V = ϕ(d) = −Ed.
The question: how much charge per unit area is there on the plates when the field is E?
figure/pp_vol.gif
Figure 1.14: Elemental volume for calculating charge/field relationship.
Answer by considering a flat elemental volume, with area A surrounding the +ve plate. Apply Gauss's law



S 
E. dS
=



A 
O + (−E) ds = −E A
=



V 
ρ

ϵ0
 d3x = Q

ϵ0
= 1

ϵ0
σ A
(1.46)
Hence
σ = − ϵ0 E . = ϵ0 V

d
(1.47)
Therefore if the total area is A, the total charge Q, and the voltage V between plates are related by
Q =
ϵ0 A

d

V
(1.48)
And the coefficient [(ϵ0A)/d] is called capacitance, C. Notice our approach:

1.3.10  Charge on an arbitrary conductor

Consider a conductor, electrostatically charged.
figure/arb_vol.gif
Figure 1.15: Arbitrary-shaped conductor possesses only surface charges related to the local normal field.
Current is zero.
So E is zero, anywhere inside because of conductivity.
Choose any volume internally: E = 0 ⇒ ∇. E = 0 ⇒ρ = 0. There is no internal charge. It all resides on surface. At the surface there is an E just outside. E is perpendicular to surface ds because surface is an equipotential (& E = − ∇ϕ). Hence applying Gauss's law to a pill box



V 
∇. E d3x =


S 
E. dS
=
E ds
=
ρ

ϵ0
 d3 x =
σ

ϵ0
ds
=
σ

ϵ0
ds
(1.49)
where σ = surface charge density Hence σ = ϵ0 E. Of course, in this general case E (= Enormal) is not uniform on the surface but varies from place to place. Again procedure would be: solve ϕ externally from ∇2ϕ = 0; then deduce σ; rather than the other way around.

1.3.11  Visualizing Electric Potential and Field


E= − ∇ϕ
(1.50)
figure/visualize.gif
Figure 1.16: Contours of potential and corresponding field-lines (marked with arrows). Only the field-lines emanating from the larger elliptical conductor are drawn.
Consider a (2-D) contour plot of ϕ. The value of ϕ can be thought of as the potential energy of a charge of 1 Coulomb. Thus there is a perfect analogy to gravitational potential energy and height contours. The force at any point on the hill is downward (on a +ve charge), which is perpendicular to the contours of constant ϕ. The strength of the force (E) is proportional to the steepness of the hill: i.e. how close together the contours are (of ϕ). When plotting field-lines, i.e. lines following the electric field direction, we generally also consider the electric field intensity to be the number of field-lines per unit area. So also the closeness of field-lines indicates field strength. In charge-free regions ∇. E = 0 implies field-lines have no beginning or end. However if ρ ≠ 0 then electric field lines do possibly have ends (on the charges). The potential contours never have ends.

1.3.12  Complex Potential Representation 2-D

In charge-free region, ∇. E = 0 ⇒ ∇2 ϕ = 0. This causes there to be an intimate relationship between field-lines and ϕ-contours. In 2 dimensions this relationship allows complex analysis to be used to do powerful analysis of potential problems. Consider a complex function f(z) = ϕ(z) + i ψ(z) where z = x + iy is the complex argument with real and imaginary parts x & y; and f has real and imaginary parts ϕ & ψ. f is "analytic" if there exists a well defined complex derivative [df/dz] (which is also analytic), defined in the usual way as limz′→ z ( [(f( z′) − f ( z ))/(z′− z )] ) . In order for this limit to be the same no matter what direction (x,y) it is taken in, f must satisfy the "Cauchy-Riemann relations"
∂ϕ

∂x
= ∂ψ

∂y
       ;        ∂ϕ

∂y
= − ∂ψ

∂x
(1.51)
Which, by substitution imply ∇2ϕ = 0, ∇2ψ = 0, and also
∇ϕ. ∇ψ = 0
(1.52)
regarding x,y as 2-d coordinates. This shows that
  1. The real part of an analytic function solves ∇2ϕ = 0.
  2. The contours of the corresponding imaginary part, ψ, then coincide with the electric field-lines.
Finding complex representations of potential problems is one of the most powerful analytic solution techniques. However, for practical calculations, numerical solution techniques are now predominant.

1.4  Electric Current in Distributed Media

Ohms law, V = IR, relates voltage current and resistance for a circuit or discrete element. However we often care not just about the total current but about the current density in finite-sized conductors (e.g. electromagnets). This requires a local Ohm's law which is
E= ηj
(1.53)
where η is the medium's electric resistivity. Often the conductivity σ = 1/η is used. j = σE, (but I'll try to avoid confusion with surface charge density σ). Such a linear relationship applies in most metals.

1.4.1  Steady State Conduction

Conservation of charge can be written
∇. j = − ∂ρ

∂t
(1.54)
so, in steady state, ∇. j = 0, i.e.
∇.
1

η
E
= ( E. ∇) 1

η
+ 1

η
∇. E= 0
(1.55)
If conductivity is uniform (∇[1/(η)] = 0) or invariant along E, we therefore have ∇. E = 0 ⇒ ρ = 0. "Uniform conductivity conductors acquire zero volume charge density in steady state".

1.4.2  Conductor Boundary Conditions (Steady Currents)

figure/conductor.gif
Figure 1.17: A distributed conductor of finite conductivity, carrying current.
If currents are flowing so that E ≠ 0 in the conductor then conductors are not equipotential surfaces for solutions of Laplace's equation outside. Surface charges (only) are present on the conductor (uniform η). No current flows through the conductor surface (except at contacts) so
j . n = 0 ⇒ E. n = 0
(1.56)
inside conductor, while outside we have
E. n = ϵ0 σ
(1.57)
surface charge density. Normal components
figure/cond_bcs.gif
Figure 1.18: Boundary conditions across a conductor/vacuum (or insulator) interface.

En|inside = 0 [ En ]insideoutside = σ/ϵ0
(1.58)
Tangential components
[ Et ]outsideinside = 0
(1.59)
In particular, for solving ∇2 ϕ = 0 inside uniform η conductor, at conductor boundary:
∇ϕ. n = 0        (Neumann B.C.)
(1.60)
unlike the usual electrostatic B.C. ϕ = given. At electrical contacts ϕ given might be appropriate. A general approach to solving a distributed steady-current problem with uniform-η media:
  1. Solve Laplace's equation ∇2ϕ = 0 inside conductors using Dirichlet (ϕ-given) or possibly inhomogeneous Neumann (∇ϕ|n = given) BC's at contacts and ∇ϕ. n = 0 at insulating boundaries.
  2. Solve Laplace's equation ∇2 ϕ = 0 outside conductors using ϕ = given (Dirichlet) B.C. with the ϕ taken from the internal solution.

1.5  Magnetic Potential

Magnetic field has zero divergence ∇. B = 0. For any vector field4 B, ∇. B = 0 is a necessary and sufficient condition that B can be written as the curl of a vector potential B = ∇∧A.

1.5.1  ∇.B =0 Necessary


∇. ( ∇∧A)
=

∂x

∂Az

∂y
(z)

 
∂Ay

∂z
(y)

 

(1.61)
+

∂y

∂Ax

∂z
(x)

 
∂Az

∂x
(z)

 

(1.62)
+

∂z

∂Ay

∂x
(y)

 
∂Ax

∂y
(x)

 

       = 0
(1.63)
So only divergenceless fields can be represented.

1.5.2  ∇.B =0 Sufficient (outline proof by construction)

Consider the quantity
K(x) =
B( x′)

4 π|xx′|
 d3 x′       ,
(1.64)
a vector constructed from the integral of each Cartesian component of B. Applying our knowledge of the Green function solution of Poisson's equation, we know:
2 K = − B       .
(1.65)
Vector operator theorem (for any v):
∇∧( ∇∧v ) = ∇( ∇. v ) − ∇2 v        .
(1.66)
Hence
B= − ∇2 K = ∇∧( ∇∧K) − ∇( ∇. K )
(1.67)
We have proved Helmholtz's theorem that any vector field can be represented as the sum of grad + curl.] When ∇. B = 0 and |B| → 0 (fast enough) as |x| → ∞, one can show that ∇. K = 0 and so we have constructed the required vector potential
A= ∇∧K = ∇∧
B( x′)

4 π
 d3x′

|xx′|
       .
(1.68)
Notice that we have constructed A such that ∇. A = 0. However A is undetermined from B because we can add to it the gradient of an arbitrary scalar without changing B, since ∇∧∇χ = 0. So in effect we can make ∇. A equal any desired quantity ψ(x) by adding to A ∇χ such that ∇2χ = ψ. Choosing ∇. A is known as choosing a "Gauge" ∇. A = 0 is the "Coulomb Gauge".

1.5.3  General Vector Potential Solution (Magnetostatic)

Static Ampere's law ∇∧B = μ0 j. Now
μ0 j
=
∇∧( ∇∧A)
=
∇( ∇. A) − ∇2 A= − ∇2 A      (Coulomb Gauge).
(1.69)
Hence Cartesian components of A are solutions of Poisson equation
2 Ai = − μ0 ji
(1.70)
Using our general solution of Poisson's equation (see eq 1.38):

A(x) = μ0



j ( x′)

|xx′|
 d3x′
(1.71)

Resulting B:
B
=
∇∧A= μ0



∇∧ j ( x′)

|xx′|
 d3x′
=
μ0



j (x′) ∧∇ 1

|xx′|
 d3x′ = μ0



j( x′) ∧( xx′)

|xx′|3
 d3x′
(1.72)
This is the distributed-current version of the law of Biot and Savart (dating from  ∼ 1820). For a wire carrying current I the integral over volume j is replaced by the integral I dl i.e.

B= μ0



( xx′)

|xx′|3
j  d3 x = μ0



( xx′)

|xx′|3
∧I dl        .
(1.73)

The Biot-Savart law gives us a direct means to calculate B by integrating over j(x′), numerically if necessary. However this integration brute-force method is excessively computationally intensive and if symmetries are present in the problem we can use them to simplify.

1.5.4  Cartesian Translational Symmetry (2-d x,y)

figure/filament.gif
Figure 1.19: (a) The coordinates with respect to an infinite straight filament carrying current I, and (b) the contour and surface for use with Ampere's law.
If we consider a situation where ∂/∂z = 0, corresponding to infinite straight parallel currents in z-direction j = j(x,y) z. Our general vector potential solution shows us immediately that A = A z, Ax = Ay = 0. (Assuming A, B→ 0 at ∞, i.e. no `external' sources.) That fact tells us that Bz = (∇∧A)z = 0. We can consider the elementary building block of this problem to be the single infinitesimal filament. Formally j = I δ(x) δ(y). We could calculate B(x) by integrating over this filament. Far easier to use Ampere's Law directly



S 
( ∇∧B) . ds =


C 
B. dl =


S 
μ0 j . ds = μ0 I
(1.74)
By symmetry (∫)C B. dl = 2πr Bθ
So
Bθ = μ0 I

2πr
    .
(1.75)
Also Br = 0 by applying Gauss's Theorem to a volume (of unit length in z-dir)
0 =


V 
∇. B d3x =


S 
B. dS = 2πr Br
(1.76)
by symmetry.
Thus Maxwell's equations immediately show us what the 2-d Green function solving ∇∧B = μ0 I z δ(xx′) is
B=
^
θ
 
μ0 I

2 π
1




( x−x′)2 + ( y−y′)2
(1.77)
Any general j(x,y) can be handled by 2-d integration using this function.

1.5.5  Cylindrical Symmetry (Circular Loops with common axis)

If there exist cylindrical coordinates (r, θ, z) such that ∂/∂θ = 0, j = j θ. Then by symmetry A = Aθ θ, Bθ = 0. This situation turns out to be soluble analytically but only in terms of the special functions known as Elliptic Integrals.
figure/hoop.gif
Figure 1.20: Cylindrical Coordinates near a circular current-carrying filament.
If
jθ = I δ( r − a ) δ( z )
(1.78)
then
Aθ (r,z) = μ0

4 π
2   ⎛


a

r
 

( 2 − k2 ) K(k) − 2 E(k)

k

(1.79)
where
k2 4 ra

( r + a )2 + z2
(1.80)
and K, E are the complete elliptical integrals of the first and second kind. This general form is so cumbersome that it does not make general analytic calculations tractable but it makes numerical evaluation easier by using canned routines for K(k) & E(k). On axis (r=0) the field is much simpler
B= B
^
z
 
= μ0 I

4 π
2 πa2

( z2 + a2 )3/2
^
z
 
       .
(1.81)

1.5.6  General Property of Symmetry Situations: Flux Function

When there is a symmetry direction, the component of B perpendicular to that direction can be expressed in terms of a "flux function". The magnetic flux between two positions is defined as the B-field flux crossing a surface spanning the gap (per unit length if translational).
figure/translation.gif
Figure 1.21: Path from a reference point to a field point defines a surface to which Stokes' theorem is applied, in a situation of translational symmetry.
Since ∇. B = 0 it does not matter how the surface gets from the ref-point to P (provided it stays symmetric). So the function
ψ ≡


S 
B. dS
(1.82)
is well defined. For translational (z) symmetry, a consequence is
B = −
^
z
 
∧∇ψ
(1.83)
This arises because
ψ =


S 
B. dS =
∇∧A. dS = Az(P) − Az(0)
(1.84)
So really ψ is identical to the z-component of the vector potential and
B= ∇∧A
=
Bz
^
z
 
+ ∇∧
Az
^
z
 

= Bz
^
z
 
+ B
B
=
Az ∇∧
^
z
 
+ ( ∇Az )∧
^
z
 
(1.85)
=
^
z
 
∧∇Az = −
^
z
 
∧∇ψ
B is the part of the field perpendicular to z. There could also be Bz.
For cylindrical symmetry some more variations arise from curvilinear coordinate system. There are even other symmetries, for example helical!

1.6  Electromagnetism and Magnets

1.6.1  Simple Solenoid

figure/solenoid.gif
Figure 1.22: Idealized long solenoid magnet coil.
A `Long' solenoid has a translational symmetry so B is independent of z, as well as of θ. (Except near ends). So
0 = ∇. B= 1

r

∂r
r Br +


r ∂θ

=0
 
Bθ +


∂z

=0
 
Bz
(1.86)
So
rBr = const.        and hence        Br = 0     .
(1.87)
Also
∇∧B=
1

r

∂r
r Bθ
^
z
 
+

∂r
Bz
^
θ
 
= μ0 j        (steady)
(1.88)
Inside the bore of the magnet, j = 0 so
rBθ = const.        and hence        Bθ = 0     .
(1.89)
(actually if jz = 0 everywhere then Bθ = 0 everywhere, as may be seen immediately from the Biot-Savart law). Also
∂Bz

∂r
= 0        and hence       Bz = const        .
(1.90)
Use the surface and bounding curve shown and write



S 
μ0 j . dS =


S 
∇∧B. ds =
(⎜)



C 
B. dl
(1.91)
So μ0 × current per unit length (denoted Jθ) gives
μ0 Jθ = Bz inside − Bz outside
(1.92)
But (by same approach) if B = 0 at infinity Bz outside = 0. So, inside
Bz = μ0 Jθ
(1.93)
Profile of field in coil:
figure/sol_profile.gif
Figure 1.23: The field profile within the conductor region of the coil depends on the current-density profile.
is determined by the current density in the coil:
dBz

dr
= −μ0 jθ
(1.94)

Bzb − Bza = −
b

a 
μ0 jθ dr = − μ0 Jθ     .
(1.95)
(as before). Notice that all this is independent of coil thickness (b−a). Coils are usually multi-turn so
Bz = μ0 n I
(1.96)
where n is turns per unit length, I is current in each turn.
Jθ = n I        .
(1.97)

1.6.2  Solenoid of Arbitrary Cross-Section

figure/arb_solen.gif
Figure 1.24: Solenoid of arbitrary cross-section.


∂z
= 0
(1.98)
Consider Biot-Savart Law, expressed as vector potential:
A(r) = μ0



j ( r′)

|rr′|
 d3 r′       .
(1.99)
If currents all flow in azimuthal direction, i.e. jz = 0, then Az = 0.
⇒        Bx = By = 0        (everywhere.)
(1.100)
Then integral form of Ampere's law is still
Bz (inside) = μ0 Jp
(1.101)
where Jp is total current in azimuthal direction per unit length.

1.6.3  Coil Types

(a) Wire (Filament):

figure/wire_coil.gif
Figure 1.25: Section through a wire-wound magnet coil.
Multiple layers wound on a former. Usually only for low-field low-current work.

(b) Tape wound:

figure/tape_coil.gif
Figure 1.26: Tape-wound coils are stacked to produce a solenoid.
Each coil consists of a spiral-wound tape, nt turns. Many coils stacked to form a solenoid. Say nc coils per unit length n = nt . nc.

(c) Pancake:

Similar to tape but using square or rectangular conductor. (Fewer turns/coil).

(d) Plate Coils:

figure/plate_coil.gif
Figure 1.27: A picture-frame type plate coil and the configuration of a solenoid.
Each turn is made of plate. Whole is a single helix (topologically). Plates may be spaced by air or solid insulator gap. n = nc.
There are many other configurations of electromagnet, designed for a tremendous variety of applications. Most require numerical computation to determine the field and its spatial variation.

1.6.4  Magnetic Dipole

figure/local_coord.gif
Figure 1.28: Currents localized to a small region close to the origin, with the field point far away.
The magnetic field from a "localized" current distribution. Suppose we want the field at a point x which is far from the currents, in the sense that for all points x′ where j(x′) is non-negligible, |x′| << |x|, (relative to an origin near the currents). The general formula for A:
A(x) = μ0



j ( x′)

|xx′|
 d3x′
(1.102)
can be approximated by writing
1

|xx′|
= 1

(x2 − 2 x. x′+ x′2 )1/2
1

|x|

1 + x. x

|x|2
+ ...
(1.103)
so
A μ0


1

|x|


j ( x′)  d3x′+ 1

|x|2

x. xj ( x′) d3x′
   .
(1.104)
Now we convert these integrals into more convenient expressions using ∇. j = 0. Actually the first one is zero. This follows immediately from the identity
∇. ( j x) = x( ∇. j ) +( j . ∇) x= j
(1.105)
(which uses ∇x = I i.e. ∂xi/∂xj = δij, and ∇. j = 0). So

j  d3x′ =
∇′. ( j x′)  d3x′ =


S 
xj . dS = 0   ,
(1.106)
for any surface S that encloses all currents so that j = 0 on S. An intuitive way to see this is that ∫jd3x is the average velocity of charges, and must be zero. The second term is simplified using the same identity but being careful to distinguish between x and x′, and using notation ∇′ to denote the gradient operator that operates on x′, j(x), not on x.

(x. x′) j ( x′) d3x′
=

( x. x′) ∇′. ( j x′)  d3x′
=

∇′. ( j x′( x. x′)) − xj . ∇′( x′. x) d3x′
=

xj. ( ∇′x′) . x d3x′
=

xj . I . x d3x′ = −
x′( j . x)  d3 x′
(1.107)
But
x∧( x′∧j ) = ( j . x) x′− ( x. x′) j    .
(1.108)
So

x∧( x′∧j )  d3x′ = −2
x. xj  d3 x′    ,
(1.109)
by the integral relation just proved. [This identity is true for any x]. Therefore our approximation for A is
A( x) = − μ0


x

|x|3

1

2

x′∧j ( x′) d3 x′
(1.110)
or

A= μ0


mx

|x|3
    ,
(1.111)

where the Magnetic Dipole Moment of the localized current distribution is
m 1

2

x′∧j   d3x′    .
(1.112)
We have derived this expression for an arbitrary j distribution; but if the localized current is a loop current filament,
figure/dipole_loop.gif
Figure 1.29: Current-carrying loop integration to give dipole moment.

m = 1

2

x′∧j   d3x′ = 1

2

x∧I dl     .
(1.113)
If the loop is planar,
1

2
xdl = ds       ,
(1.114)
where ds is the element of surface. So m is (current × area) for a planar filament. The magnetic field is obtained from B = ∇∧A
B= μ0

4 π

3 x

|x|

x

|x|
. m
m
1

|x|3
(1.115)

1.6.5  Revisionist History of Electromagnetic Induction

Michael Faraday first showed the effect of induction: a transient current can be induced in one circuit by changes in another. This was  ∼  1830. [Faraday knew no mathematics beyond the idea of proportionality EMF ∝ rate of change of B-flux]. Suppose history had been different and we knew only the Lorentz force law:
F = q ( E + vB)
(1.116)
we could have "proved" the necessity of induction by "pure thought".
Assume Galilean Invariance: physical laws must be invariant under changes to moving coordinate systems x′ = x −vt, t′ = t. [Universally assumed in Faraday's time. Einstein doesn't come till 1905!] Consider a rigid (wire) circuit moved past a magnet:
figure/moving_coil.gif
Figure 1.30: A rigid coil moving past a steady magnet.
Each electron in the circuit (revisionist!) feels a Lorentz force
F = q ( vB)
(1.117)
as it is dragged through the magnetic field. The electric field in the rest frame of the magnet is zero. And the total electromotive force (integrated force per unit charge) round the entire circuit is
1

q

(⎜)



C 
F . dl =
(⎜)



C 
vB. dl
(1.118)
This is generally a non-zero quantity. In fact, this quantity can be transformed on the basis of purely geometrical considerations.
figure/moving_circuit.gif
Figure 1.31: Surface elements in the application of Gauss's theorem to succeeding instants of time.
Let's calculate the rate of change of total magnetic flux due to circuit motion, in a static B-field. Apply Gauss' theorem to volume shown in Fig 1.31.
0
=



V 
∇. B d3 x =


Stotal 
B. dS =


S′ 



S 
+


ribbon 
B. ds
=



S′ 
B. dS


S 
B. dS +
B. ( dlvdt )
=
dΦ+ dt
(⎜)



C 
(vB) . dl
(1.119)
[where dΦ is change in flux]. So
d Φ

dt
= −
(⎜)



C 
( vB) . dl    .
(1.120)
(pure geometry when ∂B/∂t = 0).
This equation can alternatively be obtained algebraically by writing
d Φ

dt
=
dB

dt
. dS =
( v. ∇) B. dS
(1.121)
and using
∇∧( Bv) = ( v. ∇) B+ (∇.v) B− (B.∇) v− ( ∇. B) v = ( v. ∇) B   .
(1.122)
So

dt
=
∇∧( Bv) .dS =
(⎜)



C 
( Bv) . dS
(1.123)
Anyway EMF is
1

q

(⎜)



C 
F . dl =
(⎜)



C 
(vB) . dS = −

dt
(1.124)
Now we consider the whole situation when the frame of reference is changed to one in which the circuit is stationary and the magnet is moving. By Galilean invariance the total EMF is the same, and
1

q

(⎜)

F . dl = −

dt
(1.125)
But now v = 0, and instead B is changing so

dt
=
B

∂t
. dS     .
(1.126)
In this case also the Lorentz force on the charges is
F = q ( E + vB) = q E        (since v =0)
(1.127)
There has to be an electric field in this frame of reference. And also
1

q

(⎜)

F . dl =
(⎜)

E . dl = −

dt
= −
B

∂t
. dS
(1.128)
Apply Stokes' theorem to the E . dl integral:



S 

∇∧E + B

∂t

. dS = 0     .
(1.129)
But this integral has to be zero for all S (and C) which can be true only if its integrand is everywhere zero:
∇∧E = − B

∂t
(1.130)
"Faraday's" Law (expressed in differential form) (which Faraday understood intuitively but could not have formulated in math)

1.6.6  Inductance

Suppose we have a set of circuits with currents Ii (i = 1 ... N). These are inductively coupled if the current in one gives rise to flux linking the others. Because Ampere's law is linear (Bj), the flux linking circuit j from current Ii is proportional to Ii. Consequently, the total flux linking circuit j can be written
Φj =

i 
Mji Ii
(1.131)
(Summation over Ii) different currents. M is a matrix. The element Mij is an inductance between currents i and j.. Its units are
flux

current
Wb

A
Henrys     .
(1.132)
The electromotive force or voltage Vj induced in the j'th circuit is then:
Vj = d

dt
Φj =

i 
Mji

I
 

i 
    .
(1.133)
For the simplest case N=1 circuit. Mii → L the self inductance
V = L

I
 
    .
(1.134)
It can be shown from Maxwell's equations that Mij is symmetric.

HEAD NEXT