27.07.2014 Views

Topics in Classical Electrodynamics

Topics in Classical Electrodynamics

Topics in Classical Electrodynamics

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Utrecht Summer School for Theoretical Physics<br />

<strong>Topics</strong> <strong>in</strong> <strong>Classical</strong> <strong>Electrodynamics</strong><br />

by Gleb Arutyunov<br />

Recorded by Jeroen Burgers and Marc<strong>in</strong> Dukalski


1 Summary of Electrostatics<br />

<strong>Classical</strong> electrodynamics is a theory of electric and magnetic fields caused by<br />

macroscopic distributions of electric charges and currents. In these lectures,<br />

we recapitulate the basic concepts of classical electrodynamics. With<strong>in</strong> the<br />

field of electrodynamics, one can study electromagnetic fields under certa<strong>in</strong><br />

static conditions lead<strong>in</strong>g to electrostatics (electric fields <strong>in</strong>dependent of time)<br />

and magnetostatics (magnetic fields <strong>in</strong>dependent of time). First, we focus<br />

on the laws of electrostatics. Then we derive Maxwell’s equations and study<br />

some of their solutions. We end up with the discussion of two classical<br />

radiation problems.<br />

1.1 Laws of Electrostatics<br />

Electrostatics is the study of electric fields produced by static charges. It<br />

is based entirely on Coulomb’s law. This law def<strong>in</strong>es the force that two<br />

electrically charged bodies (po<strong>in</strong>t charges) exert on each other<br />

⃗F (⃗x) = k q 1 q 2<br />

⃗x 1 − ⃗x 2<br />

|⃗x 1 − ⃗x 2 | 3 , (1)<br />

where k is Coulomb’s constant (depends on the system of units used 1 ), q 1 and<br />

q 2 are the magnitudes of the two charges, and ⃗x 1 and ⃗x 2 are their position<br />

vectors (as presented <strong>in</strong> Figure 1).<br />

One can <strong>in</strong>troduce the concept of an electric field E ⃗ as the force per unit<br />

charge<br />

F ⃗ (⃗x)<br />

⃗E (⃗x) = lim .<br />

q→0 q<br />

We have used the limit<strong>in</strong>g procedure to <strong>in</strong>troduce a test charge such that it<br />

will only measure the electric field at a certa<strong>in</strong> po<strong>in</strong>t and not create its own<br />

field. Hence, us<strong>in</strong>g Coulomb’s law, we obta<strong>in</strong> an expression for the electric<br />

1 In SI units, the Coulomb’s constant is k = 1<br />

4πɛ 0<br />

, while force is measured <strong>in</strong> newtons,<br />

charge <strong>in</strong> coulombs, length <strong>in</strong> meters, and the vacuum permittivity ɛ 0 is given by<br />

ɛ 0 = 107<br />

4πc<br />

= 8.8542 · 10 −12 F/m . Here, F <strong>in</strong>dicates farad, a unit of capacitance be<strong>in</strong>g equal<br />

2<br />

to one coulomb per volt. One can also use the Gauss system of units (CGS). In CGS units,<br />

force is expressed <strong>in</strong> dynes, charge <strong>in</strong> statcoulombs, length <strong>in</strong> centimeters, and the vacuum<br />

permittivity then reduces to ɛ 0 = 1<br />

4π . 2


q 1<br />

⃗x 1<br />

♣✁ ✁✁✁✁✁✕ ⃗x 2<br />

✲<br />

q 2<br />

Figure 1: Two charges q 1 and q 2 and their respective position vectors ⃗x 1 and ⃗x 2 .<br />

The charges exert an electric force on one another.<br />

field of a po<strong>in</strong>t charge<br />

⃗x − ⃗x′<br />

⃗E (⃗x) = kq<br />

|⃗x − ⃗x ′ | . 3<br />

S<strong>in</strong>ce ⃗ E is a vector quantity, for multiple charges we can apply the pr<strong>in</strong>ciple<br />

of l<strong>in</strong>ear superposition. Consequently, the field strength will simply be a sum<br />

of all of the contributions, which we can write as<br />

⃗E (⃗x) = k ∑ i<br />

q i<br />

⃗x − ⃗x i<br />

|⃗x − ⃗x i | . (2)<br />

3<br />

Introduc<strong>in</strong>g the electric charge density ρ (⃗x), the electric field for a cont<strong>in</strong>uous<br />

distribution of charge is given by<br />

∫<br />

⃗E (⃗x) = k ρ (⃗x ′ ⃗x − ⃗x′<br />

)<br />

|⃗x − ⃗x ′ | 3 d3 x ′ . (3)<br />

The Dirac delta-function (distribution) allows one to write down the electric<br />

charge density which corresponds to local charges<br />

N∑<br />

ρ (⃗x) = q i δ (⃗x − ⃗x i ) . (4)<br />

i=1<br />

Substitut<strong>in</strong>g this formula <strong>in</strong>to eq.(3), one recovers eq.(2). However, eq.(3)<br />

is not very convenient for f<strong>in</strong>d<strong>in</strong>g the electric field. For this purpose, one<br />

typically turns to another <strong>in</strong>tegral relation known as the Gauss theorem,<br />

which states that the flux through an arbitrary surface is proportional to the<br />

charge conta<strong>in</strong>ed <strong>in</strong>side it. Let us consider the flux of E ⃗ through a small<br />

region of surface dS, represented graphically <strong>in</strong> Figure 1.2,<br />

dN = ( )<br />

E ⃗<br />

q<br />

· ⃗n dS = (⃗r · ⃗n) dS<br />

r3 = q r cos (⃗r, ⃗n) dS = q 2 r 2 dS′ ,<br />

3


❈❈ ❈<br />

❈ ❈<br />

❈ ❈<br />

❈<br />

⃗E<br />

❈<br />

✓ ✓✓✼<br />

✟ ✟✟✯ ⃗n<br />

❈ ❈<br />

❈ ❈<br />

❈ ❈<br />

q ✟✟✟✟✟✟✟✟✟<br />

✂ ✂✂✂✂✂✂✂<br />

Figure 2: The electric flux through a surface, which is proportional to the charge<br />

with<strong>in</strong> the surface.<br />

where on the first step we have used that ⃗ E = q ⃗r<br />

r 3 . By the def<strong>in</strong>ition of dS ′ ,<br />

we observe that it is positive for an angle θ between ⃗ E and ⃗n less than π 2 and<br />

negative otherwise. We <strong>in</strong>troduce the solid angle dΩ ′<br />

dΩ ′ = dS′<br />

r 2 . (5)<br />

Plugg<strong>in</strong>g this relation <strong>in</strong>to eq.(5) leaves us with the follow<strong>in</strong>g expression for<br />

the flux<br />

dN = q · dΩ ′ . (6)<br />

By <strong>in</strong>tegrat<strong>in</strong>g eq.(6), we obta<strong>in</strong> the follow<strong>in</strong>g equation for the flux N<br />

∫<br />

{<br />

( ) ⃗E 4πq if q is <strong>in</strong>side the surface<br />

· ⃗n dS =<br />

0 otherwise<br />

S<br />

Equivalently, us<strong>in</strong>g the fact that the <strong>in</strong>tegral of the charge distribution over<br />

∫<br />

volume V is equal to the total charge enclosed <strong>in</strong> the volume, i.e. q =<br />

ρ (x) V d3 x, one f<strong>in</strong>ds a similar expression<br />

∫<br />

∫<br />

( )<br />

N = ⃗E · ⃗n dS = 4π ρ(x) d 3 x .<br />

S<br />

By mak<strong>in</strong>g use of the Gauss-Ostrogradsky theorem, one may rewrite the<br />

above <strong>in</strong>tegral <strong>in</strong> terms of the volume <strong>in</strong>tegral of the divergence of the vector<br />

field E ⃗ ∫<br />

∫<br />

( ) ⃗E · ⃗n dS = div E ⃗ (⃗x) d 3 x .<br />

S<br />

4<br />

V


Recall<strong>in</strong>g that the left hand side is equal to 4πq, a relation between the<br />

divergence of the electric field and the charge density arises<br />

∫ [<br />

0 = div E ⃗ ]<br />

(⃗x) − 4πρ (⃗x) d 3 x .<br />

V<br />

S<strong>in</strong>ce the relation holds for any chosen volume, then the expression <strong>in</strong>side<br />

the <strong>in</strong>tegral must equal to zero. The result<strong>in</strong>g equation is then<br />

div ⃗ E (⃗x) = 4πρ (⃗x) .<br />

This is known as the differential form of the Gauss (law) theorem for electrostatics.<br />

This is the first equation from the set of four Maxwell’s equations,<br />

the latter be<strong>in</strong>g the essence of electrodynamics.<br />

The Gauss theorem is not enough, however, to determ<strong>in</strong>e all the components<br />

of ⃗ E. A vector field ⃗ A is known if its divergence and its curl, denoted<br />

as div ⃗ A and rot ⃗ A respectively, are known. Hence, some <strong>in</strong>formation is necessary<br />

about the curl of electric field. This is <strong>in</strong> fact given by the second<br />

equation of electrostatics<br />

rot ⃗ E = 0 . (7)<br />

The second equation of electrostatics is known as Faraday’s law <strong>in</strong> the absence<br />

of time-vary<strong>in</strong>g magnetic fields, which are obviously not present <strong>in</strong><br />

electrostatics (s<strong>in</strong>ce we required all fields to be time <strong>in</strong>dependent). We will<br />

derive this equation <strong>in</strong> the follow<strong>in</strong>g way. Start<strong>in</strong>g from the def<strong>in</strong>ition of the<br />

electric field (Coulomb’s law) given by equation (3), we rewrite it <strong>in</strong> terms<br />

of a gradient and pull the differential operator outside of the <strong>in</strong>tegral<br />

∫<br />

∫<br />

⃗E (⃗x) = ρ (⃗x ′ ⃗x − ⃗x′<br />

)<br />

|⃗x − ⃗x ′ | 3 d3 x ′ = − ρ (⃗x ′ ) ∇ ⃗ 1<br />

x<br />

|⃗x − ⃗x ′ | d3 x ′<br />

∫<br />

∫<br />

= −∇ ⃗ x<br />

ρ (⃗x ′ )<br />

|⃗x − ⃗x ′ | d3 x ′ = −grad<br />

ρ(⃗x ′ )<br />

|⃗x − ⃗x ′ | d3 x ′ . (8)<br />

From vector calculus we know that the curl of gradient is always equal to<br />

zero, such that<br />

rot (grad f) = 0 ⇒ rot ⃗ E = 0 .<br />

5


This derivation shows that the vanish<strong>in</strong>g of rot ⃗ E is not related to the <strong>in</strong>verse<br />

square law. It also shows that the electric field is the m<strong>in</strong>us gradient of some<br />

scalar potential<br />

⃗E(⃗x) = −grad ϕ .<br />

From the above, it then follows that this scalar potential is given by<br />

∫<br />

ρ(x ′ )<br />

ϕ(x) =<br />

|x − x ′ | d3 x ′ ,<br />

where the <strong>in</strong>tegration is carried out over the entire space. Obviously, the<br />

scalar potential is def<strong>in</strong>ed up to an additive constant; add<strong>in</strong>g any constant<br />

to a given ϕ(x) does not change the correspond<strong>in</strong>g electric field ⃗ E.<br />

What is the physical <strong>in</strong>terpretation of ϕ(x)? Consider the work which has<br />

to be done to move a test charge along a path from po<strong>in</strong>t A to B through an<br />

electric field E ⃗<br />

∫ B<br />

W = −<br />

A<br />

∫ B<br />

⃗F · d ⃗ l = −q<br />

A<br />

⃗E · d ⃗ l .<br />

The m<strong>in</strong>us sign represents the fact that the test charge does work aga<strong>in</strong>st<br />

the electric forces. By associat<strong>in</strong>g the electric field as the gradient of a scalar<br />

potential, one obta<strong>in</strong>s<br />

∫ B<br />

∫ B<br />

W = q gradϕ · d ⃗ l = q<br />

A<br />

A<br />

= q<br />

∫ B<br />

A<br />

∂ϕ<br />

∂x<br />

∂ϕ ∂ϕ<br />

dx + dy +<br />

∂y ∂z dz<br />

dϕ = q (ϕ B − ϕ A ) .<br />

The result is just a difference between the potentials at the end po<strong>in</strong>ts of the<br />

path. This implies that the potential energy of a test charge is given by<br />

V = q ϕ .<br />

In other words, the potential energy does not depend on the choice of path<br />

(hence, the electric force is a conservative force). If a path is chosen such<br />

that it is closed, i.e. A = B, the <strong>in</strong>tegral reduces to zero<br />

∮<br />

⃗E · d ⃗ l = 0 .<br />

6


B<br />

d ⃗ l<br />

✐<br />

✄✎<br />

✄ ⃗E<br />

<br />

A<br />

✎ ✎ ✎ ✎ ✎ ✎<br />

Figure 3: The work that has to be done over a charged particle to move it along<br />

the path from A to B through an electric field ⃗ E.<br />

This result can also be obta<strong>in</strong>ed from Stokes’ theorem<br />

∮ ∫<br />

( ) ⃗E · d ⃗ l = rot E ⃗ · dS = 0 ,<br />

where we have used the fact that rot ⃗ E = 0.<br />

S<br />

To summarize, we have derived two laws of electrostatics <strong>in</strong> the differential<br />

form<br />

⃗∇ · ⃗E (⃗x) = div ⃗ E (⃗x) = 4πρ (⃗x) , (9)<br />

⃗∇ × ⃗ E (⃗x) = rot ⃗ E (⃗x) = 0 . (10)<br />

1.2 Charged Surfaces<br />

The electric field ⃗ E (⃗x) of charged surfaces can be computed by us<strong>in</strong>g the<br />

Gauss theorem. Let us def<strong>in</strong>e the surface charge density<br />

∆q<br />

σ (⃗x) = lim<br />

∆S→0 ∆S = dq<br />

dS .<br />

One considers the flux N through such an arbitrary surface, and the charge<br />

enclosed <strong>in</strong> this surface S is q = S · σ: the charge density times the area of<br />

the surface. As shown <strong>in</strong> Figure 1.4, we consider a prism through which the<br />

electric field passes. The height of the prism is denoted by the paramater<br />

7


⃗n 2<br />

✻<br />

❅ ❅ <br />

∆S<br />

✚ ✚✚✚✚✚✚✚✚✚✚✚ ❄<br />

∆q ⃗n 1<br />

✚ ✚✚✚✚✚✚✚✚✚✚✚<br />

Figure 4: The flux through a small surface element ∆S with charge ∆q.<br />

dl. From this, the flux can be computed. First, consider the Gauss theorem<br />

which allows one to write<br />

∮<br />

N = E n dS = 4πq = 4πS σ (⃗x) .<br />

As before, we can express the flux <strong>in</strong> terms of its components as follows<br />

N = E 1 cos( ⃗ E 1 , ⃗n 1 )S + E 2 cos( ⃗ E 2 , ⃗n 2 ) S + N ′ .<br />

Here N ′ is the contribution from the sides (horizontal flux). Note that the<br />

vectors ⃗n 1 = −⃗n and ⃗n 2 = ⃗n which can be observed <strong>in</strong> the figure. Now by<br />

lett<strong>in</strong>g the height dl of the prism approach zero, we ma<strong>in</strong>ta<strong>in</strong> that the prism<br />

stretches above and below the surface, yet the horizontal contributions N ′<br />

become negligible. This leaves the follow<strong>in</strong>g simplified relation<br />

N = (−E 1n + E 2n ) · S = 4πσ · S ,<br />

where E 1n and E 2n are projections of ⃗ E on ⃗ N. The flux is then just a measure<br />

of the jump <strong>in</strong> the electric fields through the charged surface. By remov<strong>in</strong>g<br />

common factors on both sides, we arrive at the follow<strong>in</strong>g expression<br />

E n2 − E n1 = 4πσ(⃗x) .<br />

8


Thus, normal components of ⃗ E at two close po<strong>in</strong>ts separated by a charged<br />

surface differ from each other by 4πσ.<br />

1.3 Laplace and Poisson Equations<br />

In the previous section it was shown that the curl of the electric field is equal<br />

to zero, thus the field is simply the gradient of some scalar function, which<br />

can be written as<br />

rot ⃗ E (⃗x) = 0 ⇒ ⃗ E (⃗x) = −∇ϕ (⃗x) .<br />

Substitut<strong>in</strong>g the right hand side of this expression <strong>in</strong>to equation (9), we<br />

obta<strong>in</strong><br />

div ∇ϕ (⃗x) = −4πρ (⃗x) .<br />

This gives<br />

∇ 2 ϕ (⃗x) ≡ ∆ϕ (⃗x) = −4πρ (⃗x) . (11)<br />

Equation (11) is known as the Poisson equation. In case ρ (⃗x) = 0, i.e. <strong>in</strong> a<br />

region of no charge, the left hand side of (11) is zero, which is known as the<br />

Laplace equation. Substitut<strong>in</strong>g <strong>in</strong>to (11) the form scalar potential ϕ, given<br />

by (8) , we get<br />

∇ 2 ϕ (⃗x) = ∇ 2 ∫<br />

ρ(⃗x ′ ∫<br />

)<br />

|⃗x − ⃗x ′ | d3 x ′ =<br />

( ) 1<br />

d 3 x ′ ρ(⃗x ′ )∇ 2 .<br />

|⃗x − ⃗x ′ |<br />

Without loss of generality we can take x ′ = 0, which is equivalent to choos<strong>in</strong>g<br />

the orig<strong>in</strong> of our coord<strong>in</strong>ate system. By switch<strong>in</strong>g to spherical coord<strong>in</strong>ates,<br />

we can show that<br />

∇ 2 1<br />

|⃗x − ⃗x ′ | = ∇2 1 r = 1 r<br />

d<br />

(r 2 1 )<br />

= 0 .<br />

dr 2 r<br />

This is true everywhere except for r = 0, for which the expression above is<br />

undeterm<strong>in</strong>ed. To determ<strong>in</strong>e its value at r = 0 we can use the follow<strong>in</strong>g trick.<br />

Integrat<strong>in</strong>g over volume V and us<strong>in</strong>g the Gauss law, one obta<strong>in</strong>s<br />

∫<br />

∇ 2 ( 1<br />

r<br />

)<br />

d 3 x =<br />

∫<br />

div ∇<br />

( 1<br />

r<br />

)<br />

d 3 x =<br />

∮<br />

⃗n · ∇ 1 r · dS<br />

V<br />

=<br />

V<br />

∮<br />

S<br />

⃗n ·<br />

( )<br />

∂ 1<br />

⃗ndS =<br />

∂r r<br />

S<br />

∮<br />

S<br />

( )<br />

∂ 1<br />

r 2 · dΩ = −4π .<br />

∂r r<br />

9


Therefore,<br />

or<br />

∇ 2 x<br />

( ) 1<br />

∇ 2 = −4πδ(x) ,<br />

r<br />

1<br />

|⃗x − ⃗x ′ | = −4πδ (⃗x − ⃗x′ ) .<br />

Thus, we f<strong>in</strong>d<br />

∫<br />

∇ 2 ϕ =<br />

ρ(x ′ ) (−4πδ(x − x ′ )) d 3 x ′ = −4πρ(x) .<br />

Hence, we have proved that 1 solves the Poisson equation with the po<strong>in</strong>t<br />

r<br />

charge source. In general, the functions satisfy<strong>in</strong>g ∇ ⃗ 2 ϕ = 0 are called harmonic<br />

functions.<br />

1.4 The Green Theorems<br />

If <strong>in</strong> electrostatics we would always deal with discrete or cont<strong>in</strong>uous distributions<br />

of charges without any boundary surfaces, then the general expression<br />

(where one <strong>in</strong>tegrates over all of space)<br />

∫<br />

ϕ(x) =<br />

ρ(x ′ d 3 x ′<br />

)<br />

|x − x ′ |<br />

(12)<br />

would be the most convenient and straightforward solution of the problem. In<br />

other words, given some distribution of charge, one can f<strong>in</strong>d the correspond<strong>in</strong>g<br />

potential and, hence, the electric field ⃗ E = −∇ϕ.<br />

In reality, most of the problems deals with f<strong>in</strong>ite regions of space (conta<strong>in</strong><strong>in</strong>g<br />

or not conta<strong>in</strong><strong>in</strong>g the charges), on the boundaries of which def<strong>in</strong>ite<br />

boundary conditions are assumed. These boundary conditions can be created<br />

by a specially chosen distribution of charges outside the region <strong>in</strong> question.<br />

In this situation our general formula (12) can not be applied with the exception<br />

of some particular cases (as <strong>in</strong> the method of images). To understand<br />

boundary problems, one has to <strong>in</strong>voke the Green theorems.<br />

10


Consider an arbitrary vector field 2 A. ⃗ We have<br />

∫<br />

∮<br />

div A ⃗ ( )<br />

d 3 x = ⃗A · ⃗n dS . (13)<br />

V<br />

S<br />

Let us assume that ⃗ A has the follow<strong>in</strong>g specific form<br />

⃗A = ϕ · ⃗∇ψ ,<br />

where ψ and ϕ are arbitrary functions. Then<br />

div ⃗ A = div<br />

(<br />

ϕ · ⃗∇ψ<br />

)<br />

= div<br />

= ⃗ ∇ϕ · ⃗∇ψ + ϕ∇ 2 ψ .<br />

(<br />

ϕ ∂ψ )<br />

= ∂ (<br />

ϕ ∂ψ )<br />

∂x i ∂x i ∂x i<br />

Substitut<strong>in</strong>g this back <strong>in</strong>to eq.(13), we get<br />

∫ ( ∮ ( ) ∮<br />

⃗∇ϕ · ∇ψ ⃗ + ϕ∇ ψ)<br />

2 d 3 x = ϕ · ⃗∇ψ · ⃗n dS =<br />

V<br />

S<br />

S<br />

ϕ<br />

( ) dψ<br />

dS .<br />

dn<br />

which is known as the first Green formula. When we <strong>in</strong>terchange ϕ for ψ <strong>in</strong><br />

the above expression and take a difference of these two we obta<strong>in</strong> the second<br />

Green formula<br />

∫<br />

V<br />

(<br />

ϕ∇ 2 ψ − ψ∇ 2 ϕ ) ∮<br />

d 3 x =<br />

S<br />

(<br />

ϕ dψ<br />

dn − ψ dϕ )<br />

dS . (14)<br />

dn<br />

By us<strong>in</strong>g this formula, the differential Poisson equation can be reduced to an<br />

<strong>in</strong>tegral equation. Indeed, consider a function ψ such that<br />

ψ ≡ 1 R = 1<br />

|⃗x − ⃗x ′ |<br />

⇒ ∇ 2 ψ = −4πδ (⃗x) . (15)<br />

Substitut<strong>in</strong>g it <strong>in</strong>to the second Green formula (14) and assum<strong>in</strong>g x is <strong>in</strong>side<br />

the space V <strong>in</strong>tegrated over, one gets<br />

∫ (<br />

) ∮ [<br />

−4πϕ(⃗x ′ )δ (⃗x − ⃗x ′ ) + 4πρ(⃗x′ )<br />

d 3 x ′ = ϕ d ( ) 1<br />

− 1 ]<br />

dϕ<br />

dS ′ .<br />

|⃗x − ⃗x ′ |<br />

dn ′ R R dn ′<br />

V<br />

S ′<br />

2 Now <strong>in</strong>troduced for mathematical convenience, but it will later prove to be of greater<br />

importance.<br />

11


Here we have chosen ϕ (⃗x ′ ) to satisfy the Poisson equation ∆ϕ (⃗x ′ ) = −4πρ (⃗x ′ ).<br />

By us<strong>in</strong>g the sampl<strong>in</strong>g property of the delta function, i.e. ∫ ϕ V (⃗x′ ) δ (⃗x − ⃗x ′ ) =<br />

ϕ (⃗x), the expression above allows one to express ϕ(x) as<br />

∫<br />

ϕ (⃗x) =<br />

V<br />

ρ (⃗x ′ )<br />

R d3 x ′ + 1 ∮ [ 1<br />

4π S R<br />

∂ϕ<br />

∂n − ϕ ∂ ( 1 ′ ∂n ′ R<br />

)]<br />

dS ′ , (16)<br />

which is the general solution for the scalar potential. The terms <strong>in</strong>side the<br />

<strong>in</strong>tegrals are equal to zero if x lies outside of V .<br />

Consider the follow<strong>in</strong>g two special cases:<br />

• If S goes to ∞ and the electric field vanishes on it faster than 1 R ,<br />

then the surface <strong>in</strong>tegral turns to zero and ϕ(x) turns <strong>in</strong>to our general<br />

solution given by eq.(12).<br />

• For a volume which does not conta<strong>in</strong> charges, the potential at any po<strong>in</strong>t<br />

(which gives a solution of the Laplace equation) is expressed <strong>in</strong> terms<br />

of the potential and its normal derivative on the surface enclos<strong>in</strong>g the<br />

volume. This result, however, does not give a solution of the boundary<br />

problem, rather it represents an <strong>in</strong>tegral equation, because given ϕ and<br />

∂ϕ<br />

∂n<br />

(Cauchy boundary conditions) we overdeterm<strong>in</strong>ed the problem.<br />

Therefore, the question arises which boundary conditions should be imposed<br />

to guarantee a unique solution to the Laplace and Poisson equations.<br />

Experience shows that given a potential on a closed surface uniquely def<strong>in</strong>es<br />

the potential <strong>in</strong>side (e.g. a system of conductors on which one ma<strong>in</strong>ta<strong>in</strong>s<br />

different potentials). Giv<strong>in</strong>g the potential on a closed surface corresponds to<br />

the Dirichlet boundary conditions.<br />

Analogously, given an electric field (i.e. normal derivative of a potential)<br />

or likewise the surface charge distribution (E ∼ 4πσ) also def<strong>in</strong>es a unique<br />

solution. These are the Neumann boundary conditions 3 .<br />

One can prove, with the help of the first Green formula, that the Poisson<br />

equation<br />

⃗∇ 2 ϕ = −4πρ ,<br />

3 Note that both Dirichlet as well as Neumann boundary conditions are not only limited<br />

to electrodynamics, but are more general and appear throughout the field of ord<strong>in</strong>ary or<br />

partial differential equations.<br />

12


<strong>in</strong> a volume V has a unique solution under the Dirichlet or the Neumann<br />

conditions given on a surface S enclos<strong>in</strong>g V . To do so, assume there exist two<br />

different solutions ϕ 1 and ϕ 2 which both have the same boundary conditions.<br />

Consider<br />

U = ϕ 2 − ϕ 1 .<br />

It solves ∇ 2 U = 0 <strong>in</strong>side V and has either U = 0 on S (Dirichlet) or ∂U = 0 ∂n<br />

on S (Neumann). In the first Green formula one plugs ϕ = ψ = U, so that<br />

∫<br />

V<br />

(<br />

|∇U| 2 + U∇ 2 U ) ∮<br />

d 3 x =<br />

S<br />

U<br />

( ∂U<br />

∂n<br />

)<br />

dS . (17)<br />

Here the second term <strong>in</strong> the <strong>in</strong>tegral vanishes as ⃗ ∇ 2 U = 0 by virtue of be<strong>in</strong>g<br />

the solution to the Laplace equation and the right hand side is equal to<br />

zero, s<strong>in</strong>ce we have assumed that the value of the potential (Dirichlet) or its<br />

derivative (Neumann) vanish at the boundary. This equation is true iff 4<br />

∫<br />

V<br />

|∇U| 2 = 0 −→ |∇U| = 0<br />

−→ ⃗ ∇U = 0 (18)<br />

Thus, <strong>in</strong>side V the function U is constant everywhere. For Dirichlet boundary<br />

conditions U = 0 on the boundary and so it is zero uniformly, such that ϕ 1 =<br />

ϕ 2 everywhere, i.e. there is only one solution. Similarly, the solution under<br />

Neumann boundary conditions is also unique up to unessential boundary<br />

terms.<br />

1.5 Method of Green Functions<br />

This method is used to f<strong>in</strong>d solutions for many second order differential<br />

equations and provides a formal solution to the boundary problems. The<br />

method is based on an impulse from a source, which is later <strong>in</strong>tegrated with<br />

the source function over entire space. Recall<br />

4 “If and only if”.<br />

∇ 2 1<br />

|⃗x − ⃗x ′ | = −4πδ (⃗x − ⃗x′ ) . (19)<br />

13


However, the function<br />

1<br />

|⃗x−⃗x ′ | is just one of many functions which obeys ∇2 ψ =<br />

−4πδ (⃗x − ⃗x ′ ). The functions that are solutions of this second order differential<br />

equation are known as Green’s functions. In general,<br />

⃗∇ 2 G (⃗x, ⃗x ′ ) = −4πδ (⃗x − ⃗x ′ ) , (20)<br />

where G (⃗x, ⃗x ′ ) = 1 + F (⃗x, |⃗x−⃗x ′ | ⃗x′ ), so that ∇ ⃗ 2 F (⃗x, ⃗x ′ ) = 0, i.e. it obeys the<br />

Laplace equation <strong>in</strong>side V . The po<strong>in</strong>t is now to f<strong>in</strong>d such F (⃗x, ⃗x ′ ), that gets<br />

rid of one of the terms <strong>in</strong> the <strong>in</strong>tegral equation (16) we had for ϕ (⃗x). Lett<strong>in</strong>g<br />

ϕ = ϕ (⃗x) and ψ = G (⃗x, ⃗x ′ ), we then get<br />

∫<br />

ϕ (⃗x) =<br />

V<br />

ρ (⃗x ′ ) G (⃗x, ⃗x ′ ) d 3 x ′ + 1<br />

4π<br />

∮<br />

S<br />

[<br />

G (⃗x, ⃗x ′ ) ∂ϕ (⃗x′ )<br />

− ϕ (⃗x ′ ) ∂G (⃗x, ]<br />

⃗x′ )<br />

dS ′ .<br />

∂n ′ ∂n ′<br />

By us<strong>in</strong>g the arbitrar<strong>in</strong>ess <strong>in</strong> the def<strong>in</strong>ition of the Green function we can leave<br />

<strong>in</strong> the surface <strong>in</strong>tegral the desired boundary conditions. For the Dirichlet case<br />

we can choose G boundary (⃗x, ⃗x ′ ) = 0, when ⃗x ′ ∈ S, then ϕ(⃗x) simplifies to<br />

∫<br />

ϕ (⃗x) =<br />

V<br />

ρ (⃗x ′ ) G (⃗x, ⃗x ′ ) d 3 x ′ − 1<br />

4π<br />

∫<br />

S<br />

ϕ (⃗x ′ ) ∂G (⃗x, ⃗x′ )<br />

∂n ′ dS ′ ,<br />

where G (⃗x, ⃗x ′ ) is referred to as the bulk-to-bulk propagator and ∂G(⃗x,⃗x′ )<br />

is<br />

∂n ′<br />

the bulk-to-boundary propagator.<br />

For the Neumann case we could try to choose ∂G(⃗x,⃗x′ )<br />

= 0 when ⃗x ′ ∈ S.<br />

∂n<br />

However, one has<br />

∮ ∂G (⃗x, ⃗x ′ ∫<br />

)<br />

( ) ∫<br />

∫<br />

dS ′ = ⃗∇G · ⃗n dS ′ = div∇G ∂n ⃗ d 3 x ′ = ∇ 2 G d 3 x ′<br />

′ S<br />

∫<br />

= −4π δ(x − x ′ ) d 3 x ′ = −4π . (21)<br />

For this reason we can not demand ∂G(⃗x,⃗x′ )<br />

= 0. Instead, one chooses another<br />

∂n<br />

simple condition ∂G(⃗x,⃗x′ )<br />

= − 4π , where S is the total surface area, and the<br />

∂n<br />

S<br />

left hand side of the equation is referred to as the Neumann Green function.<br />

Us<strong>in</strong>g this condition:<br />

∫<br />

ϕ (⃗x) = ρ (⃗x ′ ) G N (x, x ′ ) d 3 x ′ + 1 ∮<br />

G N (⃗x, ⃗x ′ ) ∂ϕ (⃗x′ )<br />

dS ′<br />

V<br />

4π S<br />

∂n ′<br />

+ 1 ∫<br />

ϕ (⃗x ′ ) dS ′ (22)<br />

S<br />

S<br />

14


S 1<br />

S 2<br />

Figure 5: For an arbitrary choice of surfaces S 1 and S 2 , where S is the area<br />

between them, then when we let them expand then the last term <strong>in</strong> equation 22<br />

would vanish.<br />

The last term represents 〈ϕ〉, the averaged value of the potential on S. If one<br />

takes the limit S = S 1 +S 2 → ∞, where S 1 and S 2 are two surfaces enclos<strong>in</strong>g<br />

the volume V and such that S 2 tends to <strong>in</strong>f<strong>in</strong>ity, this average disappears.<br />

1.6 Electrostatic Problems with Spherical Symmetry<br />

Frequently, when deal<strong>in</strong>g with electrostatics, one encounters the problems<br />

exhibit<strong>in</strong>g spherical symmetry. As an example, take the Coulomb law (1),<br />

which depends on the radial distance only and has no angular dependence.<br />

When encounter<strong>in</strong>g a symmetry of that sort, one often chooses a set of convenient<br />

coord<strong>in</strong>ates which greatly simplifies the correspond<strong>in</strong>g problem. It is<br />

no surprise that <strong>in</strong> this case, we will be mak<strong>in</strong>g use of spherical coord<strong>in</strong>ates,<br />

which <strong>in</strong> terms of the Cartesian coord<strong>in</strong>ates, are given by<br />

r = √ x 2 + y 2 + z 2 ,<br />

(<br />

)<br />

z<br />

θ = arccos √ , (23)<br />

x2 + y 2 + z 2<br />

( y<br />

φ = arctan ,<br />

x)<br />

To obta<strong>in</strong> the Cartesian coord<strong>in</strong>ates from the spherical ones, we use<br />

x = r s<strong>in</strong> θ cos φ ,<br />

y = r s<strong>in</strong> θ s<strong>in</strong> φ , (24)<br />

z = r cos θ .<br />

15


z<br />

θ<br />

P( r, θ,<br />

φ)<br />

φ<br />

r<br />

y<br />

x<br />

Figure 6: Spherical coord<strong>in</strong>ate system.<br />

In terms of spherical coord<strong>in</strong>ates our differential operators look different.<br />

The one we will be most <strong>in</strong>terested <strong>in</strong>, the Laplace operator, becomes<br />

⃗∇ 2 = 1 ( ∂ ∂ )<br />

r 2 ∂r r2 + 1 ( ∂<br />

∂r r 2 s<strong>in</strong> θ ∂θ s<strong>in</strong> θ ∂ )<br />

1 ∂ 2<br />

+<br />

∂θ r 2 s<strong>in</strong> 2 θ ∂φ . 2<br />

Hence, <strong>in</strong> these coord<strong>in</strong>ates the Laplace equation reads as<br />

⃗∇ 2 ϕ = 1 ∂ 2<br />

r ∂r (rϕ) + 1 (<br />

∂<br />

s<strong>in</strong> θ ∂ϕ )<br />

1 ∂ 2 ϕ<br />

+<br />

2 r 2 s<strong>in</strong> θ ∂θ ∂θ r 2 s<strong>in</strong> 2 θ ∂φ = 0 . 2<br />

We use the ansatz that ϕ (r, θ, φ) = U(r) P (θ) Q (φ). Upon substitut<strong>in</strong>g this<br />

r<br />

r<br />

<strong>in</strong>to the Laplace equation and multiply<strong>in</strong>g both sides by<br />

2 s<strong>in</strong> 2 θ<br />

, one<br />

U(r)P (θ)Q(φ)<br />

obta<strong>in</strong>s<br />

(( 1<br />

⃗∇ 2 φ = r 2 s<strong>in</strong> 2 θ<br />

U<br />

)<br />

∂ 2 U<br />

+<br />

∂r 2<br />

1<br />

r 2 s<strong>in</strong> θP<br />

( ∂<br />

∂θ<br />

))<br />

∂P<br />

s<strong>in</strong> θ + 1 ∂ 2 Q<br />

∂θ Q ∂φ . 2<br />

S<strong>in</strong>ce we only have φ dependence <strong>in</strong> the last term we can state that, s<strong>in</strong>ce<br />

there are no other terms with φ, then this term has to be constant (chosen<br />

here for convenience with anticipation of the solution)<br />

1 ∂ 2 Q<br />

Q ∂φ = 2 −m2 .<br />

16


Hence the solution is Q = e ±imφ , where m is an <strong>in</strong>teger such that Q is s<strong>in</strong>gle<br />

valued. This leaves us with two separated equations. For P (θ) the equation<br />

simplifies to<br />

(<br />

1 d<br />

s<strong>in</strong> θ dP )<br />

]<br />

+<br />

[l(l + 1) − m2<br />

s<strong>in</strong> θ dθ dθ<br />

s<strong>in</strong> 2 P = 0 ,<br />

θ<br />

and for U (r) one obta<strong>in</strong>s<br />

d 2 U l (l + 1)<br />

− U = 0 ,<br />

dr2 r 2<br />

where we have just aga<strong>in</strong> conveniently picked l(l + 1) to be the <strong>in</strong>tegration<br />

constant such that <strong>in</strong> our solution it will appear <strong>in</strong> a convenient form. It is<br />

easy to verify that the solution to the equation for U(r) is given by<br />

U (r) = Ar l+1 + Br −l ,<br />

where l is assumed to be positive and A and B are arbitrary constants. The<br />

second equation, on the other hand, is a bit more complicated and upon<br />

substitution cos θ = x it transforms <strong>in</strong>to<br />

d<br />

dx<br />

[ (1<br />

− x<br />

2 ) dP<br />

dx<br />

]<br />

+<br />

[l(l + 1) − m2<br />

1 − x 2 ]<br />

P = 0 ,<br />

which one can recognize as the so-called generalized Legendre equation. Its<br />

solutions are the associated Legendre functions. For m 2 = 0, we obta<strong>in</strong> the<br />

Legendre equation<br />

[<br />

d<br />

(1 − x 2 ) dP ]<br />

+ l(l + 1)P = 0 . (25)<br />

dx dx<br />

The solutions to this equation are referred to as the Legendre polynomials.<br />

In order for our solution to have physical mean<strong>in</strong>g, it must be f<strong>in</strong>ite and<br />

cont<strong>in</strong>uous on the <strong>in</strong>terval −1 ≤ x ≤ 1. We try as a solution the follow<strong>in</strong>g<br />

power series<br />

P (x) = x α<br />

∞<br />

∑<br />

j=0<br />

a j x j , (26)<br />

17


where α is unknown. Substitut<strong>in</strong>g our trial solution (26) <strong>in</strong>to the Legendre<br />

equation (25), we obta<strong>in</strong><br />

∞∑<br />

(<br />

(α + j) (α + j − 1) a j x α+j−2<br />

j=0<br />

− [(α + j) (α + j + 1) − l (l + 1)] a j x α+j )<br />

= 0 .<br />

For j = 0 and j = 1, the first term will have x α−2 and x α−1 and the second<br />

term will have x α and x α+1 respectively, which will never make the equation<br />

equal to zero unless<br />

• a 0 ≠ 0, then α (α − 1) = 0 so that (A) α = 0 or α = 1<br />

• a 1 ≠ 0, then α (α + 1) = 0 so that (B) α = 0 or α = −1<br />

For other j, one obta<strong>in</strong>s a recurrence relation<br />

a j+2 =<br />

(α + j) (α + j + 1) − l (l + 1)<br />

a j<br />

(α + j + 1) (α + j + 2)<br />

Cases (A) and (B) are actually equivalent. We will consider case (A) for<br />

which α = 0 or 1. The expansion conta<strong>in</strong>s only even powers of x for α = 0<br />

and only odd powers of x for α = 1. We note two properties of this series:<br />

1. The series is convergent for x 2 < 1 for any l.<br />

2. The series is divergent at x = ±1 unless it is truncated.<br />

It is obvious from the recurrent formula that the series is truncated <strong>in</strong><br />

the case that l is a non-negative <strong>in</strong>teger. The correspond<strong>in</strong>g polynomials are<br />

normalized <strong>in</strong> such a way that they are all equal to identity at x = 1. These<br />

are the Legendre polynomials P l (x):<br />

P 0 (x) = 1 ;<br />

P 1 (x) = x ;<br />

P 2 (x) = 1 (<br />

3x 2 − 1 ) ;<br />

2<br />

P 3 (x)<br />

· · ·<br />

= 1 (<br />

5x 3 − 2x ) ;<br />

3<br />

P l (x) = 1 d l (<br />

x 2 − 1 ) l<br />

.<br />

2 l l! dx l<br />

18


∆ϕ = 0<br />

S<br />

Figure 7: The field ϕ (⃗x), which obeys the Laplace equation, has no maximum or<br />

m<strong>in</strong>imum <strong>in</strong>side a region S.<br />

The general expression given <strong>in</strong> the last l<strong>in</strong>e is also known as the Rodriges<br />

formula.<br />

The Legendre polynomials form a complete system of orthogonal functions<br />

on −1 ≤ x ≤ 1. To check whether they are <strong>in</strong>deed orthogonal, one<br />

takes the differential equation for P l , multiplies it by P l ′, and then <strong>in</strong>tegrates<br />

or<br />

∫ 1<br />

P l ′<br />

−1<br />

∫ 1<br />

−1<br />

[ d<br />

dx (1 − x2 ) dP l<br />

dx + l(l + 1)P l<br />

]<br />

dx = 0 ,<br />

[<br />

(x 2 − 1) dP ]<br />

l ′ dP l<br />

dx dx + l(l + 1)P l ′P l) dx = 0 .<br />

Now subtract the same equation, but with the <strong>in</strong>terchange of l and l ′ ,<br />

such that the follow<strong>in</strong>g expression is left<br />

∫ 1<br />

[(l ′ (l ′ + 1) − l(l + 1)] P l ′P l = 0 .<br />

−1<br />

The equation above shows that for l ≠ l ′ the polynomials are orthogonal<br />

∫ 1<br />

−1<br />

P l ′P l = 0 .<br />

By us<strong>in</strong>g the Rodriges formula, one can get an identity<br />

∫ 1<br />

−1<br />

P l ′(x)P l (x)dx = 2<br />

2l + 1 δ l ′ ,l .<br />

19


For any function def<strong>in</strong>ed on −1 ≤ x ≤ 1<br />

f(x) =<br />

∞∑<br />

A l P l (x) ,<br />

l=0<br />

A l = 2l + 1<br />

2<br />

∫ 1<br />

−1<br />

f(x)P l (x)dx .<br />

Note that this expansion and its coefficients are not different to any other<br />

set of orthogonal functions <strong>in</strong> the function space. In situations where there<br />

is azimuthal symmetry, one can take m = 0. Thus,<br />

ϕ (r, θ) =<br />

∞∑ (<br />

Al r l + B l r −(l+1)) P l (cos θ) .<br />

l=0<br />

If charge is absent anywhere <strong>in</strong> the vic<strong>in</strong>ity of the coord<strong>in</strong>ate system, one<br />

should take B l = 0. Take a sphere of radius a with the potential V (θ). Then<br />

V (θ) =<br />

∞∑<br />

A l a l P l (cos θ)<br />

l=0<br />

so that<br />

A l = 2l + 1<br />

2a l<br />

∫ π<br />

0<br />

V (θ)P l (cos θ) s<strong>in</strong> θdθ .<br />

Example: f<strong>in</strong>d the potential of an empty sphere of radius r = a which<br />

has two semi-spheres with separate potentials V (θ), such that the potential<br />

is equal to V for 0 ≤ θ < π and equal to −V for π < θ ≤ π. For such a<br />

2 2<br />

system, the scalar potential is given by<br />

ϕ(r, θ) = √ V ∑ ∞<br />

(−1) j−1 (2j − 1)Γ(j − 1)<br />

(<br />

2 2 a<br />

) 2j<br />

P2j−1 (cos θ)<br />

π j! r<br />

j=1<br />

[ 3<br />

( r<br />

= V P 1 (cos θ) −<br />

2 a)<br />

7 ( r<br />

) 3<br />

P3 (cos θ) + 11 ( r<br />

) 5<br />

P5 (cos θ) − . . .]<br />

.<br />

8 a<br />

16 a<br />

Here Γ (z) for R (z) > 0 is def<strong>in</strong>ed as<br />

Γ (z) =<br />

∫ ∞<br />

0<br />

20<br />

t z−1 e −t dt .


F<strong>in</strong>ally, we would like to comment on the solutions of the Laplace equation<br />

△ϕ = 0. It is not difficult to show that one cannot have an absolute m<strong>in</strong>imum<br />

or maximum <strong>in</strong> the region (<strong>in</strong> both directions, x and y) because for an<br />

< 0<br />

imply<strong>in</strong>g that <strong>in</strong> the other direction the second derivative must have an opposite<br />

sign.<br />

extremum to exist one requires ∂ϕ<br />

∂x i<br />

= 0 which results <strong>in</strong> ∂2 ϕ<br />

∂x 2 i<br />

> 0 or ∂2 ϕ<br />

∂x 2 i<br />

2 <strong>Electrodynamics</strong><br />

Here we will treat electrodynamics as a classical relativistic field theory. We<br />

will rewrite the basic equations of electrodynamics <strong>in</strong> the manifestly Lorentzcovariant<br />

form. We will also study the correspond<strong>in</strong>g solutions.<br />

2.1 Relativistic Particle <strong>in</strong> Electormagnetic Field<br />

Let us first revisit some of the basics of special relativity written us<strong>in</strong>g tensor<br />

notation. The M<strong>in</strong>kowski metric η µν that we will use has the signature<br />

(+, −, −, −) and we will use the convention that the Lat<strong>in</strong> <strong>in</strong>dices run only<br />

over the space coord<strong>in</strong>ates (i.e. i, j, k... = 1, 2, 3), whereas the Greek <strong>in</strong>dices<br />

will <strong>in</strong>clude both time and space coord<strong>in</strong>ates (i.e. µ, ν, σ, ρ... = 0, 1, 2, 3). Additionally,<br />

<strong>in</strong> special relativity we will have to dist<strong>in</strong>guish between 3-vectors<br />

(those with only space components) and 4-vectors (hav<strong>in</strong>g both space and<br />

time components). The convention that we will use is that ⃗ A will denote a<br />

3-vector, whereas A µ will denote a 4-vector. Us<strong>in</strong>g these def<strong>in</strong>itions, we can<br />

def<strong>in</strong>e the Lorentz <strong>in</strong>variant relativistic <strong>in</strong>terval given by the expression<br />

ds 2 = x µ x µ = c 2 dt 2 − ( dx i) 2<br />

. (27)<br />

The action for a relativistic particle has the follow<strong>in</strong>g form<br />

Rewrit<strong>in</strong>g (27), we get<br />

S = −α<br />

∫ b<br />

a<br />

√<br />

∫ b<br />

ds2 = −α ds .<br />

a<br />

ds =<br />

√<br />

dx<br />

µ<br />

dt<br />

dx µ<br />

dt dt2 =<br />

√<br />

dx<br />

µ<br />

dt<br />

dx µ<br />

dt<br />

dt . (28)<br />

21


A<br />

Figure 8: The simplest form of action is given by the length of the space-time<br />

<strong>in</strong>terval between po<strong>in</strong>ts A and B.<br />

<br />

B<br />

Here we have used the convention V µ V µ = η µν V µ V ν , where η µν is the<br />

M<strong>in</strong>kowski metric.<br />

dx µ<br />

= (c, ⃗v) , ds = √ √<br />

c<br />

dt<br />

− ⃗v 2 = c 1 − ⃗v2<br />

c . 2<br />

Therefore,<br />

∫ √<br />

t1<br />

S = −αc 1 − ⃗v2<br />

t 0<br />

c dt , 2<br />

where <strong>in</strong> non-relativistic physics we assume ⃗v2 ≪ 1. In general, S =<br />

∫ c 2<br />

t1<br />

t 0<br />

L dt where L is the so-called Lagrangian of the system, which <strong>in</strong> the<br />

non-relativistic limit is given by:<br />

√<br />

(<br />

)<br />

L = −αc 1 − ⃗v2<br />

c ≈ −αc 1 − ⃗v2<br />

2 2c + · · · ≈ −αc + α ⃗v2<br />

2 2c . (29)<br />

If we want to recover the usual form of the Lagrangian L = K<strong>in</strong> Energy−<br />

V ext for a free particle V ext = 0 (hence L = 1 2 m⃗v2 ), we need to set α = mc.<br />

When we do so, equation (29) turns <strong>in</strong>to<br />

Thus, one can rewrite L as<br />

L = −mc 2 + 1 2 m⃗v2 .<br />

L = −mc √ ẋ µ ẋ µ .<br />

When we use the canonical momentum p µ def<strong>in</strong>ed as the derivative of L with<br />

respect to ẋ µ , we get<br />

p µ = ∂L<br />

∂ẋ µ = −mc ẋ µ<br />

√ẋν ẋ ν .<br />

22


A<br />

(<br />

ϕ, ⃗ A<br />

)<br />

Figure 9: In the presence of the vector potential A µ = ( ϕ, ⃗ A ) the action of a<br />

charged particle conta<strong>in</strong>s an additional term describ<strong>in</strong>g an <strong>in</strong>teraction with the<br />

vector potential.<br />

<br />

B<br />

Now when we take<br />

p 2 ≡ p µ p µ = m 2 c 2 ẋ µ ẋ µ<br />

(√ẋν ẋ ν) 2 = m2 c 2 .<br />

Hence, the particle trajectories which m<strong>in</strong>imize the action must satisfy the<br />

constra<strong>in</strong>t p 2 − m 2 c 2 = 0, which is referred to as the mass-shell condition.<br />

Let us now def<strong>in</strong>e the vector potential, which is an underly<strong>in</strong>g field (a<br />

Lorentz <strong>in</strong>variant 4-vector) <strong>in</strong> electrodynamics that we will base our further<br />

derivations on. It reads<br />

(<br />

A µ = ϕ (x) , A ⃗ )<br />

(x) .<br />

Notice that<br />

A µ → A µ = η µν A ν =<br />

(<br />

ϕ (x) , −A ⃗ )<br />

(x) .<br />

The properties of a charged particle with respect to its <strong>in</strong>teraction with<br />

electromagnetic field are characterized by a s<strong>in</strong>gle parameter: the electric<br />

charge e. The properties of the electromagnetic field itself are determ<strong>in</strong>ed by<br />

the vector A µ , the electromagnetic potential <strong>in</strong>troduced above. Us<strong>in</strong>g these<br />

quantities, one can <strong>in</strong>troduce the action of a charged particle <strong>in</strong> electromagnetic<br />

field, which has the form<br />

S = −mc<br />

∫ b<br />

a<br />

ds − e c<br />

∫<br />

A µ dx µ .<br />

Us<strong>in</strong>g Hamilton’s pr<strong>in</strong>ciple, stat<strong>in</strong>g that particles follow paths that m<strong>in</strong>imize<br />

their action (δS = 0), we can derive the equations of motion <strong>in</strong> which we<br />

neglect the back reaction of the charge on the electromagnetic field<br />

∫ dxµ<br />

0 = δS = −mc<br />

ds d(δxµ ) − e ∫<br />

[(δA µ )dx µ + A µ d(δx µ )] . (30)<br />

c<br />

23


Us<strong>in</strong>g (28), the term δs <strong>in</strong> the first <strong>in</strong>tegral becomes δds = √ dx µdδx µ<br />

, dxµ dx µ<br />

whereas <strong>in</strong> the second <strong>in</strong>tegral we have simply used the product rule of differentiation.<br />

Let us consider for a moment the U µ = dxµ term, which we will<br />

ds<br />

refer to as 4-velocity. The explicit form of U µ is<br />

⎛<br />

⎞<br />

U µ = dxµ<br />

ds = dx µ<br />

√ = ⎝ 1 ⃗v<br />

√ , √ ⎠ . (31)<br />

c 1 − ⃗v2 dt 1 − ⃗v2 c 1 − ⃗v2<br />

c 2 c 2 c 2<br />

and it has an <strong>in</strong>terest<strong>in</strong>g property that<br />

U µ U µ = dx µ<br />

ds<br />

dx µ<br />

ds = 1 .<br />

Note that this result is only valid for the signature of the metric that we<br />

chose. If we were to <strong>in</strong>vert the signature, the result would be −1 <strong>in</strong>stead.<br />

Us<strong>in</strong>g the fact that δA µ = A µ (x ν + δx ν ) − A µ (x ν ) = ∂ ν A µ δx ν + · · · , we can<br />

rewrite equation (30) as follows<br />

∫<br />

δS = mc dU µ δx µ + e ∫<br />

(∂ ν A µ dx ν δx µ − ∂ ν A µ δx ν dx µ ) = 0 .<br />

c<br />

This imposes the follow<strong>in</strong>g condition for the extremum<br />

mc dU µ<br />

ds + e c (∂ νA µ − ∂ µ A ν ) U ν = 0 .<br />

Identify<strong>in</strong>g the tensor F νµ of the electromagnetic field<br />

∂ ν A µ − ∂ µ A ν = F νµ = −F µν ,<br />

we can write the equation of motion of the charge <strong>in</strong> the electromagnetic field<br />

as follows<br />

mc dU µ<br />

ds = e c F µν U ν . (32)<br />

This expression can also be written <strong>in</strong> a more suggestive form if we def<strong>in</strong>e<br />

the momentum p µ = mcU µ (which is consistent with the requirement p 2 =<br />

m 2 c 2 s<strong>in</strong>ce U 2 = 1), so that one can express the acceleration term dU µ<br />

as<br />

= d2 x µ<br />

ds ds 2<br />

dp µ<br />

ds = dpµ dt<br />

dt ds = e c F µν U ν , (33)<br />

24


where the right hand side of the equation is referred to as the Lorentz force,<br />

whereas the left hand side is simply the rate of change of momentum with<br />

respect to the relativistic <strong>in</strong>terval. This equation is comparable with the<br />

Newtonian statement: force is the rate of change of momentum. Note that<br />

this derivation has assumed that the electromagnetic field is given (fixed)<br />

and that we vary the trajectory of the particle only (the endpo<strong>in</strong>ts rema<strong>in</strong><br />

fixed).<br />

2.2 Gauge Invariance and Maxwell’s Equations<br />

All the physical properties of the electromagnetic field as well as the properties<br />

of charge <strong>in</strong> the electromagnetic field are determ<strong>in</strong>ed not by A µ , but<br />

rather by F µν . The underly<strong>in</strong>g reason for this is that electrodynamics exhibits<br />

an important new type of symmetry 5 . To understand this issue, we<br />

may decide to change the vector potential <strong>in</strong> the follow<strong>in</strong>g way<br />

A µ → A µ + ∂ µ χ , (34)<br />

which can be rewritten <strong>in</strong> a less abstract form of space and time components<br />

separately:<br />

⃗A → ⃗ A + ⃗ ∇χ and ϕ → ϕ − 1 c<br />

∂χ<br />

∂t . (35)<br />

These transformations are referred to as the gauge transformations. Let us<br />

see what effect they have on the tensor of the electromagnetic field:<br />

δF µν = ∂ µ (A ν + ∂ ν χ) − ∂ ν (A µ + ∂ µ χ) − F µν<br />

= ∂ µ ∂ ν χ − ∂ ν ∂ µ χ = 0 . (36)<br />

Thus, the transformation (34) does not change the form of the electromagnetic<br />

field tensor. For this reason electromagnetism is a gauge <strong>in</strong>variant<br />

theory! The tensor of the electromagnetic field can be then written as<br />

⎛<br />

⎞<br />

0 E x E y E z<br />

F µν = ⎜ −E x 0 −H z H y<br />

⎟<br />

⎝ −E y H z 0 −H x<br />

⎠ (37)<br />

−E z −H y H x 0<br />

5 This symmetry extends to many other physical theories besides electrodynamics.<br />

25


and, therefore,<br />

⎛<br />

F µν = η µσ η νρ F σρ<br />

⎜<br />

⎝<br />

0 −E x −E y −E z<br />

E x 0 −H z H y<br />

E y H z 0 −H x<br />

E z −H y H x 0<br />

⎞<br />

⎟<br />

⎠ , (38)<br />

where we have def<strong>in</strong>ed the F 0i components to be the electric fields and the<br />

F ij components to the magnetic fields. From the electric and magnetic fields<br />

one can make <strong>in</strong>variants, i.e. objects that rema<strong>in</strong> unchanged under Lorentz<br />

transformations. In terms of the tensor of th electromagnetic field two such<br />

<strong>in</strong>variants are<br />

F µν F µν = <strong>in</strong>v ; (39)<br />

ε µνρσ F µν F ρσ = <strong>in</strong>v . (40)<br />

Let us <strong>in</strong>spect the gauge <strong>in</strong>variance of the electric and magnetic fields ⃗ E and<br />

⃗H, which from the form and their <strong>in</strong> terms of the electromagnetic field tensor<br />

components can be expressed <strong>in</strong> terms of the vector potential as<br />

⃗E = − ⃗ ∇ϕ − 1 c<br />

∂ ⃗ A<br />

∂t<br />

and ⃗ H = rot ⃗ A . (41)<br />

One can easily see that <strong>in</strong> the first case an extra ϕ term cancels with an extra<br />

⃗A term and <strong>in</strong> the second case we have the gauge transformation contribution<br />

vanish<strong>in</strong>g due to the fact that rot gradχ = 0. We look back at the expression<br />

for the Lorentz force and try to write it <strong>in</strong> terms of electric and magnetic<br />

fields. Rearrang<strong>in</strong>g (33), we get<br />

dp i<br />

dt<br />

=<br />

=<br />

( e<br />

c F i0 U 0 + e ) ds<br />

c F ij U j<br />

dt =<br />

⎛<br />

⎞<br />

⎝ e 1<br />

c Ei √ + e √<br />

1 −<br />

c F ij ⃗v<br />

√ ⎠ c 1 − ⃗v2<br />

⃗v2 c 1 −<br />

c . (42)<br />

2 ⃗v2<br />

c 2 c 2<br />

We can thus rewrite the expression for the Lorentz force as<br />

dp i<br />

dt = eEi + e [<br />

⃗v, H<br />

c<br />

⃗ ]<br />

. (43)<br />

26


Concern<strong>in</strong>g this result,it is <strong>in</strong>terest<strong>in</strong>g to po<strong>in</strong>t out that<br />

dE k<strong>in</strong><br />

dt<br />

= d mc 2<br />

√<br />

dt<br />

1 − v2<br />

c 2<br />

= ⃗v · dpi<br />

dt = e( ⃗ E · ⃗v<br />

)<br />

.<br />

This is the work of the electromagnetic field on the charge. Hence, the<br />

magnetic field does not play any role is k<strong>in</strong>etic energy changes, but rather<br />

only affects the direction of the movement of the particle! Us<strong>in</strong>g basic vector<br />

calculus and the def<strong>in</strong>itions of the electric and magnetic fields (41), the first<br />

two Maxwell’s equations are atta<strong>in</strong>ed<br />

div ⃗ H = div rot ⃗ A = 0 ⇒ div ⃗ H = 0 ; (44)<br />

rot E ⃗ = − 1 c rot grad ϕ − 1 ∂<br />

c ∂t rot A ⃗ ⇒ rot E ⃗ = − 1 ∂H<br />

⃗<br />

c ∂t . (45)<br />

Equation (44) is known as the no magnetic monopole rule and (45) is referred<br />

to as Faraday’s law, which we have already encountered <strong>in</strong> the previous<br />

section, but then the right hand side was suppressed due to time <strong>in</strong>dependence<br />

requirement. Together these two equations constitute the first pair of<br />

Maxwell’s equations. Notice that these are 4 equations <strong>in</strong> total, as Faraday’s<br />

law represents three equations - one for every space direction. Additionally,<br />

notice that Faraday’s law is consistent with electrostatics; if the magnetic<br />

field is time <strong>in</strong>dependent then the right hand side of the equation is equal 0,<br />

which is exactly equation (10). These equations also have an <strong>in</strong>tegral form.<br />

Integrat<strong>in</strong>g (45) over a surface S with the boundary ∂S and us<strong>in</strong>g Stokes’<br />

theorem, we arrive at<br />

∮<br />

∮<br />

rot E ⃗ · dS ⃗ = ⃗E · d ⃗ l = − 1 ∮<br />

∂<br />

⃗Hd S<br />

c ∂t<br />

⃗ . (46)<br />

S<br />

S<br />

∂S<br />

For eq.(44) one <strong>in</strong>tegrates both sides over the volume and uses the Gauss-<br />

Ostrogradsky theorem to arrive at<br />

∫<br />

∫<br />

div HdV ⃗ = ⃗H · dS ⃗ = 0 . (47)<br />

V<br />

∂V<br />

2.3 Fields Produced by Mov<strong>in</strong>g Charges<br />

Let us now consider the case where the mov<strong>in</strong>g particles produce the fields<br />

themselves. The new action will be then<br />

S = S particles + S <strong>in</strong>t + S field ,<br />

27


where we have added a new term S field , which represents the <strong>in</strong>teraction<br />

between the particles and the field that they have produced themselves. We<br />

will write is as<br />

∫<br />

∫<br />

S field ∼ F µν F µν d 4 x = F µν F µν cdt d 3 x .<br />

Then add<strong>in</strong>g the proportionality constants the total action is written as<br />

∫<br />

S = −mc ds − e ∫<br />

A µ dx µ − 1 ∫<br />

F µν F µν cdt d 3 x ,<br />

c<br />

16πc<br />

where we have adopted the Gauss system of units, i.e. µ 0 = 4π and ε 0 = 1 . 4π<br />

Note that we can rewrite the second term as<br />

e<br />

c<br />

∫<br />

∫<br />

A µ dx µ = 1 c<br />

= 1 ∫<br />

c<br />

∫<br />

ρA µ dx µ dV = 1 c<br />

j µ A µ dV dt = 1 ∫<br />

c 2<br />

dx µ<br />

ρA µ dV dt<br />

dt<br />

j µ A µ d 4 x , (48)<br />

where <strong>in</strong> the second l<strong>in</strong>e we have <strong>in</strong>troduced, the current j i = ρ dxi = (ρc, ρ⃗v).<br />

dt<br />

Includ<strong>in</strong>g this, we can now write the action of the mov<strong>in</strong>g test charge as<br />

∫<br />

S = −mc ds − 1 ∫<br />

j ρ A<br />

c 2 ρ d 4 x − 1 ∫<br />

F µν F µν cdtd 3 x .<br />

16πc<br />

Keep<strong>in</strong>g sources constant and the path unchanged (i.e. δj µ = 0 and δs = 0),<br />

we can write the deviation from the action as follows<br />

δS = − 1 ∫<br />

j ρ δA<br />

c 2 ρ d 4 x − 1 ∫<br />

F µν δF µν cdtd 3 x<br />

8πc<br />

= − 1 [ ∫ 1<br />

j ρ δA ρ d 4 x + 1 ∫ ]<br />

∂F<br />

µν<br />

c c<br />

4π ∂x δA µcdtd 3 x , (49)<br />

ν<br />

where <strong>in</strong> the last term <strong>in</strong> the first l<strong>in</strong>e, we have used that<br />

δF µν = ∂ µ δA ν − ∂ ν δA µ .<br />

To f<strong>in</strong>d the extremum, we need to satisfy δS = 0, which due to eq.(49),<br />

is equivalent<br />

− 1 c 2 jµ − 1<br />

4πc ∂µ F µν = 0 .<br />

28


Upon rearrangement, this gives us the second pair of Maxwell’s equations<br />

∂F µν<br />

∂x ν = − 4π c jµ .<br />

Notice that for vanish<strong>in</strong>g currents, these equation resemble the first pair of<br />

Maxwell’s equations given by (54), when currents are to vanish (i.e. j µ = 0).<br />

Identify<strong>in</strong>g the respective components of the electromagnetic tensor we<br />

can rewrite the second pair of Maxwell’s equations <strong>in</strong> a more familiar form<br />

rot ⃗ H = 4π c ⃗ j + 1 c<br />

∂ ⃗ E<br />

∂t<br />

and div ⃗ E = 4πρ , (50)<br />

where 4π⃗ j and 4πρ are the sources and 1 ∂E<br />

⃗ is the so-called displacement<br />

c c ∂t<br />

current. The first expression is Ampére’s law (also known as the Biot-Savart<br />

law), whereas the second one is Coulomb’s law, which we have already found<br />

before, but us<strong>in</strong>g a different pr<strong>in</strong>ciple. F<strong>in</strong>ally, we notice that the covariant<br />

conservation of the current ∂jµ = 0 is equivalent to the cont<strong>in</strong>uity equation<br />

∂x µ<br />

∂ρ<br />

∂t + divj = 0 .<br />

Below we <strong>in</strong>clude here a short digression on the tensor of the electromagnetic<br />

field. It is easy to check that, us<strong>in</strong>g the def<strong>in</strong>ition of the tensor, the<br />

follow<strong>in</strong>g is true:<br />

dF = ∂F µν<br />

∂x + ∂F νσ<br />

σ ∂x + ∂F σµ<br />

µ ∂x ν = 0 . (51)<br />

With a change of <strong>in</strong>dices, this takes the form<br />

ε µνσρ ∂F νσ<br />

∂x ρ = 0 , (52)<br />

which are four equations <strong>in</strong> disguise, s<strong>in</strong>ce we are free to pick any value of<br />

the <strong>in</strong>dex µ. Let us <strong>in</strong>troduce the so-called dual tensor<br />

Then we can rewrite equation (52) as<br />

F ∗µν = 1 2 εµνρσ F ρσ . (53)<br />

∂F ∗µν<br />

∂x ν = 0 . (54)<br />

29


Omitt<strong>in</strong>g the currents <strong>in</strong> the second pair, the first and second pair of<br />

Maxwell’s equations are similar. Indeed, we have<br />

∂F ∗µν<br />

∂x µ = 0 ,<br />

∂F µν<br />

∂x µ = 0 .<br />

The ma<strong>in</strong> difference between them is that the first pair it never <strong>in</strong>volves any<br />

currents. This has a deeper mean<strong>in</strong>g. The magnetic field, as opposed to<br />

the electric field, is an axial vector, i.e. one that does not change sign under<br />

reflection of all coord<strong>in</strong>ate axes. Thus, if there would be sources for the<br />

first pair of Maxwell equations, they should be an axial vector and a pseudoscalar<br />

6 . The classical description of particles does not allow to construct<br />

such quantities from dynamical variables associated to particle.<br />

2.4 Electromagnetic Waves<br />

When the electric charge source and current terms are absent, we obta<strong>in</strong> the<br />

electromagnetic wave solutions. In this case the Maxwell equations reduce<br />

to<br />

rotE ⃗ = − 1 ∂H<br />

⃗<br />

c ∂t , div E ⃗ = 0 ,<br />

rotH ⃗ = 1 ∂E<br />

⃗<br />

c ∂t , div H ⃗ = 0 .<br />

These equations can have non-zero solutions mean<strong>in</strong>g that the electromagnetic<br />

fields can exist without any charges or currents. Electromagnetic<br />

fields, which exist <strong>in</strong> the absence of any charges, are called electromagnetic<br />

waves. Start<strong>in</strong>g with the def<strong>in</strong>itions of the electric and magnetic fields given<br />

<strong>in</strong> terms of the vector potential <strong>in</strong> equation (41), one can choose a gauge, i.e.<br />

fix A µ , which will simplify the mathematical expressions as well as the calculations,<br />

we will be deal<strong>in</strong>g with. The reason why we are allowed to make<br />

this choice is that gauge symmetry transforms one solution <strong>in</strong>to another,<br />

both solutions be<strong>in</strong>g physically equivalent 7 . By mak<strong>in</strong>g a gauge choice one<br />

6 A physical quantity that behaves like a scalar, only it changes sign under parity<br />

<strong>in</strong>version e.g. an improper rotation.<br />

7 Both solutions belong the same gauge orbit.<br />

30


eaks the gauge symmetry. This removes the excessive, unphysical degrees<br />

of freedom, which make two physically equivalent solutions to the equations<br />

of motion appear different. Obviously the simplicity of these equations and<br />

their solutions drastically depends on the gauge choice.<br />

One of the convenient gauge choices <strong>in</strong>volves sett<strong>in</strong>g ∂ µ A µ = 0, which is<br />

the covariant gauge choice known as the Lorenz gauge 8 . This however is not<br />

a complete gauge choice, because, as will be shown later, there are still the<br />

gauge transformations that leave the electromagnetic field tensor unchanged.<br />

A further specification of the Lorenz gauge known as the Coulomb gauge sets<br />

the divergence of the vector or the scalar potential equal to zero, i.e. div ⃗ A = 0<br />

and ϕ = 0. We will return back to the comparison of these gauge choices<br />

later.<br />

To see the process of gauge fix<strong>in</strong>g and how we can use it to simplify the<br />

equations of motion, consider the gauge transformations<br />

⃗A → A ⃗ + ∇f ⃗ ,<br />

ϕ → ϕ − 1 ∂f<br />

c ∂t .<br />

If f does not depend on t, ϕ will not change, however ⃗ A will. On the other<br />

hand, div ⃗ A does not depend on t by the Maxwell equations 9 . Thus, <strong>in</strong> this<br />

gauge, equations (41) become<br />

⃗E = − ⃗ ∇ϕ − 1 c<br />

⃗H = rot ⃗ A .<br />

∂A<br />

⃗<br />

∂t = −1 ∂A<br />

⃗<br />

c ∂t ,<br />

Plugg<strong>in</strong>g this <strong>in</strong>to (50), our Maxwell’s equation describ<strong>in</strong>g the curl of the<br />

8 Often erroneously referred to as the Lorentz gauge, due to the similarity with the name<br />

Lorentz as <strong>in</strong> Lorentz transformations, developed by Dutch physicist Hendrik Lorentz.<br />

However it was a Danish physicist, Ludvig Lorenz, who actually <strong>in</strong>troduced the Lorenz<br />

gauge.<br />

9 Under the gauge transformation with the time-<strong>in</strong>dependent function f we have<br />

div ⃗ A → div ⃗ A + div∇f, therefore, the function f should be determ<strong>in</strong>ed from the Poisson<br />

equation ∆f = −div ⃗ A.<br />

31


magnetic field, we obta<strong>in</strong><br />

rot ⃗ H = rot rot ⃗ A = 1 c<br />

⇒<br />

(<br />

∂<br />

∂t<br />

−1<br />

c<br />

∂A<br />

⃗ )<br />

= − 1 ∂ 2 A ⃗<br />

∂t c ∂t , 2<br />

−∆ ⃗ A + grad div ⃗ A = −1<br />

c 2 ∂ 2 ⃗ A<br />

∂t 2 .<br />

In this gauge we can choose f, such that the term <strong>in</strong>volv<strong>in</strong>g the divergence of<br />

⃗A disappears. The equation that rema<strong>in</strong>s is known as d’Alembert’s equation<br />

(or the wave equation)<br />

∆ ⃗ A − 1 c 2 ∂ 2 ⃗ A<br />

∂t 2 = 0 .<br />

When we only consider the plane-wave solutions (i.e. only x-dependence),<br />

then the equation reduces to<br />

∂ 2 f<br />

∂x − 1 ∂ 2 f<br />

2 c 2 ∂t = 0 . 2<br />

It can be further written <strong>in</strong> the factorized form<br />

( ∂<br />

∂t − c ∂ ) ( ∂<br />

∂x ∂t + c ∂ )<br />

f = 0 .<br />

∂x<br />

With a change of variables ξ = t − x and η = t + x ⇒ ∂2 f<br />

c c<br />

solution to the equation is<br />

∂ξ∂η<br />

= 0. Hence, the<br />

f = f (ξ) + f (η) .<br />

Chang<strong>in</strong>g our variables back to x and t, we f<strong>in</strong>d that the general solution for<br />

f is given by<br />

(<br />

f = f 1 t − x ) (<br />

+ f 2 t + x )<br />

.<br />

c<br />

c<br />

Notice that this solution simply represents the sum of right- and left-mov<strong>in</strong>g<br />

plane waves of any arbitrary profile, respectively.<br />

Let us return to the issue of the Coulomb versus the Lorenz gauge choice,<br />

and first consider the later. The Lorentz gauge condition reads as follows<br />

0 = ∂Aµ<br />

∂x µ<br />

= div ⃗ A + 1 c<br />

∂φ<br />

∂t .<br />

32


We see that under gauge transformations the Lorenz gauge condition transforms<br />

as<br />

∂ µ (A µ + ∂ µ χ) = ∂Aµ<br />

∂x µ + ∂ µ∂ µ χ<br />

and it rema<strong>in</strong>s unchanged provided ∂ µ ∂ µ χ = 0. Thus, the Lorenz gauge does<br />

not kill the gauge freedom completely. We still have a possibility to perform<br />

gauge transformations of the special type ∂ µ ∂ µ χ = 0. Hence there will be still<br />

an excessive number of solutions that are physically equivalent and transform<br />

<strong>in</strong>to each other under gauge transformations <strong>in</strong>volv<strong>in</strong>g harmonic functions.<br />

This problem is fixed with the <strong>in</strong>troduction of the complete gauge choice.<br />

Start<strong>in</strong>g over, one can always fix ϕ = 0 by choos<strong>in</strong>g a suitable function<br />

χ (⃗x, t), i.e. a function such that ϕ = 1 ∂χ<br />

. Under the gauge transformations<br />

c ∂t<br />

we have<br />

ϕ → ϕ − 1 ∂χ<br />

⇒ ϕ = 0 .<br />

c ∂t<br />

Transform<strong>in</strong>g the new ϕ = 0 with a new, only space-dependent function<br />

˜χ (x, y, z), we obta<strong>in</strong> 10<br />

S<strong>in</strong>ce ⃗ E = − ⃗ ∇ϕ − 1 c<br />

0 = ϕ → ϕ − 1 c<br />

∂ ⃗ A<br />

∂t<br />

div ⃗ E = − 1 c<br />

∂ ˜χ<br />

∂t = 0 and ⃗ A → ⃗ A + ∇˜χ .<br />

and ϕ = 0, we f<strong>in</strong>d<br />

∂<br />

∂t div ⃗ A and div ⃗ E = 0 ,<br />

where the right hand side has to be equal to zero from our orig<strong>in</strong>al assumption<br />

- lack of sources of electromagnetic fields. From the above equation we can<br />

<strong>in</strong>fer that ∂ ∂t div ⃗ A = 0. We can use yet another gauge freedom to set the<br />

space-dependent and time-<strong>in</strong>dependent ˜χ, such that div ⃗ A = −div ⃗ ∇˜χ, which<br />

means that we have reached the Coulomb gauge<br />

div ⃗ A → div ⃗ A + div ⃗ ∇˜χ = 0 .<br />

Hav<strong>in</strong>g fixed the gauge, let us now consider plane wave solution to the<br />

d’Alambert equation. In this case the derivatives of the y and z component<br />

of the vector potential with respect to y and z components respectively<br />

10 Note that ∂ ˜χ<br />

∂t = 0. 33


direction of propagation<br />

H<br />

E <br />

should vanish as we will only look at oscillations <strong>in</strong> the x direction. This<br />

implies that<br />

div ⃗ A = 0 = ∂A x<br />

∂x + ∂A y<br />

∂y + ∂A z<br />

∂z ⇒ ∂A x<br />

∂x = 0 .<br />

If ∂A x<br />

∂x<br />

form<br />

= 0 everywhere, then ∂2 A x<br />

∂x 2<br />

= 0, which leaves the wave equation <strong>in</strong> the<br />

∂ 2 A x<br />

∂x 2<br />

− 1 c 2 ∂ 2 A x<br />

∂t 2 = 0<br />

− 1 ∂ 2 A x<br />

= 0 ⇒ ∂2 A x<br />

c 2 ∂t 2 ∂t 2<br />

= 0 ⇒ ∂A x<br />

∂t<br />

= const.<br />

S<strong>in</strong>ce we are not <strong>in</strong>terested <strong>in</strong> a constant electric field E x , we need to fix<br />

A x = 0. S<strong>in</strong>ce E ⃗ = − 1 ∂A<br />

⃗ and H ⃗ = rot A, ⃗ then<br />

c ∂t<br />

[<br />

⃗H = ⃗∇(<br />

t− x c )<br />

A]<br />

, ⃗ = − 1 [<br />

⃗n, ∂ ]<br />

A<br />

c ∂t ⃗ = [ ⃗n, E ⃗ ] ,<br />

where [ ]<br />

A, ⃗ B ⃗ denotes the cross-product of two vectors. From the def<strong>in</strong>ition of<br />

the cross product one can see that the electric field E ⃗ and the magnetic field<br />

⃗H are perpendicular to each other. Waves with this property are referred to<br />

as transversal waves.<br />

Electromagnetic waves are known to carry energy; we can def<strong>in</strong>e the energy<br />

flux to be<br />

⃗S = c [ ] ⃗E, H ⃗<br />

c [ [ ]]<br />

= ⃗E, ⃗n, E ⃗ .<br />

4π 4π<br />

34


S<strong>in</strong>ce [ ⃗a, [ ⃗ b,⃗c<br />

]]<br />

= ⃗ b<br />

(<br />

⃗a,⃗c<br />

)<br />

− ⃗c<br />

(<br />

⃗a, ⃗ b<br />

)<br />

, where<br />

(<br />

⃗a, ⃗ b<br />

)<br />

denotes the scalar product<br />

between vectors ⃗a and ⃗ b, we f<strong>in</strong>d the follow<strong>in</strong>g result<br />

⃗S = c<br />

4π ⃗n ⃗ E 2 ,<br />

where due to orthogonality of ⃗n and ⃗ E the contribution of the second term<br />

vanishes. The energy density is given by<br />

W = 1 ( ⃗E 2 + H<br />

8π<br />

⃗ 2) .<br />

For electromagnetic waves ∣ ∣ ⃗ E<br />

∣ ∣ =<br />

∣ ∣ ⃗H<br />

∣ ∣, so that W =<br />

1<br />

4π ⃗ E 2 . Hence, there exists<br />

a simple relationship<br />

⃗S = cW⃗n .<br />

We def<strong>in</strong>e the momentum associated to the electromagnetic wave to be<br />

⃗p = ⃗ S<br />

c 2 = W c ⃗n .<br />

For a particle mov<strong>in</strong>g along ⃗n, we have p = W . Consider a particle<br />

c<br />

mov<strong>in</strong>g with velocity ⃗v. We then have p = vE which for v → c becomes<br />

c 2<br />

p = E ; the dispersion relation for a relativistic particle mov<strong>in</strong>g at the speed<br />

c<br />

of light (photon).<br />

Cont<strong>in</strong>u<strong>in</strong>g, we are now <strong>in</strong>terested <strong>in</strong> the case of fields created by mov<strong>in</strong>g<br />

charges. So far we have discussed<br />

1. Time-<strong>in</strong>dependent fields created by charges at rest<br />

2. Time-dependent fields but without charges<br />

We will now study time-dependent fields <strong>in</strong> the presence of arbitrary mov<strong>in</strong>g<br />

charges 11 . Consider<br />

∂<br />

∂x ν (∂µ A ν − ∂ ν A µ ) =<br />

∂F µν<br />

∂x ν = − 4π c jµ ,<br />

∂ 2<br />

A ν −<br />

∂2<br />

A µ = − 4π ∂x ν ∂x µ ∂x ν ∂x ν c jµ .<br />

11 The motion of the charges has to be strictly def<strong>in</strong>ed, i.e. even though the charges<br />

produce an electromagnetic field, their motion will not be <strong>in</strong>fluenced by the presence of<br />

external electromagnetic fields.<br />

35


Impos<strong>in</strong>g the Lorenz condition<br />

we obta<strong>in</strong> from the previous equation<br />

∂A ν<br />

∂x ν = 0 ,<br />

∂ 2<br />

∂x ν ∂x ν<br />

A µ = 4π c jµ .<br />

The last equation can be split <strong>in</strong>to two<br />

∆ ⃗ A − 1 c 2 ∂ 2 ⃗ A<br />

∂t 2 = − 4π c ⃗ j ,<br />

∆ϕ − 1 c 2 ∂ 2 ϕ<br />

∂t 2 = − 4π c ρ .<br />

These wave equations represent a structure, which is already familiar to us,<br />

namely<br />

∆ψ − 1 ∂ 2 ψ<br />

= −4πf (⃗x, t) . (55)<br />

c 2 ∂t2 To solve this problem, as <strong>in</strong> electrostatics, it is useful to first f<strong>in</strong>d the Green’s<br />

function G (⃗x, t; ⃗x ′ , t ′ ), def<strong>in</strong>ed as a solution of the follow<strong>in</strong>g equation<br />

(∆ x − 1 )<br />

∂ 2<br />

G (⃗x, t; ⃗x ′ , t ′ ) = −4πδ (⃗x − ⃗x ′ ) δ (t − t ′ ) . (56)<br />

c 2 ∂t 2<br />

Note that G (⃗x, t; ⃗x ′ , t ′ ) is not unique and it has to be specified <strong>in</strong> a number<br />

of ways. Additionally, it is referred to as the propagator (especially <strong>in</strong> the<br />

field of quantum electrodynamics). The solution to equation (55) reads<br />

∫<br />

ψ (⃗x, t) = G (⃗x, t; ⃗x ′ , t ′ ) f (⃗x ′ , t ′ ) d 3 x ′ dt .<br />

To check that this is actually the solution, one can apply the operator ∆ x −<br />

1 ∂ 2<br />

and move it <strong>in</strong>to the <strong>in</strong>tegral - two delta functions will emerge by virtue<br />

c 2 ∂t 2<br />

of (56), which upon <strong>in</strong>tegration will turn f (⃗x ′ , t ′ ) <strong>in</strong>to f (⃗x, t). In what follows<br />

we will need the Fourier transforms of all the elements of equation (56)<br />

δ (⃗x − ⃗x ′ ) δ (t − t ′ ) = 1<br />

(2π) 4 ∫ ∞<br />

G (⃗x, t; ⃗x ′ , t ′ ) =<br />

∫ ∞<br />

−∞<br />

d 3 k<br />

−∞<br />

∫ ∞<br />

−∞<br />

d 3 k<br />

36<br />

∫ ∞<br />

dΩ g<br />

−∞<br />

dΩ e i⃗ k·(⃗x−⃗x ′) e −iω(t−t′) ,<br />

(<br />

⃗k, ω<br />

)<br />

e i⃗ k·(⃗x−⃗x ′ )−iω(t−t ′) .


Plugg<strong>in</strong>g these <strong>in</strong>to the equation, we obta<strong>in</strong><br />

which amounts to<br />

g ( ⃗ k, ω<br />

) ( −k 2 + ω2<br />

c 2 )<br />

= −4π 1<br />

(2π) 4 = − 1<br />

4π 3 ,<br />

g ( ⃗ k, ω<br />

)<br />

=<br />

1<br />

4π 3 1<br />

⃗ k2 − ω2<br />

c 2 .<br />

From this one can f<strong>in</strong>d an <strong>in</strong>tegral expression for G (⃗x, t; ⃗x ′ , t ′ )<br />

G (⃗x, t; ⃗x ′ , t ′ ) = 1<br />

4π 3 ∫ ∞<br />

−∞<br />

∫ ∞<br />

d 3 k<br />

−∞<br />

dΩ ei⃗ k·(⃗x−⃗x ′ )−iω(t−t ′ )<br />

.<br />

⃗ k2 − ω2<br />

c 2<br />

The complex function <strong>in</strong>side the <strong>in</strong>tegral is s<strong>in</strong>gular at ⃗ k 2<br />

= ω2<br />

c 2<br />

and thus<br />

has two first order poles at ω = ±c ∣ ∣⃗ k<br />

∣ ∣. We have to f<strong>in</strong>d the proper way to<br />

treat this s<strong>in</strong>gularity. This is done by us<strong>in</strong>g the follow<strong>in</strong>g physical reason<strong>in</strong>g.<br />

The Green function is a wave-type perturbation produced by a po<strong>in</strong>t source<br />

sitt<strong>in</strong>g at x ′ and emanat<strong>in</strong>g dur<strong>in</strong>g an <strong>in</strong>f<strong>in</strong>itesimal time at t = t ′ . We can<br />

expect that this wave propagates with the speed of light as a spherical wave.<br />

Thus, we should require that<br />

a) G = 0 <strong>in</strong> the whole space for t < t ′<br />

b) G is a diverg<strong>in</strong>g wave for t > t ′<br />

We shall see that the above only represents one of the possible Green’s<br />

functions, s<strong>in</strong>ce a different treatment of the poles produces different Green’s<br />

functions - an advanced or a retarded one:<br />

Retarded Green function states G = 0 if t < t ′<br />

Advanced Green function states G = 0 if t > t ′<br />

Notice that the difference of the two G adv − G ret , called the Pauli Green’s<br />

function G P auli , satisfies the homogenous equation .<br />

Consider the retarded Green’s function. For t > t ′ , it should give a wave<br />

propagat<strong>in</strong>g from a po<strong>in</strong>t-like source. Let us def<strong>in</strong>e τ = t − t ′ , ⃗ R = ⃗x − ⃗x ′<br />

and R = ∣ ∣ ⃗ R<br />

∣ ∣. Then we have<br />

e −iω(t−t′) ∼ e Iωτ ,<br />

37


s<strong>in</strong>ce τ > 0. Thus we need to require that Iω < 0 <strong>in</strong> order to have a decay<strong>in</strong>g<br />

function at large ω, hence we have to <strong>in</strong>tegrate over the lower complex plane.<br />

In other words, for t < t ′ , the contour over which we <strong>in</strong>tegrate <strong>in</strong> the upper<br />

half of the complex plane should give zero contribution due to the aforementioned<br />

physical reasons. As a result, one could <strong>in</strong>f<strong>in</strong>itesimally shift the<br />

poles <strong>in</strong>to the lower half plane when perform<strong>in</strong>g the analytic cont<strong>in</strong>uation.<br />

Accord<strong>in</strong>g to this prescription, the Green’s function is specified as follows<br />

G(⃗x, t; ⃗x ′ , t ′ ) = 1 ∫ ∫<br />

d 3 e i⃗ kR−iωτ<br />

k dω<br />

4π 3 k 2 − 1 (ω + iε) .<br />

c 2 2<br />

We can conveniently rewrite the previous statement, by mak<strong>in</strong>g use of partial<br />

fractions<br />

G (⃗x, t; ⃗x ′ , t ′ ) = (57)<br />

= 1 ∫ ∞ ∫ ∞<br />

d 3 k dωe i⃗ kR c [<br />

]<br />

1<br />

4π 3 2k ck − iε − ω − 1<br />

e −iωτ .<br />

−ck − iε − ω<br />

−∞<br />

−∞<br />

In the limit ε → 0, us<strong>in</strong>g Cauchy’s theorem, we f<strong>in</strong>d<br />

G (⃗x, t; ⃗x ′ , t ′ ) = 1<br />

4π 3 ∫ ∞<br />

−∞<br />

= c<br />

2π 2 ∫ ∞<br />

= 2c<br />

πR<br />

= 1<br />

πR<br />

= − 1<br />

=<br />

−∞<br />

∫ ∞<br />

0<br />

∫ ∞<br />

4πR<br />

1<br />

2πR<br />

−∞<br />

∫ ∞<br />

−∞<br />

∫ ∞<br />

d 3 ke i⃗ k· ⃗R 2πi c [<br />

e −ickτ − e ickτ] (58)<br />

2k<br />

d 3 k ei⃗ k· ⃗R<br />

k<br />

s<strong>in</strong> ckτ<br />

dk s<strong>in</strong>(kR) s<strong>in</strong>(ckτ) (59)<br />

( ) (ck) R<br />

d (ck) s<strong>in</strong> s<strong>in</strong> ((ck) τ) (60)<br />

c<br />

(<br />

)<br />

dx e ix R c − e<br />

−ix (e R<br />

c ixτ − e −ixτ) (61)<br />

−∞<br />

= 1 (<br />

R δ τ − R c<br />

= 1 (<br />

R δ τ − R c<br />

dx<br />

(e ix (τ− R c ) − e<br />

ix(τ+ R c ) )<br />

)<br />

− 1 (<br />

R δ τ + R )<br />

c<br />

)<br />

(62)<br />

(63)<br />

Note that <strong>in</strong> the meantime we have used: partial fractions (57), the Cauchy<br />

theorem <strong>in</strong> (57-58), switched to spherical coord<strong>in</strong>ates and <strong>in</strong>tegrated over the<br />

38


angles(59), substituted ck = x (60), expanded the trigonometric functions <strong>in</strong><br />

terms of their complex exponentials (61), and identified Fourier transforms<br />

of delta funtions (62). On the last step we have rejected δ ( )<br />

τ + R c , because<br />

for τ, R, c > 0, the result will always be zero. Substitut<strong>in</strong>g back our orig<strong>in</strong>al<br />

variables, we get<br />

(<br />

)<br />

δ t ′ + |⃗x−⃗x′ |<br />

− t<br />

G ret (⃗x, t; ⃗x ′ , t ′ c<br />

) =<br />

.<br />

|⃗x − ⃗x ′ |<br />

The result can be understood as the signal propagat<strong>in</strong>g at the speed of light,<br />

which was emitted at t ′ and will travel for |⃗x−⃗x′ |<br />

and will be observed at time<br />

c<br />

t. Thus, this Green function reflects a natural causal sequence of events. The<br />

time t is then expressed <strong>in</strong> terms of the retarded time t ′<br />

t = t ′ + |⃗x − ⃗x′ |<br />

c<br />

Substitut<strong>in</strong>g this solution and <strong>in</strong>tegrat<strong>in</strong>g over t ′ , we obta<strong>in</strong> the “retarded”<br />

potentials<br />

)<br />

∫ δ<br />

(t ′ + |⃗x−⃗x′ |<br />

− t<br />

c<br />

ϕ (⃗x, t) =<br />

ρ (⃗x ′ , t ′ ) d 3 x ′ dt<br />

|⃗x − ⃗x ′ |<br />

)<br />

∫ ρ<br />

(⃗x ′ , t − |⃗x−⃗x′ |<br />

c<br />

=<br />

d 3 x ′ + ϕ<br />

|⃗x − ⃗x ′ 0 , (64)<br />

|<br />

⃗A (⃗x, t) = 1 c<br />

(<br />

= 1 ∫ ⃗j<br />

c<br />

∫ δ<br />

(t ′ + |⃗x−⃗x′ |<br />

c<br />

|⃗x − ⃗x ′ |<br />

⃗x ′ , t − |⃗x−⃗x′ |<br />

c<br />

|⃗x − ⃗x ′ |<br />

.<br />

)<br />

− t<br />

⃗j (⃗x ′ , t ′ ) d 3 x ′ dt<br />

)<br />

d 3 x ′ + ⃗ A 0 , (65)<br />

where ϕ 0 and A ⃗ 0 are the solutions of the homogeneous d’Alembert equations<br />

(those correspond<strong>in</strong>g to the free electromagnetic field).<br />

Note that for ϕ <strong>in</strong> the case of time-<strong>in</strong>dependent ρ and j we have<br />

∫<br />

ϕ =<br />

ρ(⃗x ′ )<br />

|⃗x − ⃗x ′ | d3 x ′ .<br />

39


This is just the electrostatic formula for the scalar potential. Moreover, if<br />

the current j is time-<strong>in</strong>dependent, we obta<strong>in</strong><br />

⃗A(x) = 1 c<br />

∫<br />

⃗j(⃗x ′ )<br />

|⃗x − ⃗x ′ | d3 x ′ .<br />

This potential def<strong>in</strong>es the follow<strong>in</strong>g magnetic field<br />

∫ [ ]<br />

⃗H = rot xA ⃗<br />

1 rot x<br />

⃗j(⃗x ′ ) 1<br />

= + ∇<br />

c |⃗x − ⃗x ′ x<br />

| |⃗x − ⃗x ′ | × ⃗j(⃗x ′ )<br />

Note the use above of the follow<strong>in</strong>g identity<br />

rot(ϕ⃗a) = ϕ rot⃗a + ⃗ ∇ϕ × ⃗a .<br />

d 3 x ′ . (66)<br />

The first term <strong>in</strong> (66) vanishes, because curl is taken with respect to coord<strong>in</strong>ates<br />

x, while the current ⃗j depends on x ′ . This leaves<br />

⃗H = − 1 ∫ ⃗R × ⃗j(⃗x ′ )<br />

d 3 x ′ = 1 ∫ [ ]<br />

⃗j(⃗x ′ ), ⃗x − ⃗x ′ d 3 x ′ .<br />

c R 3 c |⃗x − ⃗x ′ | 3<br />

This is the famous law of Biot-Savart, which relates magnetic fields to their<br />

source currents.<br />

Let us now show that G ret is Lorentz <strong>in</strong>variant. We write<br />

(<br />

)<br />

δ t ′ + |⃗x−⃗x′ |<br />

− t<br />

G ret (⃗x, t; ⃗x ′ , t ′ ) = Θ (t − t ′ c<br />

)<br />

.<br />

|⃗x − ⃗x ′ |<br />

Here the extra term Θ (t − t ′ ) ensures that G ret (⃗x, t; ⃗x ′ , t ′ ) = 0 for t < t ′ ,<br />

because<br />

{ 0, t < t<br />

Θ (t − t ′ ′<br />

) =<br />

1, t ′ ≥ t<br />

When we use<br />

δ (f (x)) = ∑ i<br />

δ (x)<br />

|f ′ (x o )| .<br />

In the last formula the derivative is evaluated at the set of po<strong>in</strong>ts x o , such<br />

that f ( o ) = 0. Realis<strong>in</strong>g that for a wave propagat<strong>in</strong>g at the speed of light<br />

40


ds 2 = 0 and us<strong>in</strong>g some algebraic trickery, we get<br />

G ret (⃗x, t; ⃗x ′ , t ′ ) = 2cΘ (t − t ′ ) δ (|⃗x − ⃗x′ | − c (t − t ′ ))<br />

2 |⃗x − ⃗x ′ |<br />

= 2cΘ (t − t ′ ) δ (|⃗x − ⃗x′ | − c (t − t ′ ))<br />

|⃗x − ⃗x ′ | + c (t − t ′ )<br />

(<br />

= 2cΘ (t − t ′ ) δ |⃗x − ⃗x ′ | 2 − c 2 (t − t ′ ) 2) ,<br />

where the argument of the delta function is the 4-<strong>in</strong>terval between two events<br />

(⃗x, t) and (⃗x ′ , t ′ ), which is a Lorentz <strong>in</strong>variant object. From this we can<br />

conclude that the Green’s function is <strong>in</strong>variant under proper orthochronical<br />

(ones that ma<strong>in</strong>ta<strong>in</strong> causality) Lorentz transformations.<br />

2.5 Causality Pr<strong>in</strong>ciple<br />

A quick word on <strong>in</strong>tervals. A spacetime <strong>in</strong>terval we have already def<strong>in</strong>ed as<br />

ds 2 = c 2 dt 2 − dx 2 i (67)<br />

We refer to them differently depend<strong>in</strong>g on the sign of ds 2 :<br />

time-like <strong>in</strong>tervals if ds 2 > 0<br />

space-like <strong>in</strong>tervals if ds 2 < 0<br />

light-like <strong>in</strong>tervals (also called null <strong>in</strong>tervals) if ds 2 = 0<br />

Consider Figure 1.9 represent<strong>in</strong>g the light-cone built over a po<strong>in</strong>t X.<br />

Signals <strong>in</strong> X can come only from po<strong>in</strong>ts X ′ , which are <strong>in</strong> the pastlight-cone<br />

of X. We say X > X ′ (X is later than X ′ ). The <strong>in</strong>fluence of a current j<br />

<strong>in</strong> X ′ on potential A at X is a signal from X ′ to X. Thus, the causality<br />

pr<strong>in</strong>ciple is reflected <strong>in</strong> the fact that A(X) can depend on 4-currents j(X ′ )<br />

only for those X ′ for which X > X ′ . Thus,<br />

δA(X)<br />

δj(X ′ ) ∼ G(X − X′ ) = 0 (68)<br />

for X < X ′ or po<strong>in</strong>ts X ′ that are space-like to X.<br />

pr<strong>in</strong>ciple for the Green function is<br />

Hence, the causality<br />

G(X ′ − X) = 0 , (69)<br />

<strong>in</strong> terms of the conditions described above. The retarded Green’s function is<br />

the only relativistic Green’s function which has this property.<br />

41


❏<br />

❏<br />

❏<br />

❏<br />

❏<br />

✡<br />

✡<br />

✡<br />

✡<br />

absolute<br />

future<br />

light-like<br />

✡ ✡✡✡<br />

❏✡ ✡✡ X space-like<br />

✡❏ ✡ ❏❏❏❏❏❏<br />

past<br />

X ′<br />

time-like<br />

Figure 10: At every po<strong>in</strong>t <strong>in</strong> time every observer has his past light cone, which is<br />

a set of all events that could have <strong>in</strong>fluenced his presence, and a future light cone,<br />

the set of events which the observer can <strong>in</strong>fluence. The boundaries of the light<br />

cones also def<strong>in</strong>e the split between different k<strong>in</strong>ds of space-time <strong>in</strong>tervals. On the<br />

light cone itself the <strong>in</strong>tervals are all light-like, time-like on the <strong>in</strong>side and space-like<br />

on the outside.<br />

2.6 Applicability of <strong>Classical</strong> <strong>Electrodynamics</strong><br />

We conclude this section by po<strong>in</strong>t<strong>in</strong>g out the range of applicability of classical<br />

electrodynamics.<br />

The energy of the charge distribution <strong>in</strong> electrodynamics is given by<br />

U = 1 ∫<br />

dV ρ(x)ϕ(x) .<br />

2<br />

Putt<strong>in</strong>g electron at rest, one can assume that the entire energy of the electron<br />

co<strong>in</strong>cides with its electromagnetic energy (electric charge is assumed to be<br />

homogeneously distributed over a ball of the radius r e )<br />

mc 2 ∼ e2<br />

r e<br />

,<br />

where m and e are the mass and the charge of electron. Thus, we can def<strong>in</strong>e<br />

the classical radius of electron<br />

r e =<br />

e2<br />

mc 2 ∼ 2.818 · 10−15 m .<br />

42


In SI units it reads as r e = 1<br />

electrodynamics is not applicable.<br />

e 2<br />

4πɛ 0 mc 2<br />

. At distances less than r e , the classical<br />

In reality, due to quantum effects the classical electrodynamics fails even<br />

at larger distances. The characteristic scale is the Compton wavelength, which<br />

is the fundamental limitation on measur<strong>in</strong>g the position of a particle tak<strong>in</strong>g<br />

both quantum mechanics and special relativity <strong>in</strong>to account. Its theoretical<br />

value is given by<br />

<br />

mc ∼ 137 r e ∼ 10 −13 m ,<br />

where α = 1 = e2 is the f<strong>in</strong>e structure constant for electromagnetism. The<br />

137 c<br />

most recent experimental measurement of campton wavelenght (CODATA<br />

2002) is one order of magnitude larger and is approximately equal to 2.426 ·<br />

10 −12 m<br />

3 Radiation<br />

The last part of these lectures will treat two classical radiation problems:<br />

Liénard-Wiechert potentials and the dipole radiation.<br />

3.1 Liénard-Wiechert Potentials<br />

The charge distribution <strong>in</strong> space and time of a s<strong>in</strong>gle charge is be given by<br />

ρ (⃗x, t) = eδ (⃗x − ⃗r (t)) ,<br />

⃗j (⃗x, t) = e⃗vδ (⃗x − ⃗r (t)) .<br />

Here ⃗x is the position of the observer, ⃗r (t) is the trajectory of the charge<br />

and ⃗v = ṙ (t), its velocity. The potential then reads<br />

)<br />

∫ δ<br />

(t ′ + |⃗x−⃗x′ |<br />

− t<br />

c<br />

ϕ (⃗x, t) =<br />

eδ (⃗x ′ − ⃗r (t ′ )) d 3 x ′ dt ′ (70)<br />

|⃗x − ⃗x ′ |<br />

Let us take ⃗x ′ = ⃗r (t ′ ), because only then the <strong>in</strong>tegrand is non-zero. Then<br />

eq.(70) can be <strong>in</strong>tegrated over ⃗x ′ and we get<br />

(<br />

)<br />

∫ δ t ′ + |⃗x−⃗r(t′ )|<br />

− t<br />

c<br />

ϕ (⃗x, t) = e<br />

dt ′ . (71)<br />

|⃗x − ⃗r (t ′ )|<br />

43


Take f (t ′ ) = t ′ + |⃗x−⃗r(t′ )|<br />

−t and use δ (f (x)) = δ(x) , where f ′ (x) is evaluated<br />

c<br />

|f ′ (x)|<br />

at the po<strong>in</strong>t were f (x) = 0, i.e. at t ′ which solves t ′ + |⃗x−⃗r(t′ )|<br />

− t = 0<br />

c<br />

df (t ′ )<br />

dt ′<br />

= 1 − 1 c<br />

(⃗x − ⃗r(t ′ )) · ˙⃗r(t ′ )<br />

|⃗x − ⃗r (t ′ )|<br />

= 1 − 1 c<br />

⃗R · ⃗v<br />

R .<br />

In the last equation we have used the fact that R ⃗ = ⃗x − ⃗r (t ′ ) and ⃗v = ṙ (t).<br />

The potential then becomes<br />

1 e<br />

ϕ (⃗x, t) = e =<br />

R 1 − 1 ⃗R·⃗v<br />

c R<br />

We can use the same l<strong>in</strong>e of reason<strong>in</strong>g to show<br />

R − ⃗ R·⃗v<br />

c<br />

. (72)<br />

⃗A (⃗x, t) = e ⃗v<br />

c R·⃗v<br />

(R − ⃗ c (73)<br />

The formulae (72) and (73) are the Liénard-Wiechert potentials. Let us<br />

compute the correspond<strong>in</strong>g electric and magnetic fields. We have<br />

⃗E = − 1 ∂A<br />

c ∂t − ∇ϕ ⃗ ;<br />

⃗H = rotA ⃗ .<br />

Moreover, R(t ′ ) is given by the difference <strong>in</strong> the times t and t ′ with an overall<br />

factor of c<br />

R (t ′ ) = c (t − t ′ ) .<br />

Therefore,<br />

∂R (t ′ )<br />

∂t<br />

From this relation, it follows that<br />

= ∂R (t′ ) ∂t ′<br />

∂t ′ ∂t = − R ⃗ ( )<br />

· ⃗v ∂t ′<br />

R ∂t = c 1 − ∂t′ . (74)<br />

∂t<br />

∂t ′<br />

∂t = 1<br />

.<br />

R·⃗v<br />

1 − ⃗ Rc<br />

Analogously, one can also start from the expressions R(t ′ ) = c(t − t ′ ) and<br />

t ′ = t ′ (t, ⃗x), such that<br />

⃗∇R (t ′ ) = −c∇t ⃗ ′ ⇒ ∇t ⃗ ′ = − 1 ∇R<br />

c ⃗ (t ′ ) = − 1 ∇<br />

c ⃗ x |⃗x − ⃗r (t ′ (⃗x, t))|<br />

( )<br />

= − 1 ⃗R<br />

c R + ∂R ∇<br />

∂t ⃗ (t ′ ) ,<br />

′<br />

44


where one can aga<strong>in</strong> identify ∂R with the previous result from (74) and f<strong>in</strong>ally<br />

∂t ′<br />

obta<strong>in</strong><br />

⃗R<br />

⃗∇t ′ = − ( ) .<br />

R·⃗v<br />

c R − ⃗ c 2<br />

Now we have all the necessary <strong>in</strong>gredients, which we can use to f<strong>in</strong>d ⃗ E and<br />

⃗H, i.e. to obta<strong>in</strong> the Liénard-Wiechert fields<br />

⃗H = 1 [ ]<br />

⃗R, E ⃗ ,<br />

R(<br />

) ( )<br />

1 − v2 ⃗R<br />

c −<br />

⃗v<br />

⃗E R = e<br />

2 c<br />

( )<br />

R·⃗v<br />

R − ⃗ 3<br />

+<br />

c<br />

e<br />

[<br />

⃗R,<br />

[<br />

⃗R −<br />

⃗v<br />

c<br />

c 2 (R − ⃗ R·⃗v<br />

c<br />

]]<br />

R, ˙⃗v<br />

) 3<br />

.<br />

Notice that <strong>in</strong> the last equation the first term only depends on the velocity<br />

of the mov<strong>in</strong>g particle and is proportional to 1 (short distance), whereas<br />

R 2<br />

the second term depends on acceleration and is proportional to 1 provid<strong>in</strong>g,<br />

therefore, the long-distance dom<strong>in</strong>at<strong>in</strong>g contribution, the so-called wave-<br />

R<br />

zone. Note also that flux is proportional to E ⃗ 2 hence is also proportional to<br />

1<br />

. Therefore,<br />

R 2 ∫ ∫ 1<br />

⃗E 2 dV =<br />

R 2 R2 dΩ = 4π ,<br />

which is a constant flux of ⃗ E at large distances. It is worth stress<strong>in</strong>g that<br />

there is no energy (radiation) com<strong>in</strong>g from a charge mov<strong>in</strong>g at a constant<br />

velocity, because we can always choose a frame where it is stationary, hence<br />

⃗H = 0 ⇒ ⃗ E · ⃗H = 0, consequently it cannot emit energy.<br />

3.2 Dipole Radiation<br />

Field of a neutral system is expressed with the help of the so-called electric<br />

moment given <strong>in</strong> its discretised form as<br />

⃗p =<br />

N∑<br />

e iRi ⃗ , (75)<br />

i=1<br />

where e i is the magnitude of a charge at some distance R i taken from an<br />

arbitrary po<strong>in</strong>t, <strong>in</strong> this case chosen to be the orig<strong>in</strong>. For a neutral system we<br />

45


V<br />

0<br />

l<br />

R r<br />

0<br />

P( x, y,<br />

z)<br />

R r<br />

x r '<br />

( x', y', z'<br />

)<br />

Figure 11: A diagramatic representation of a dipole<br />

require that<br />

N∑<br />

e i = 0 .<br />

i=1<br />

Note that for such a system, electric moment does not depend on the choice<br />

of the orig<strong>in</strong> of the reference frame, i.e. shift<strong>in</strong>g all ⃗ R i → ⃗ R i − ⃗a gives<br />

⃗p ⃗a =<br />

N∑<br />

i=1<br />

( )<br />

e ⃗Ri i − ⃗a =<br />

N∑<br />

e iRi ⃗ − ⃗a<br />

i=1<br />

N∑<br />

e i =<br />

i=1<br />

N∑<br />

e iRi ⃗ = ⃗p .<br />

i=1<br />

Let us now consider a neutral system of mov<strong>in</strong>g charges. From diagram 11<br />

us<strong>in</strong>g Pythagorean theorem and assum<strong>in</strong>g that ⃗ l ≪ R 0 , l be<strong>in</strong>g the charac-<br />

46


teristic size, we get<br />

√ (<br />

R = ⃗R0 − R ⃗ ) 2<br />

√<br />

′ = ⃗R 0 2 − 2R ⃗ 0 · ⃗R ′ + R ⃗ ′2 ≈<br />

(<br />

≈ √ R ⃗ 2<br />

R<br />

0 1 − 2 ⃗ 0 · ⃗R<br />

) (<br />

′<br />

R<br />

≈ R 0 1 − ⃗ 0 · ⃗R<br />

)<br />

′ R<br />

≈ R 0 − ⃗ 0 · ⃗R ′<br />

.<br />

⃗R 0<br />

2 ⃗R 0<br />

2 R 0<br />

By us<strong>in</strong>g (64), we then f<strong>in</strong>d the retarded scalar potential<br />

∫ ( )<br />

ρ x ′ , t − R c<br />

ϕ =<br />

d 3 x ′ =<br />

R<br />

∫<br />

= d 3 x ′ ρ ( )<br />

x ′ , t − R 0<br />

c<br />

R<br />

− ⃗ 0 · ⃗R ′ ∂ ρ ( )<br />

x ′ , t − R 0<br />

c<br />

· · · =<br />

R 0 ⃗R 0<br />

∂R 0 R 0<br />

R<br />

= ⃗ ∫ (<br />

0 ∂ 1<br />

· d 3 x ′ R ⃗ ′ ρ x ′ , t − R )<br />

0<br />

,<br />

R 0 ∂R 0 R 0 c<br />

where the first term vanishes because it is proportional the complete charge<br />

of the system, which we have set to zero, by def<strong>in</strong><strong>in</strong>g the system to be neutral.<br />

In the rema<strong>in</strong><strong>in</strong>g term we will write the <strong>in</strong>tegral as ⃗p ( )<br />

t − R 0<br />

c<br />

, the electric<br />

moment at time t − R 0<br />

Therefore,<br />

Further, we f<strong>in</strong>d<br />

so that<br />

div ⃗p ( x ′ , t − R c<br />

R<br />

c<br />

, which is just a cont<strong>in</strong>uous version of (75)<br />

(<br />

⃗p t − R ) ∫ (<br />

0<br />

= d 3 x ′ R ⃗ ′ ρ x ′ , t − R 0<br />

c<br />

c<br />

ϕ = ⃗ R<br />

R ·<br />

)<br />

∂ 1<br />

(x<br />

∂R R ⃗p ′ , t − R )<br />

.<br />

c<br />

)<br />

. (76)<br />

= ⃗p · ⃗∇ 1 R + 1 · ⃗R<br />

div⃗p = −⃗p + 1 R R 3 R div⃗p ,<br />

∂x = R ⃗ ∂⃗p<br />

i R ∂R ,<br />

div⃗p = ∂p i<br />

∂x = ∂p i ∂R<br />

i ∂R<br />

div ⃗p ( x ′ , t − R c<br />

R<br />

)<br />

⃗p · ⃗R R<br />

= − + ⃗ ∂⃗p<br />

R 3 R 2 ∂R .<br />

47


On the other hand,<br />

Thus,<br />

ϕ =<br />

⃗p · ⃗R<br />

R 3<br />

− ⃗ R<br />

R 2 ∂⃗p<br />

∂R .<br />

ϕ = −div ⃗p ( x ′ , t − R c<br />

.<br />

R<br />

Here divergence is taken over coord<strong>in</strong>ates of the po<strong>in</strong>t P (x, y, z) where the<br />

observer is located. Us<strong>in</strong>g expression (65), the vector potential becomes<br />

⃗A = 1 c<br />

∫<br />

=<br />

∫ ⃗ j ( x ′ , t − R c<br />

)<br />

R<br />

d 3 x ′ =<br />

d 3 x ′ [⃗j ( x ′ , t − R 0<br />

c<br />

)<br />

R 0<br />

− ⃗ R 0 · ⃗R ′<br />

⃗R 0<br />

)<br />

∂ ⃗j ( x ′ , t − R 0 ]<br />

· · · .<br />

∂R 0 R 0<br />

First <strong>in</strong>tegral can also be expressed via electric moment, which can be achieved<br />

by us<strong>in</strong>g the cont<strong>in</strong>uity equation<br />

(<br />

∂<br />

∂t ρ x ′ , t − R ) (<br />

0<br />

= −div ′ ⃗j x ′ , t − R )<br />

0<br />

.<br />

c<br />

c<br />

Multiply<strong>in</strong>g both sides of this equation by time <strong>in</strong>dependent R ⃗ ′ , <strong>in</strong>tegrat<strong>in</strong>g<br />

over entire space and us<strong>in</strong>g the def<strong>in</strong>ition (76), we can then state that<br />

(<br />

∂<br />

∂t ⃗p x ′ , t − R ) ∫<br />

(<br />

0<br />

= − d 3 x ′ R ⃗ ′ div ′ ⃗j x ′ , t − R )<br />

0<br />

.<br />

c<br />

c<br />

To proceed, let us sidetrack and consider an arbitrary unit vector ⃗a, i.e.<br />

|⃗a| = 1. Then<br />

(<br />

⃗a ⃗ R<br />

′ ) div⃗j = div<br />

(<br />

⃗ j ( ⃗a ⃗ R ′)) − ⃗j · ⃗∇ ′( ⃗a ⃗ R ′)<br />

c<br />

)<br />

= div<br />

(<br />

⃗ j ( ⃗a ⃗ R ′)) − ⃗j · ⃗a ,<br />

where the last step follows from ⃗a be<strong>in</strong>g a constant and ∇ ′ R ⃗ ′ = 1. Based on<br />

that we can write<br />

⃗a · ∂ (<br />

∂t ⃗p x ′ , t − R ) ∫ (<br />

0<br />

= − d 3 x ′ div ′ ⃗ j ( ⃗a R<br />

c<br />

⃗ ′)) ∫ (<br />

+ ⃗a · d 3 x ′ ⃗j x ′ , t − R )<br />

0<br />

.<br />

c<br />

48


S<strong>in</strong>ce currents do not leave the volume V , we f<strong>in</strong>d that<br />

∫ [<br />

d 3 x ′ div ′ ⃗ j<br />

(⃗a R ⃗ )] ∮<br />

′ ∼ (aR ′ ) j n dS = 0<br />

as the normal component j n of the current vanishes. This gives<br />

⃗a · ∂ (<br />

∂t ⃗p x ′ , t − R ) ∫ (<br />

0<br />

= ⃗a · d 3 x ′ ⃗j x ′ , t − R )<br />

0<br />

.<br />

c<br />

c<br />

S<strong>in</strong>ce the last relation is valid for any unit vector ⃗a we obta<strong>in</strong> that<br />

(<br />

∂<br />

∂t ⃗p x ′ , t − R ) ∫ (<br />

0<br />

= d 3 x ′ ⃗j x ′ , t − R )<br />

0<br />

.<br />

c<br />

c<br />

F<strong>in</strong>ally, we arrive at<br />

⃗A = 1 Rc · ∂ (<br />

∂t ⃗p x ′ , t − R )<br />

0<br />

.<br />

c<br />

We see that both the scalar and the vector potential of any arbitrary neutral<br />

system on large distances are def<strong>in</strong>ed via the electric moment of this system.<br />

The simplest system of this type is a dipole i.e. two opposite electric<br />

charges separated by a certa<strong>in</strong> distance from each other. A dipole whose<br />

moment ⃗p changes <strong>in</strong> time is called an oscillator (or a vibrator).<br />

Radiation of an oscillator plays an important role <strong>in</strong> the electromagnetic<br />

theory (radiotelegraphic antennae, radiation bodies, proton-electron systems,<br />

etc.). To advance our <strong>in</strong>vestigation of a dipole, let us <strong>in</strong>troduce the Hertz<br />

vector<br />

⃗P (t, R) = ⃗p ( )<br />

t − R c<br />

. (77)<br />

R<br />

It is <strong>in</strong>terest<strong>in</strong>g to see that<br />

∆ ⃗ P (t, R) = ⃗ ∇ 2 ⃗ P (t, R) =<br />

1<br />

c 2 ∂ 2 ⃗ P<br />

∂t 2 .<br />

This can be derived as follows. First, we notice that<br />

∂<br />

P<br />

∂x ⃗ = − 1 ∂R<br />

R 2 ∂x ⃗p − 1<br />

cR<br />

∂⃗p ∂R<br />

∂t ∂x = − x R ⃗p − 3<br />

49<br />

x ∂⃗p<br />

cR 2 ∂t ,


s<strong>in</strong>ce ∂R = x . Differentiat<strong>in</strong>g once aga<strong>in</strong>, we get<br />

∂x R<br />

so that<br />

∂ 2<br />

P<br />

∂x ⃗ = − 1 x2<br />

⃗p + 3 2 R3 R ⃗p + 3 x 2 ∂⃗p<br />

5 c R 4 ∂t − 1 ∂⃗p<br />

cR 2 ∂t + 1 x 2 ∂ 2 ⃗p<br />

c 2 R 3 ∂t , 2<br />

3∑<br />

i=1<br />

∂ 2<br />

∂x 2 i<br />

⃗P = 1 ∂ 2 ⃗p<br />

c 2 R ∂t , 2<br />

which represents the spherically symmetric solution of the wave equation.<br />

Consider the retarded potentials<br />

ϕ (t) = −divP ⃗ (t, R) A ⃗<br />

1 ∂P ⃗ (t, R)<br />

(t) ; = ;<br />

c ∂t<br />

⃗H = rotA ⃗ (t) = rot 1 ∂P ⃗ (t, R)<br />

= 1 ∂<br />

c ∂t c ∂t rot P ⃗ (t, R) ;<br />

⃗E = − 1 c<br />

∂ ⃗ A (t)<br />

∂t<br />

− ⃗ ∇φ = − 1 c 2 ∂ 2 ⃗ P (t, R)<br />

∂t 2 − ⃗ ∇div ⃗ P (t, R)<br />

= − 1 c 2 ∂ 2 ⃗ P (t, R)<br />

∂t 2 + ⃗ ∇ 2 ⃗ P (t, R) + rot rot ⃗ P (t, R) .<br />

On the last l<strong>in</strong>e the sum of the first two terms is equal to zero by virtue of<br />

the wave equation. This results <strong>in</strong><br />

⃗E = rot rot ⃗ P (t, R) . (78)<br />

Assume that the electric moment changes only its magnitude, but not its<br />

direction i.e.,<br />

⃗p (t) = ⃗p 0 f (t) .<br />

This is not a restriction because moment ⃗p of an arbitrary oscillator can be<br />

decomposed <strong>in</strong>to three mutually orthogonal directions and a field <strong>in</strong> each<br />

direction can be studied separately. Based on this we have<br />

f ( )<br />

t −<br />

⃗P R c<br />

(t, R) = ⃗p 0<br />

R<br />

rot ⃗ P = f R rot ⃗p 0 +<br />

,<br />

[<br />

⃗∇ f R , ⃗p 0<br />

]<br />

50<br />

= ∂<br />

∂R<br />

( (<br />

f t −<br />

r<br />

c<br />

R<br />

))<br />

− 1 [ ]<br />

⃗R, ⃗p0 .<br />

R


In the spherical coord<strong>in</strong>ate system we compute the correspond<strong>in</strong>g components<br />

]∣<br />

∣<br />

∣[<br />

⃗R, ∣∣<br />

⃗p0 = Rp0 s<strong>in</strong> θ ,<br />

[ ] [ ]<br />

⃗R, ⃗p0 = ⃗R, ⃗p0 = 0 ,<br />

R<br />

θ<br />

[ ]<br />

⃗R, ⃗p0 = −Rp 0 s<strong>in</strong> θ .<br />

φ<br />

and get<br />

(<br />

rot P ⃗ )<br />

R<br />

(<br />

rot P ⃗ )<br />

φ<br />

=<br />

(<br />

rot P ⃗ )<br />

θ<br />

= 0 ,<br />

= −p 0 s<strong>in</strong> θ ∂<br />

∂R<br />

( (<br />

f t −<br />

r<br />

c<br />

R<br />

))<br />

= − s<strong>in</strong> θ ∂<br />

∂R ⃗ P Hertz (t, R) .<br />

S<strong>in</strong>ce the magnetic field components are the components of the curl of the<br />

vector potential, the latter is written <strong>in</strong> terms of the Hertz vector (77), where<br />

we f<strong>in</strong>d<br />

H R = H θ = 0<br />

H φ = − s<strong>in</strong> θ 1 c<br />

∂ 2 ⃗ PHertz (t, R)<br />

∂t ∂R<br />

The components of curl of any vector field ⃗a <strong>in</strong> spherical coord<strong>in</strong>ates are<br />

given by<br />

(<br />

1 ∂<br />

(rot ⃗a) R<br />

=<br />

R s<strong>in</strong> θ ∂θ (s<strong>in</strong> θa φ) − ∂a )<br />

θ<br />

;<br />

∂R<br />

(<br />

1 ∂aR<br />

(rot ⃗a) θ<br />

=<br />

R s<strong>in</strong> θ ∂φ − ∂<br />

)<br />

∂R (R s<strong>in</strong> θa φ) ;<br />

(rot ⃗a) φ<br />

= 1 ( ∂<br />

R ∂R (Ra θ) − ∂a )<br />

R<br />

.<br />

∂θ<br />

Us<strong>in</strong>g these formulae together with equation (78), we also f<strong>in</strong>d the compo-<br />

.<br />

51


nents of the electric field<br />

[<br />

1 ∂<br />

E R =<br />

s<strong>in</strong> θ (− s<strong>in</strong> θ) ∂<br />

]<br />

P<br />

R s<strong>in</strong> θ ∂θ<br />

∂R ⃗ Hertz (t, R)<br />

[<br />

= − 1 ∂<br />

s<strong>in</strong> 2 θ ∂ ⃗ ]<br />

P Hertz<br />

= − 2 cos θ ∂P ⃗ Hertz<br />

R s<strong>in</strong> θ ∂θ ∂R R ∂R ;<br />

E θ = − 1<br />

R s<strong>in</strong> θ s<strong>in</strong> θ ∂ [<br />

R (− s<strong>in</strong> θ) ∂<br />

]<br />

P<br />

∂R<br />

∂R ⃗ Hertz (t, R) =<br />

(<br />

= s<strong>in</strong> θ ∂<br />

R ∂ ⃗ )<br />

P Hertz<br />

;<br />

R ∂R ∂R<br />

E φ = 0 .<br />

From the above expressions we can see that electric and magnetic fields are<br />

always perpendicular; magnetic l<strong>in</strong>es co<strong>in</strong>cide with circles parallel to the<br />

equator, while electric field l<strong>in</strong>es are <strong>in</strong> the meridian planes. Now let us<br />

further assume that<br />

f (t) = cos ωt ⇒<br />

(<br />

⃗p t − R ) (<br />

= ⃗p 0 cos ω t − R )<br />

c<br />

c<br />

or <strong>in</strong> a complex form<br />

(<br />

⃗p t − R )<br />

c<br />

= ⃗p 0 e iω(t− R c ) . (79)<br />

Then<br />

(<br />

∂P<br />

∂R = ∂<br />

∂R<br />

( 1<br />

= −<br />

R + iω c<br />

)<br />

p 0 e iω(t− R c )<br />

R<br />

)<br />

P (R, t) ,<br />

= − 1 R 2 p 0e iω (t− R c ) −<br />

iω<br />

c<br />

1<br />

R p 0e iω (t− R c ) =<br />

and<br />

(<br />

∂<br />

R ∂P )<br />

∂R ∂R<br />

= − ∂ [(<br />

1 + iωR ) ]<br />

P =<br />

∂R c<br />

( 1<br />

R + iω )<br />

c − ω2 R<br />

P .<br />

c<br />

52


Thus, for this particular case we get the follow<strong>in</strong>g result<br />

H φ = iω ( 1<br />

c s<strong>in</strong> θ R + iω )<br />

P (R, t) ;<br />

c<br />

( 1<br />

E R = 2 cos θ<br />

R + iω )<br />

P (R, t) ;<br />

2 cR<br />

( 1<br />

E θ = s<strong>in</strong> θ<br />

R + iω )<br />

2 cR − ω2<br />

P (R, t) .<br />

c 2<br />

These are the exact expressions for electromagnetic fields of a harmonic oscillator.<br />

They are complicated and we will look more closely only on what<br />

happens close and far away from the oscillator. To do that we will aid ourselves<br />

with the concept of a characteristic scale, which is determ<strong>in</strong>ed by the<br />

competition between<br />

1<br />

R and ω<br />

c = 2π<br />

T c = 2π λ ,<br />

where T and λ are the period and the wavelength of the electromagnetic<br />

wave, respectively.<br />

3.2.1 Close to the oscillator<br />

By “close to the oscillator” we mean:<br />

R ≪ λ<br />

2π<br />

or<br />

1<br />

R ≫ ω c = 2π λ ,<br />

i.e. distances from oscillator are smaller than the wavelength. Thus we can<br />

simplify (<br />

ω t − R )<br />

= ωt − R ω 2πR<br />

= ωt −<br />

c<br />

c λ ≈ ωt ,<br />

so that<br />

P (t, R) = p ( )<br />

t − R c<br />

≈ p (t)<br />

R R .<br />

Us<strong>in</strong>g the “close to oscillator condition”, fields are determ<strong>in</strong>ed by the electric<br />

moment p (t) and its derivative ∂p without retard<strong>in</strong>g<br />

∂t<br />

H φ ≈ iω c s<strong>in</strong> θ P R ≈ iω c<br />

s<strong>in</strong> θ<br />

p (t)<br />

R 2<br />

53<br />

= 1 c<br />

s<strong>in</strong> θ<br />

R 2<br />

∂p (t)<br />

∂t<br />

,


ecause iωp (t) = ∂p(t) , which follows from the particular choice of the time<br />

∂t<br />

dependence of the oscillator that we have made <strong>in</strong> (79). Similarly <strong>in</strong> this<br />

limit the electric field components become<br />

E R = 2 cos θ P = 2 cos θ p (t) ;<br />

R 2 R 3<br />

E θ = s<strong>in</strong> θ<br />

R P = s<strong>in</strong> θ<br />

2 R p (t) . 3<br />

At any given moment t, this is a field of a static dipole. For the magnetic<br />

field we f<strong>in</strong>d<br />

⃗H = 1 [ ] ∂⃗p (t)<br />

, R<br />

cR 3 ∂t<br />

⃗ = J [ ]<br />

⃗l, R ⃗ .<br />

cR 3<br />

Given that this <strong>in</strong>troduced current J obeys J ⃗ l = ∂⃗p(t) , this expression gives<br />

∂t<br />

the magnetic field of a current element of length l. This is known as the<br />

Biot-Savart law 12 .<br />

3.2.2 Far away from the oscillator<br />

Let us now consider the region far away from the oscillator, i.e. the region<br />

where<br />

R ≫ λ 1<br />

or<br />

2π R ≪ ω c = 2π λ .<br />

Distances greater than the wavelength are called wave-zone. In this particular<br />

limit our field components become<br />

H φ = − ω2<br />

s<strong>in</strong> θP = −ω2<br />

c2 c s<strong>in</strong> θ p ( )<br />

t − R c<br />

;<br />

2 R<br />

E R = 0 ;<br />

E θ = − ω2<br />

c s<strong>in</strong> θ p ( )<br />

t − R c<br />

= H 2 φ .<br />

R<br />

Thus summariz<strong>in</strong>g we get<br />

E R = E φ = H R = H θ = 0 ,<br />

and<br />

(<br />

E θ = H φ = − ω2 s<strong>in</strong> θ<br />

c 2 R<br />

p 0 cos ω t − R )<br />

,<br />

c<br />

12 Note that E ∼ 1 R 3 and H ∼ 1 R 2 .<br />

54


or<br />

E θ = H φ = s<strong>in</strong> θ ∂ 2 p ( )<br />

t − R c<br />

.<br />

c 2 R ∂t 2<br />

This last result is valid for any arbitrary p (t), not necessarily p 0 f (t), because<br />

we can always perform a harmonic Fourier decomposition of any function.<br />

Thus <strong>in</strong> the wave zone the electric and magnetic fields are equal to each other<br />

and vanish as 1 R . Additionally, vectors ⃗ E, ⃗ H, and ⃗ R are perpendicular 13 .<br />

Note that the phase of ⃗ E and ⃗ H, i.e. ω ( t − R c<br />

)<br />

moves with the speed of<br />

light.<br />

Thus, <strong>in</strong> the wave zone of the oscillator an electromagnetic wave is propagat<strong>in</strong>g!<br />

λ = cT = 2πc<br />

ω .<br />

This wave propagates <strong>in</strong> the radial direction i.e. phase depends on the distance<br />

to the center.<br />

Let us now look at the Poynt<strong>in</strong>g vector<br />

S = c<br />

4π<br />

[ ]∣ ∣ ⃗E, H ⃗ ∣∣ c =<br />

4π EH = 1<br />

4π<br />

(<br />

s<strong>in</strong> 2 θ ∂ 2 p ( )) 2<br />

t − R c<br />

,<br />

c 3 R 2 ∂t 2<br />

where on the first step we have used the fact that the electric and the magnetic<br />

fields are perpendicular. Additionally note that the second derivative<br />

with respect to time <strong>in</strong>side the square is an acceleration. Energy flux through<br />

the sphere of radius R is<br />

Σ =<br />

=<br />

∫ 2π ∫ π<br />

0 0<br />

∫ 2π ∫ π<br />

0<br />

0<br />

SR 2 s<strong>in</strong> θdφdθ =<br />

(<br />

∂ 2 p ( )) 2 [<br />

t − R c<br />

R 2 s<strong>in</strong> θdφdθ = 2 ∂ 2 p ( )] 2<br />

t − R c<br />

.<br />

c 3 R 2 ∂t 2 3c 3 ∂t 2<br />

1 s<strong>in</strong> 2 θ<br />

4π<br />

13 Note that ⃗ E, ⃗ H and ⃗ R have completely mismatch<strong>in</strong>g components i.e. if one vector<br />

has a particular non-zero component, for the other two this component is zero.<br />

55


For p ( t − R c<br />

)<br />

= p0 cos ω ( t − R c<br />

)<br />

the flux for one period is<br />

∫ T<br />

0<br />

Σdt = 2<br />

3c 3 p2 0ω 4<br />

∫T<br />

0<br />

(<br />

cos 2 ω t − R )<br />

dt =<br />

c<br />

= p2 0ω 4 T<br />

= 2πp2 0ω 3<br />

= 2πp2 0<br />

3c 3 3c 3 3<br />

The averaged radiation <strong>in</strong> a unit time is then<br />

( 2π<br />

λ<br />

) 3<br />

.<br />

〈 Σ 〉 = 1 T<br />

∫ T<br />

0<br />

Σdt = cp2 0<br />

3<br />

( 2π<br />

λ<br />

) 4<br />

. (80)<br />

Thus the oscillator cont<strong>in</strong>uously radiates energy <strong>in</strong>to surround<strong>in</strong>g space with<br />

average rate 〈 Σ 〉 ∼ p 2 0 1 λ 4 . In particular this expla<strong>in</strong>s that when transmitt<strong>in</strong>g<br />

radio signals by telegraph<strong>in</strong>g one should use waves of relatively short wavelengths<br />

14 (or equivalently high frequencies ω). On the other hand, radiation<br />

of low frequency currents is highly supressed, which expla<strong>in</strong>es the effect of<br />

the sky appear<strong>in</strong>g <strong>in</strong> blue, which is to the high frequency end of the visible<br />

light 15 spectrum.<br />

Lastly, let us f<strong>in</strong>ally focus on the concept of resistance to radiation, which<br />

is given by R λ such that<br />

〈 Σ 〉 = R λ 〈 J 2 〉 .<br />

Recall that we have previously def<strong>in</strong>ed J such that it obeys J ⃗ l = ∂⃗p(t− R c )<br />

∂t<br />

.<br />

Us<strong>in</strong>g this def<strong>in</strong>ition, we get<br />

〈 J 2 〉 = 1 T<br />

∫ T<br />

0<br />

= 1<br />

T l 2<br />

∫T<br />

0<br />

J 2 dt = 1<br />

T l 2<br />

p 2 0ω 2 s<strong>in</strong> 2 ω<br />

∫T<br />

0<br />

( (<br />

∂⃗p t −<br />

R<br />

c<br />

∂t<br />

(<br />

t − R c<br />

)) 2<br />

dt =<br />

)<br />

dt = p2 0ω 2<br />

T l 2<br />

π<br />

ω = πp2 0ω 2<br />

l 2 2π<br />

ω = p2 0ω 2<br />

2l .<br />

ω 2<br />

14 Generaly these range from tens of meters to tens of kilometers.<br />

15 In this case charge polarised chemical bonds between the atoms <strong>in</strong> the particles <strong>in</strong> the<br />

atmosphere act as little oscillators.<br />

56


Us<strong>in</strong>g the result (80), it is now easy to f<strong>in</strong>d R λ<br />

R λ = cp2 0<br />

3<br />

( ) 4 2π 2l 2<br />

λ p 2 0ω = 2c ( ) 4 2π 1<br />

2 3l 2 (<br />

λ 2π c) 2 = 2 3c<br />

λ<br />

( ) 2 2πl<br />

.<br />

λ<br />

4 Problem Set<br />

Problem 1<br />

A conductor is a material <strong>in</strong>side of which electric charges can freely move<br />

under the <strong>in</strong>fluence of an electric field. In the electrostatic equilibrium charge<br />

is distributed over the surface of a charged conductor. Us<strong>in</strong>g the Gauss<br />

theorem along with ∫<br />

⃗E · d ⃗ l = 0 ,<br />

show that<br />

• the electric field on the surface of a conductor is always normal to this<br />

surface;<br />

• the value of the electric field on the surface is E = 4πσ, where σ is the<br />

surface charged density.<br />

Problem 2<br />

The simplest capacitor is made of two isolated conductors A and B placed<br />

close to each other. If these two conductors are equally but oppositely<br />

charged with charges q and −q then they will acquire potentials ϕ A and<br />

ϕ B respectively. The ratio of the charge to the difference of the potentials is<br />

called a capacitance of a capacitor<br />

C =<br />

q<br />

ϕ A − ϕ B<br />

.<br />

Us<strong>in</strong>g the Gauss theorem f<strong>in</strong>d the capacitance of<br />

• two big plates of surface area S placed at a distance d from each other;<br />

57


• two concentric spheres with radii R 1 and R 2 (R 2 > R 1 );<br />

• two concentric cyl<strong>in</strong>ders of length L and radii R 1 and R 2 (R 2 > R 1 ).<br />

Problem 3<br />

F<strong>in</strong>d the value α ≡ α(d), for which a function<br />

ϕ ∼<br />

1<br />

|x − x ′ | α(d)<br />

is harmonic <strong>in</strong> d dimensions (i.e.<br />

∇ 2 ϕ = 0 outside x = x ′ ).<br />

it is a solution of the Laplace equation<br />

Problem 4<br />

Express the Lorentz <strong>in</strong>variants<br />

F µν F µν and ɛ µνρσ F µν F ρσ<br />

through the electric and magnetic fields ⃗ E and ⃗ H.<br />

Problem 5<br />

F<strong>in</strong>d the advanced Green’s function G adv for the <strong>in</strong>homogeneous wave equation<br />

△ϕ − 1 ∂ 2 ϕ<br />

= −4πf(⃗x, t) .<br />

c 2 ∂t2 F<strong>in</strong>d the Pauli Green’s function, which is the difference between the advanced<br />

and retarded Green’s functions<br />

G Pauli = G adv − G ret .<br />

Show that it solves the homogeneous wave equation.<br />

58


Literature<br />

The follow<strong>in</strong>g literature is recommended<br />

1. L. D. Landau and E. M. Lifshits, <strong>Classical</strong> Theory of Fields (3rd ed.)<br />

1971, London: Pergamon. Vol. 2 of the Course of Theoretical Physics.<br />

2. J. D. Jackson, <strong>Classical</strong> <strong>Electrodynamics</strong>, John Wiley & Sons, 1975.<br />

3. I. E. Tamm, Fundamentals of the theory of electricity, MIR (1979)<br />

(Translated from Russian).<br />

4. B. Thiedé, Electromagnetic Field Theory, Upsilon Books, Uppsala,<br />

Sweden.<br />

59

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!