You are on page 1of 31

Maxwell's equations

Electromagnetism

Electricity · Magnetism
Electrostatics
showMagnetostatics
[hide]Electrodynamics
Free space · Lorentz force law ·
emf · Electromagnetic induction ·
Faraday’s law · Lenz's law ·
Displacement current · Maxwell's
equations · EM field ·
Electromagnetic radiation · Liénard-
Wiechert Potential · Maxwell
tensor · Eddy current
[show]Electrical Network
[show]Covariant formulation
[show]Scientists
This box: view • talk • edit

Maxwell's equations are a set of four partial differential equations that relate the electric and
magnetic fields to their sources, charge density and current density. These equations can be
combined to show that light is an electromagnetic wave. Individually, the equations are
known as Gauss's law, Gauss's law for magnetism, Faraday's law of induction, and Ampère's
law with Maxwell's correction. The set of equations is named after James Clerk Maxwell.

These four equations, together with the Lorentz force law are the complete set of laws of
classical electromagnetism. The Lorentz force law itself was actually derived by Maxwell
under the name of Equation for Electromotive Force and was one of an earlier set of eight
equations by Maxwell.

Contents
 1 Conceptual description
 2 General formulation
 3 History
o 3.1 The term Maxwell's equations
o 3.2 Maxwell's On Physical Lines of Force (1861)
o 3.3 Maxwell's A Dynamical Theory of the Electromagnetic Field (1864)
o 3.4 A Treatise on Electricity and Magnetism (1873)
 4 Maxwell's equations and matter
o 4.1 Bound charge and current
 4.1.1 Proof that the two general formulations are equivalent
o 4.2 Constitutive relations
 4.2.1 Case without magnetic or dielectric materials
 4.2.2 Case of linear materials
 4.2.3 General case
 4.2.4 Maxwell's equations in terms of E and B for linear materials
 4.2.5 Calculation of constitutive relations
o 4.3 In vacuum
o 4.4 With magnetic monopoles
 5 Boundary conditions: using Maxwell's equations
 6 CGS units
 7 Special relativity
o 7.1 Historical developments
o 7.2 Covariant formulation of Maxwell's equations
 8 Potentials
 9 Four-potential
 10 Differential forms
o 10.1 Conceptual insight from this formulation
 11 Classical electrodynamics as the curvature of a line bundle
 12 Curved spacetime
o 12.1 Traditional formulation
o 12.2 Formulation in terms of differential forms
 13 See also
 14 Notes
 15 References
 16 Further reading
o 16.1 Journal articles
o 16.2 University level textbooks
 16.2.1 Undergraduate
 16.2.2 Graduate
 16.2.3 Older classics
 16.2.4 Computational techniques
 17 External links
o 17.1 Modern treatments
o 17.2 Historical
o 17.3 Other

Conceptual description
This section will conceptually describe each of the four Maxwell's equations, and also how
they link together to explain the origin of electromagnetic radiation such as light. The exact
equations are set out in later sections of this article.
 Gauss' law describes how an electric field is generated by electric charges: The
electric field tends to point away from positive charges and towards negative charges.
More technically, it relates the electric flux through any hypothetical closed
"Gaussian surface" to the electric charge within the surface.

 Gauss' law for magnetism states that there are no "magnetic charges" (also called
magnetic monopoles), analogous to electric charges.[1] Instead the magnetic field is
generated by a configuration called a dipole, which has no magnetic charge but
resembles a positive and negative charge inseparably bound together. Equivalent
technical statements are that the total magnetic flux through any Gaussian surface is
zero, or that the magnetic field is a solenoidal vector field.

An Wang's magnetic core memory (1954) is an application of Ampere's law. Each core stores
one bit of data.

 Faraday's law describes how a changing magnetic field can create ("induce") an
electric field.[1] This aspect of electromagnetic induction is the operating principle
behind many electric generators: A bar magnet is rotated to create a changing
magnetic field, which in turn generates an electric field in a nearby wire. (Note: The
"Faraday's law" that occurs in Maxwell's equations is a bit different than the version
originally written by Michael Faraday. Both versions are equally true laws of physics,
but they have different scope, for example whether "motional EMF" is included. See
Faraday's law of induction for details.)

 Ampère's law with Maxwell's correction states that magnetic fields can be generated
in two ways: by electrical current (this was the original "Ampère's law") and by
changing electric fields (this was "Maxwell's correction").

Maxwell's correction to Ampère's law is particularly important: It means that a changing


magnetic field creates an electric field, and a changing electric field creates a magnetic field.
[1][2]
Therefore, these equations allow self-sustaining "electromagnetic waves" to travel
through empty space (see electromagnetic wave equation).

The speed calculated for electromagnetic waves, which could be predicted from experiments
on charges and currents,[3] exactly matches the speed of light; indeed, light is one form of
electromagnetic radiation (as are X-rays, radio waves, and others). Maxwell understood the
connection between electromagnetic waves and light in 1864, thereby unifying the
previously-separate fields of electromagnetism and optics.
[edit] General formulation
The equations in this section are given in SI units. Unlike the equations of mechanics (for
example), Maxwell's equations are not unchanged in other unit systems. Though the general
form remains the same, various definitions get changed and different constants appear at
different places. Other than SI (used in engineering), the units commonly used are Gaussian
units (based on the cgs system and considered to have some theoretical advantages over SI[4]),
Lorentz-Heaviside units (used mainly in particle physics) and Planck units (used in
theoretical physics). See below for CGS-Gaussian units.

Two equivalent, general formulations of Maxwell's equations follow. The first separates
bound charge and bound current (which arise in the context of dielectric and/or magnetized
materials) from free charge and free current (the more conventional type of charge and
current). This separation is useful for calculations involving dielectric or magnetized
materials. The second formulation treats all charge equally, combining free and bound charge
into total charge (and likewise with current). This is the more fundamental or microscopic
point of view, and is particularly useful when no dielectric or magnetic material is present.
More details, and a proof that these two formulations are mathematically equivalent, are
given in section 4.

Symbols in bold represent vector quantities, and symbols in italics represent scalar
quantities. The definitions of terms used in the two tables of equations are given in another
table immediately following.

Formulation in terms of free charge and current

Name Differential form Integral form

Gauss's law

Gauss's law for magnetism

Maxwell–Faraday equation
(Faraday's law of induction)

Ampère's circuital law


(with Maxwell's correction)

Formulation in terms of total charge and current[note 1]
Name Differential form Integral form

Gauss's law

Gauss's law for


magnetism
Maxwell–Faraday
equation
(Faraday's law of
induction)
Ampère's circuital
law
(with Maxwell's
correction)

The following table provides the meaning of each symbol and the SI unit of measure:

Definitions and units


Meaning (first term is the most
Symbol SI Unit of Measure
common)
volt per meter or,
  electric field equivalently,
newton per coulomb
magnetic field tesla, or equivalently,
also called the magnetic induction weber per square meter,

also called the magnetic field density volt-second per square
also called the magnetic flux density meter
electric displacement field coulombs per square
  also called the electric induction meter or equivalently,
also called the electric flux density newton per volt-meter
magnetizing field
also called auxiliary magnetic field
  ampere per meter
also called magnetic field intensity
also called magnetic field
  the divergence operator per meter (factor
contributed by applying
  the curl operator
either operator)
per second (factor
partial derivative with respect to time contributed by applying
  the operator)
differential vector element of surface area
  A, with infinitesimally small magnitude square meters
and direction normal to surface S
differential vector element of path length
  meters
tangential to the path/curve
permittivity of free space, also called the
  farads per meter
electric constant, a universal constant
henries per meter, or
permeability of free space, also called the
  newtons per ampere
magnetic constant, a universal constant
squared
free charge density (not including bound coulombs per cubic

charge) meter
total charge density (including both free coulombs per cubic

and bound charge) meter
free current density (not including bound amperes per square
  current) meter
total current density (including both free amperes per square

and bound current) meter
net free electric charge within the three-
  dimensional volume V (not including coulombs
bound charge)
net electric charge within the three-
  dimensional volume V (including both coulombs
free and bound charge)
line integral of the electric field along the
boundary ∂S of a surface S (∂S is always a joules per coulomb
  closed curve).
line integral of the magnetic field over the
tesla-meters
  closed boundary ∂S of the surface S
the electric flux (surface integral of the
electric field) through the (closed) surface joule-meter per coulomb
  (the boundary of the volume V)
the magnetic flux (surface integral of the
magnetic B-field) through the (closed) tesla meters-squared or
  surface (the boundary of the volume webers
V)
magnetic flux through any surface S, not webers or equivalently,
  necessarily closed volt-seconds
electric flux through any surface S, not joule-meters per
  necessarily closed coulomb

flux of electric displacement field through
coulombs
any surface S, not necessarily closed

net free electrical current passing through


the surface S (not including bound amperes
  current)
net electrical current passing through the
surface S (including both free and bound amperes
  current)

Maxwell's equations are generally applied to macroscopic averages of the fields, which vary
wildly on a microscopic scale in the vicinity of individual atoms (where they undergo
quantum mechanical effects as well). It is only in this averaged sense that one can define
quantities such as the permittivity and permeability of a material. At microscopic level,
Maxwell's equations, ignoring quantum effects, describe fields, charges and currents in free
space—but at this level of detail one must include all charges, even those at an atomic level,
generally an intractable problem.

[edit] History
Although James Clerk Maxwell is said by some not to be the originator of these equations, he
nevertheless derived them independently in conjunction with his molecular vortex model of
Faraday's "lines of force". In doing so, he made an important addition to Ampère's circuital
law.

All four of what are now described as Maxwell's equations can be found in recognizable form
(albeit without any trace of a vector notation, let alone ∇) in his 1861 paper On Physical
Lines of Force, in his 1865 paper A Dynamical Theory of the Electromagnetic Field, and also
in vol. 2 of Maxwell's "A Treatise on Electricity & Magnetism", published in 1873, in
Chapter IX, entitled "General Equations of the Electromagnetic Field". This book by
Maxwell pre-dates publications by Heaviside, Hertz and others.

[edit] The term Maxwell's equations

The term Maxwell's equations originally applied to a set of eight equations published by
Maxwell in 1865, but nowadays applies to modified versions of four of these equations that
were grouped together in 1884 by Oliver Heaviside,[6] concurrently with similar work by
Willard Gibbs and Heinrich Hertz.[7] These equations were also known variously as the Hertz-
Heaviside equations and the Maxwell-Hertz equations,[6] and are sometimes still known as the
Maxwell–Heaviside equations.[8]

Maxwell's contribution to Science in producing these equations lies in the correction he made
to Ampère's circuital law in his 1861 paper On Physical Lines of Force. He added the
displacement current term to Ampère's circuital law and this enabled him to derive the
electromagnetic wave equation in his later 1865 paper A Dynamical Theory of the
Electromagnetic Field and demonstrate the fact that light is an electromagnetic wave. This
fact was then later confirmed experimentally by Heinrich Hertz in 1887.

The concept of fields was introduced by, among others, Faraday. Albert Einstein wrote:

The precise formulation of the time-space laws was the work of Maxwell. Imagine his
feelings when the differential equations he had formulated proved to him that electromagnetic
fields spread in the form of polarised waves, and at the speed of light! To few men in the
world has such an experience been vouchsafed . . it took physicists some decades to grasp the
full significance of Maxwell's discovery, so bold was the leap that his genius forced upon the
conceptions of his fellow-workers
—(Science, May 24, 1940)

The equations were called by some the Hertz-Heaviside equations, but later Einstein referred
to them as the Maxwell-Hertz equations.[6] However, in 1940 Einstein referred to the
equations as Maxwell's equations in "The Fundamentals of Theoretical Physics" published in
the Washington periodical Science, May 24, 1940.

Heaviside worked to eliminate the potentials (electrostatic potential and vector potential) that
Maxwell had used as the central concepts in his equations;[6] this effort was somewhat
controversial,[9] though it was understood by 1884 that the potentials must propagate at the
speed of light like the fields, unlike the concept of instantaneous action-at-a-distance like the
then conception of gravitational potential.[7] Modern analysis of, for example, radio antennas,
makes full use of Maxwell's vector and scalar potentials to separate the variables, a common
technique used in formulating the solutions of differential equations. However the potentials
can be introduced by algebraic manipulation of the four fundamental equations.

The net result of Heaviside's work was the symmetrical duplex set of four equations,[6] all of
which originated in Maxwell's previous publications, in particular Maxwell's 1861 paper On
Physical Lines of Force, the 1865 paper A Dynamical Theory of the Electromagnetic Field
and the Treatise. The fourth was a partial time derivative version of Faraday's law of
induction that doesn't include motionally induced EMF; this version is often termed the
Maxwell-Faraday equation or Faraday's law in differential form to keep clear the distinction
from Faraday's law of induction, though it expresses the same law.[10][11]

[edit] Maxwell's On Physical Lines of Force (1861)

The four modern day Maxwell's equations appeared throughout Maxwell's 1861 paper On
Physical Lines of Force:

i. Equation (56) in Maxwell's 1861 paper is .


ii. Equation (112) is Ampère's circuital law with Maxwell's displacement current added.
It is the addition of displacement current that is the most significant aspect of
Maxwell's work in electromagnetism, as it enabled him to later derive the
electromagnetic wave equation in his 1865 paper A Dynamical Theory of the
Electromagnetic Field, and hence show that light is an electromagnetic wave. It is
therefore this aspect of Maxwell's work which gives the equations their full
significance. (Interestingly, Kirchhoff derived the telegrapher's equations in 1857
without using displacement current. But he did use Poisson's equation and the
equation of continuity which are the mathematical ingredients of the displacement
current. Nevertheless, Kirchhoff believed his equations to be applicable only inside an
electric wire and so he is not credited with having discovered that light is an
electromagnetic wave).
iii. Equation (115) is Gauss's law.
iv. Equation (54) is an equation that Oliver Heaviside referred to as 'Faraday's law'. This
equation caters for the time varying aspect of electromagnetic induction, but not for
the motionally induced aspect, whereas Faraday's original flux law caters for both
aspects. Maxwell deals with the motionally dependent aspect of electromagnetic
induction, v × B, at equation (77). Equation (77) which is the same as equation (D) in
the original eight Maxwell's equations listed below, corresponds to all intents and
purposes to the modern day force law F = q ( E + v × B ) which sits adjacent to
Maxwell's equations and bears the name Lorentz force, even though Maxwell derived
it when Lorentz was still a young boy.

The difference between the and the vectors can be traced back to Maxwell's 1855 paper
entitled On Faraday's Lines of Force which was read to the Cambridge Philosophical
Society. The paper presented a simplified model of Faraday's work, and how the two
phenomena were related. He reduced all of the current knowledge into a linked set of
differential equations.
Figure of Maxwell's molecular vortex model. For a uniform magnetic field, the field lines
point outward from the display screen, as can be observed from the black dots in the middle
of the hexagons. The vortex of each hexagonal molecule rotates counter-clockwise. The small
green circles are clockwise rotating particles sandwiching between the molecular vortices.

It is later clarified in his concept of a sea of molecular vortices that appears in his 1861 paper
On Physical Lines of Force - 1861. Within that context, represented pure vorticity (spin),
whereas was a weighted vorticity that was weighted for the density of the vortex sea.
Maxwell considered magnetic permeability µ to be a measure of the density of the vortex sea.
Hence the relationship,

(1) Magnetic induction current causes a magnetic current density

was essentially a rotational analogy to the linear electric current relationship,

(2) Electric convection current

where ρ is electric charge density. was seen as a kind of magnetic current of vortices
aligned in their axial planes, with being the circumferential velocity of the vortices. With µ
representing vortex density, it follows that the product of µ with vorticity leads to the
magnetic field denoted as .

The electric current equation can be viewed as a convective current of electric charge that
involves linear motion. By analogy, the magnetic equation is an inductive current involving
spin. There is no linear motion in the inductive current along the direction of the vector.
The magnetic inductive current represents lines of force. In particular, it represents lines of
inverse square law force.
The extension of the above considerations confirms that where is to , and where is to ρ,
then it necessarily follows from Gauss's law and from the equation of continuity of charge
that is to . i.e. parallels with , whereas parallels with .

[edit] Maxwell's A Dynamical Theory of the Electromagnetic Field (1864)

Main article: A Dynamical Theory of the Electromagnetic Field

In 1864 Maxwell published A Dynamical Theory of the Electromagnetic Field in which he


showed that light was an electromagnetic phenomenon. Confusion over the term "Maxwell's
equations" is exacerbated because it is also sometimes used for a set of eight equations that
appeared in Part III of Maxwell's 1864 paper A Dynamical Theory of the Electromagnetic
Field, entitled "General Equations of the Electromagnetic Field,"[12] a confusion compounded
by the writing of six of those eight equations as three separate equations (one for each of the
Cartesian axes), resulting in twenty equations and twenty unknowns. (As noted above, this
terminology is not common: Modern references to the term "Maxwell's equations" refer to
the Heaviside restatements.)

The eight original Maxwell's equations can be written in modern vector notation as follows:

(A) The law of total currents

(B) The equation of magnetic force

(C) Ampère's circuital law

(D) Electromotive force created by convection, induction, and by static electricity. (This is in
effect the Lorentz force)

(E) The electric elasticity equation

(F) Ohm's law

(G) Gauss's law

(H) Equation of continuity

Notation
is the magnetizing field, which Maxwell called the magnetic intensity.
is the electric current density (with being the total current including
displacement current).[note 2]
is the displacement field (called the electric displacement by Maxwell).
is the free charge density (called the quantity of free electricity by Maxwell).
is the magnetic vector potential (called the angular impulse by Maxwell).
is called the electromotive force by Maxwell. The term electromotive force is
nowadays used for voltage, but it is clear from the context that Maxwell's meaning
corresponded more to the modern term electric field.
is the electric potential (which Maxwell also called electric potential).
is the electrical conductivity (Maxwell called the inverse of conductivity the
specific resistance, what is now called the resistivity).

It is interesting to note the term that appears in equation D. Equation D is therefore


effectively the Lorentz force, similarly to equation (77) of his 1861 paper (see above).

When Maxwell derives the electromagnetic wave equation in his 1865 paper, he uses
equation D to cater for electromagnetic induction rather than Faraday's law of induction
which is used in modern textbooks. (Faraday's law itself does not appear among his
equations.) However, Maxwell drops the term from equation D when he is deriving
the electromagnetic wave equation, as he considers the situation only from the rest frame.

[edit] A Treatise on Electricity and Magnetism (1873)

English Wikisource has original text related to this article:


A Treatise on Electricity and Magnetism

In A Treatise on Electricity and Magnetism, an 1873 textbook on electromagnetism written


by James Clerk Maxwell, the equations are compiled into two sets.

The first set is

The second set is

[edit] Maxwell's equations and matter


[edit] Bound charge and current

Main articles: Bound charge#Bound charge and Bound current#Magnetization current


Left: A schematic view of how an assembly of microscopic dipoles appears like a
macroscopically separated pair of charged sheets, as shown at top and bottom (these sheets
are not intended to be viewed as originating the electric field that causes the dipole alignment,
but as a representation equivalent to the dipole array); Right: How an assembly of
microscopic current loops appears as a macroscopically circulating current loop. Inside the
boundaries, the individual contributions tend to cancel, but at the boundaries no cancellation
occurs.

If an electric field is applied to a dielectric material, each of the molecules responds by


forming a microscopic electric dipole—its atomic nucleus will move a tiny distance in the
direction of the field, while its electrons will move a tiny distance in the opposite direction.
This is called polarization of the material. In an idealized situation like that shown in the
figure, the distribution of charge that results from these tiny movements turns out to be
identical (outside the material) to having a layer of positive charge on one side of the
material, and a layer of negative charge on the other side (a macroscopic separation of
charge) even though all of the charges involved are bound to individual molecules. The
volume polarization P is a result of bound charge. (Mathematically, once physical
approximation has established the electric dipole density P based upon the underlying
behavior of atoms, the surface charge that is equivalent to the material with its internal
polarization is provided by the divergence theorem applied to a region straddling the interface
between the material and the surrounding vacuum.)[13][14]

Somewhat similarly, in all materials the constituent atoms exhibit magnetic moments that are
intrinsically linked to the angular momentum of the atoms' components, most notably their
electrons. The connection to angular momentum suggests the picture of an assembly of
microscopic current loops. Outside the material, an assembly of such microscopic current
loops is not different from a macroscopic current circulating around the material's surface,
despite the fact that no individual magnetic moment is traveling a large distance. The bound
currents can be described using M. (Mathematically, once physical approximation has
established the magnetic dipole density based upon the underlying behavior of atoms, the
surface current that is equivalent to the material with its internal magnetization is provided by
Stokes' theorem applied to a path straddling the interface between the material and the
surrounding vacuum.)[15][16]

These ideas suggest that for some situations the microscopic details of the atomic and
electronic behavior can be treated in a simplified fashion that ignores many details on a fine
scale that may be unimportant to understanding matters on a grosser scale. That notion
underlies the bound/free partition of behavior.
[edit] Proof that the two general formulations are equivalent

In this section, a simple proof is outlined which shows that the two alternate general
formulations of Maxwell's equations given in Section 1 are mathematically equivalent.

The relation between polarization, magnetization, bound charge, and bound current is as
follows:

where P and M are polarization and magnetization, and ρb and Jb are bound charge and
current, respectively. Plugging in these relations, it can be easily demonstrated that the two
formulations of Maxwell's equations given in Section 2 are precisely equivalent.

Maxwell's Equations and Electromagnetic


Waves
Michael Fowler, Physics Department, UVa  5/9/09

 
The Equations
Maxwell’s four equations describe the electric and magnetic fields arising from varying
distributions of electric charges and currents, and how those fields change in time.  The
equations were the mathematical distillation of decades of experimental observations of the
electric and magnetic effects of charges and currents.  Maxwell’s own contribution is just the
last term of the last equation but realizing the necessity of that term had dramatic
consequences.  It made evident for the first time that varying electric and magnetic fields
could feed off each other these fields could propagate indefinitely through space, far
from the varying charges and currents where they originated.  Previously the fields had been
envisioned as tethered to the charges and currents giving rise to them.   Maxwell’s new term
(he called it the displacement current) freed them to move through space in a self-sustaining
fashion, and even predicted their velocity it was the velocity of light!

Here are the equations:


1. Gauss’ Law for electric fields:  

 (The integral of the outgoing electric field over


an area enclosing a volume equals the total charge inside, in appropriate units.)

2. The corresponding formula for magnetic fields:

 (No magnetic charge exists: no “monopoles”.)


3. Faraday’s Law of Magnetic Induction: 

 The first
term is integrated round a closed line, usually a wire, and gives the total voltage
change around the circuit, which is generated by a varying magnetic field threading
through the circuit.
4. Ampere’s Law plus Maxwell’s displacement current: 

 This gives the total magnetic force around


a circuit in terms of the current through the circuit, plus any varying electric field
through the circuit (that’s the “displacement current”).

The purpose of this lecture is to review the first three equations and the original Ampere’s
law fairly briefly, as they were already covered earlier in the course, then to demonstrate
why the displacement current term must be added for consistency, and finally to show,
without using differential equations, how measured values of static electrical and
magnetic attraction are sufficient to determine the speed of light.

 
Preliminaries: Definitions of µ0 and ε0, the Ampere and the
Coulomb
Ampere discovered that two long parallel wires carrying electric currents in the same
direction attract each other magnetically, the force per unit length being proportional to the
product of the currents (so oppositely directed currents repel) and decaying with distance as
1/r.  In modern (SI) notation, his discovery is written (F in Newtons)

The modern convention is that the constant  appearing


-7
here is exactly 10 , this defines our present unit of current, the ampere.  To repeat:

is not something to measure experimentally, it's


-7
just a funny way of writing the number 10 !  That's not quite fair it has dimensions to
ensure that both sides of the above equation have the same dimensionality. (Of course, there's
a historical reason for this strange convention, as we shall see later).  Anyway, if we bear in
mind that dimensions have been taken care of, and just write the equation 

it's clear that this defines the unit current one ampere as that current in a long
straight wire which exerts a magnetic force of  newtons
per meter of wire on a parallel wire one meter away carrying the same current.

However, after we have established our unit of current the ampere we have also
thereby defined our unit of charge, since current is a flow of charge, and the unit of charge
must be the amount carried past a fixed point in unit time by unit current.  Therefore, our unit
of charge the coulomb is defined by stating that a one amp current in a wire
carries one coulomb per second past a fixed point.

 To be consistent, we must do electrostatics using this same unit of charge. Now, the

electrostatic force between two charges is

 The constant appearing here,

now written , must be experimentally measured


its value turns out to be .
 To summarize:  to find the value of , two experiments
have to be performed. We must first establish the unit of charge from the unit of current by
measuring the magnetic force between two current-carrying parallel wires.  Second, we must
find the electrostatic force between measured charges. (We could, alternatively, have defined
some other unit of current from the start, then we would have had to find both  
and  by experiments on magnetic and electrostatic attraction.  In fact, the ampere
was originally defined as the current that deposited a definite weight of silver per hour in an
electrolytic cell). 

 
Maxwell's Equations
We established earlier in the course that the total flux of electric field out of a closed surface

is just the total enclosed charge multiplied by , 

This is Maxwell’s first equation.  It represents completely covering the surface with a large
number of tiny patches having areas .  (The little areas are small enough to
be regarded as flat, the vector magnitude dA is just the value of the area, the direction of the
vector is perpendicular to the area element, pointing outwards away from the enclosed
volume.)  Hence the dot product with the electric field selects the component of that field
pointing perpendicularly outwards (it would count negatively if the field were pointing
inwards) this is the only component of the field that contributes to actual electric flux
across the surface.  (Remember flux just means flow the picture of the electric field in
this context is like a fluid flowing out from the charges, the field vector representing the
direction and velocity of the flowing fluid.)

 The second Maxwell equation is the analogous one for the magnetic field, which has no
sources or sinks (no magnetic monopoles, the field lines just flow around in closed curves).
Again thinking of the force lines as representing a kind of fluid flow, the so-called "magnetic
flux", we see that for a closed surface, as much magnetic flux flows into the surface as flows
out since there are no sources.  This can perhaps be visualized most clearly by taking a
group of neighboring lines of force forming a slender tube the "fluid" inside this tube
flows round and round, so as the tube goes into the closed surface then comes out again
(maybe more than once) it is easy to see that what flows into the closed surface at one place
flows out at another. Therefore the net flux out of the enclosed volume is zero,  Maxwell’s
second equation:
The first two Maxwell's equations, given above, are for integrals of the electric and magnetic
fields over closed surfaces.  Maxwell's other two equations, discussed below, are for integrals
of electric and magnetic fields around closed curves (taking the component of the field
pointing along the curve). These represent the work that would be needed to take a charge
around a closed curve in an electric field, and a magnetic monopole (if one existed!) around a
closed curve in a magnetic field.

 The simplest version of Maxwell's third equation is for the special electrostatic case

However, we know that this is only part of the truth, because from Faraday's Law of
Induction, if a closed circuit has a changing magnetic flux through it, a circulating current
will arise, which means there is a nonzero voltage around the circuit.

 The complete Maxwell's third equation is: 

where the area integrated over on the right hand side spans the path (or circuit) on the left
hand side, like a soap film on a loop of wire.  (The best way to figure out the sign is to use
Lenz’ law: the induced current will generate a magnetic field opposing the changing of the
external  field, so if an external upward field is decreasing, the current thereby generated
around the loop will give an upward pointing field.)

 It may seem that the integral on the right hand side is not very clearly defined, because if the
path or circuit lies in a plane, the natural choice of spanning surface (the "soap film") is flat,
but how do you decide what surface to choose to do the integral over for a wire bent into a
circuit that doesn’t lie in a plane?   The answer is that it doesn’t matter what surface you
choose, as long as the wire forms its boundary.  Consider two different surfaces both having
the wire as a boundary (just as both the northern hemisphere of the earth’s surface and the
southern hemisphere have the equator as a boundary).  If you add these two surfaces together,

they form a single closed surface, and we know that for a closed surface

.  This implies that

 for one of the two surfaces bounded by the path is equal to

 for the other one, so that the two will


add to zero for the whole closed surface.  But don’t forget these integrals for the whole closed
surface are defined with the little area vectors pointing outwards from the enclosed volume.
By imagining two surfaces spanning the wire that are actually close to each other, it is clear
that the integral over one of them is equal to the integral over the other if we take the
 vectors to point in the same direction for both of them, which in terms of the
enclosed volume would be outwards for one surface, inwards for the other one. The bottom

line of all this is that the surface integral  is


the same for any surface spanning the path, so it doesn’t matter which we choose. 

The equation analogous to the electrostatic version of the third equation given above, but for
the magnetic field, is Ampere's law, 

 for the magnetostatic case,

 where the currents counted are those threading through the path we're integrating around, so
if there is a soap film spanning the path, these are the currents that punch through the film (of
course, we have to agree on a direction, and subtract currents flowing in the opposite
direction).

 We must now consider whether this equation, like the electrostatic one, has limited validity.
In fact, it was not questioned for a generation after Ampere wrote it down: Maxwell's great
contribution, in the 1860's, was to realize that it was not always valid.

 
When Does Ampere's Law Go Wrong?
 A simple example to see that something must be wrong with Ampere's Law in the general
case is given by Feynman in his Lectures in Physics (II, 18-3).  Suppose we use a hypodermic
needle to insert a spherically symmetric blob of charge in the middle of a large vat of
solidified jello (which we assume conducts electricity).  Because of electrostatic repulsion,
the charge will dissipate, currents will flow outwards in a spherically symmetric way.
Question: does this outward-flowing current distribution generate a magnetic field?  The
answer must be no , because since we have a completely spherically symmetric situation, it
could only generate a spherically symmetric magnetic field.  But the only possible such fields
are one pointing outwards everywhere and one pointing inwards everywhere, both
corresponding to non-existent monopoles.  So, there can be no magnetic field.

 However, imagine we now consider checking Ampere's law by taking as a path a horizontal
circle with its center above the point where we injected the charge (think of a halo above
someone’s head.)  Obviously, the left hand side of Ampere's equation is zero, since there can
be no magnetic field.  On the other hand, the right hand side is most definitely not zero, since
some of the outward flowing current is going to go through our circle.  So the equation must
be wrong.
 Ampere's law was established as the result of large numbers of careful experiments on all
kinds of current distributions.  So how could it be that something of the kind we describe
above was overlooked?  The reason is really similar to why electromagnetic induction was
missed for so long.  No-one thought about looking at changing fields, all the experiments
were done on steady situations.  With our ball of charge spreading outward in the jello, there
is obviously a changing electric field.  Imagine yourself in the jello near where the charge
was injected: at first, you would feel a strong field from the nearby concentrated charge, but
as the charge spreads out spherically, some of it going past you, the field will decrease with
time.

 
Maxwell's Example
Maxwell himself gave a more practical example:  consider Ampere's law for the usual
infinitely long wire carrying a steady current I , but now break the wire at some point and put
in two large circular metal plates, a capacitor, maintaining the steady current I in the wire
everywhere else, so that charge is simply piling up on one of the plates and draining off the
other.

 Looking now at the wire some distance away from the plates, the situation appears normal,
and if we put the usual circular path around the wire, application of Ampere's law tells us that
the magnetic field at distance r , from 

is just

(Reminder on field direction:  the right hand rule if you curl the fingers of your right
hand around an imaginary wire, a current flowing in the direction indicated by your thumb
will generate circular magnetic field lines in the direction indicated by your fingers.)

Recall, however, that we defined the current threading the path in terms of current punching
through a soap film spanning the path, and said this was independent of whether the soap film
was flat, bulging out on one side, or whatever. With a single infinite wire, there was no
escape no contortions of this covering surface could wriggle free of the wire going
through it (actually, if you distort the surface enough, the wire could penetrate it several
times, but you have to count the net flow across the surface, and the new penetrations would
come in pairs with the current crossing the surface in opposite directions, so they would
cancel).

 Once we bring in Maxwell's parallel plate capacitor, however, there is a way to distort the
surface so that no current penetrates it at all: we can run it between the plates!
The question then arises: can we rescue Ampere's law by adding another term just as the
electrostatic version of the third equation was rescued by adding Faraday's induction term?
The answer is of course yes:  although there is no current crossing the surface if we put it
between the capacitor plates, there is certainly a changing electric field , because the
capacitor is charging up as the current I flows in. Assuming the plates are close together, we
can take all the electric field lines from the charge q on one plate to flow across to the other
plate, so the total electric flux across the surface between the plates, 

Now, the current in the wire, I , is just the rate of change of charge on the plate, 

Putting the above two equations together, we see that 

Ampere's law can now be written in a way that is correct no matter where we put the surface
spanning the path we integrate the magnetic field around: 

This is Maxwell’s fourth equation.

Notice that in the case of the wire, either the current in the wire, or the increasing electric
field, contribute on the right hand side, depending on whether we have the surface simply
cutting through the wire, or positioned between the plates. (Actually, more complicated
situations are possible we could imaging the surface partly between the plates, then
cutting through the plates to get out!  In this case, we would have to figure out the current
actually in the plate to get the right hand side, but the equation would still apply).

 
"Displacement Current"
 Maxwell referred to the second term on the right hand side, the changing electric field term,
as the "displacement current".  This was an analogy with a dielectric material.  If a dielectric
material is placed in an electric field, the molecules are distorted, their positive charges
moving slightly to the right, say, the negative charges slightly to the left.  Now consider what
happens to a dielectric in an increasing electric field.  The positive charges will be displaced
to the right by a continuously increasing distance, so, as long as the electric field is increasing
in strength, these charges are moving: there is actually a displacement current .  (Meanwhile,
the negative charges are moving the other way, but that is a current in the same direction, so
adds to the effect of the positive charges' motion.)  Maxwell's picture of the vacuum, the
aether, was that it too had dielectric properties somehow, so he pictured a similar motion of
charge in the vacuum to that we have just described in the dielectric.  This is why the
changing electric field term is often called the "displacement current", and in Ampere's law
(generalized) is just added to the real current, to give Maxwell's fourth and final
equation. 

 
Another Angle on the Fourth Equation: the Link to
Charge Conservation
Going back for a moment to Ampere's law, we stated it as:         

 for magnetostatics

where the currents counted are those threading through the path we’re integrating around, so
if there is a soap film spanning the path, these are the currents that punch through the film.
Our mental picture here is usually of a few thin wires, maybe twisted in various ways,
carrying currents. More generally, thinking of electrolytes, or even of fat wires, we should be
envisioning a current density varying from point to point in space. In other words, we have a
flux of current and the natural expression for the current threading our path is (analogous to

the magnetic flux in the third equation) to write a surface integral of the current density

 over a surface spanning the path, giving for magnetostatics 

path integral

,  (surface integral, over surface spanning path)


The question then arises as to whether the surface integral we have written on the right hand
side above depends on which surface we choose spanning the path. From an argument exactly
parallel to that for the magnetic flux in the third equation (see above), this will be true if and

only if    for a closed surface


(with the path lying in the surface this closed surface is made up by combining two
different surfaces spanning the path).

 Now,  taken over a closed surface is just the net


current flow out of the enclosed volume. Obviously, in a situation with steady currents
flowing along wires or through conductors, with no charge piling up or draining away from
anywhere, this is zero. However, if the total electric charge q , say, enclosed by the closed
surface is changing as time goes on, then evidently 

where we put in a minus sign because, with our convention,  is a little
vector pointing outwards, so the integral represents net flow of charge out from the surface,
equal to the rate of decrease of the enclosed total charge.

 To summarize: if the local charge densities are changing in time, that is, if charge is piling up

in or leaving some region, then

 over a closed surface around that region. That implies that

 over one surface spanning the wire will be different from

 over another surface spanning the wire if these two


surfaces together make up a closed surface enclosing a region containing a changing amount
of charge.

 The key to fixing this up is to realize that although

 it can be
written as another surface integral over the same surface, using the first Maxwell equation,
that is, the integral over a closed surface
where q is the total charge in the volume enclosed by the surface. 

By taking the time rate of change of both sides, we find 

Putting this together with

 gives: 

for any closed surface, and consequently this is a surface integral that must be the same for
any surface spanning the path or circuit!  (Because two different surfaces spanning the same
circuit add up to a closed surface. We’ll ignore the technically trickier case where the two
surfaces intersect each other, creating multiple volumes there one must treat each
created volume separately to get the signs right.) 

Therefore, this is the way to generalize Ampere's law from the magnetostatic situation to the
case where charge densities are varying with time, that is to say the path integral
and this gives the same result for any surface spanning the path. 

 
A Sheet of Current: A Simple Magnetic Field
As a preliminary to looking at electromagnetic waves, we consider the magnetic field
configuration from a sheet of uniform current of large extent.  Think of the sheet as
perpendicular to this sheet of paper, the current running vertically down into the paper.   It
might be helpful to visualize the sheet as many equal parallel fine wires uniformly spaced
close together, carrying equal (small) currents: 

...................................................................................... (wires)

The magnetic field from this current sheet can be found using Ampere's law applied to a
rectangular contour in the plane of the paper, with the current sheet itself bisecting the
rectangle, so the rectangle's top and bottom are equidistant from the current sheet in opposite
directions.

B-field

Rectangular contour

Infinite current sheet:

flow direction into paper

A plane rectangular contour in the magnetic field of an infinite current sheet


 Applying Ampere’s law to the above rectangular contour, there are contributions to

 (taken clockwise) only from the top and


bottom, and they add to give 2BL if the rectangle has side L. The total current enclosed by the
rectangle is IL, taking the current density of the sheet to be I amperes per meter (how many
little wires per meter multiplied by the current in each wire). 

Thus,

 immediately gives: 

B = µ0I/2 

a magnetic field strength independent of distance d from the sheet. (This is the magnetostatic
analog of the electrostatic result that the electric field from an infinite sheet of charge is
independent of distance from the sheet.) In real life, where there are no infinite sheets of
anything, these results are good approximations for distances from the sheet small compared
with the extent of the sheet. 

 
Switching on the Sheet: How Fast Does the Field Build
Up? 
Consider now how the magnetic field develops if the current in the sheet is suddenly
switched on at time t = 0. We will assume that sufficiently close to the sheet, the magnetic
field pattern found above using Ampere's law is rather rapidly established.  

In fact, we will assume further that the magnetic field spreads out from the sheet like a tidal
wave, moving in both directions at some speed v , so that after time t the field within distance
vt of the sheet is the same as that found above for the magnetostatic case, but beyond vt there
is at that instant no magnetic field present.

Let us now apply Maxwell's equations to this guess to see if it can make sense. Certainly
Ampere's law doesn't work by itself, because if we take a rectangular path as we did in the
previous section, for d < vt everything works as before, but for a rectangle extending beyond
the spreading magnetic field, d > vt , there will be no magnetic field contribution from the top
and bottom of the rectangle, and hence 

but there is definitely enclosed current!  

We are forced to conclude that for Maxwell's fourth equation to be correct, there must also
be a changing electric field through the rectangular contour. 

Let us now try to nail down what this electric field through the contour must look like.  First,
it must be through the contour, that is, have a component perpendicular to the plane of the
contour, in other words, perpendicular to the magnetic field.  In fact, electric field
components in other directions won't affect the fourth equation we are trying to satisfy, so we
shall ignore them.  Notice first that for a rectangular contour with d < vt,  Ampere's law
works, so we don't want a changing electric field through such a contour (but a constant
electric field would be ok). 

Now apply Maxwell's fourth equation to a rectangular contour with d > vt, 

B-field

Rectangular contour

Infinite current sheet:

flow direction into paper

A plane rectangular contour in the magnetic field of an infinite current sheet just switched
on: the contour goes beyond the region the new B-field has reached.

vt
No magnetic field here yet

It is:  path integral

 (over surface spanning path).   

For the rectangle shown above, the integral on the left hand side is zero because  is

perpendicular to  along the sides, so the dot product is zero, and  


is zero at the top and bottom, because the outward moving "wave" of magnetic field hasn’t
gotten there yet. Therefore, the right hand side of the equation must also be zero. 
We know ,

so we must have:  

 
Finding the Speed of the Outgoing Field Front: the
Connection with Light
So, as long as the outward moving front of magnetic field, travelling at v , hasn't reached the
top and bottom of the rectangular contour, the electric field through the contour increases
linearly with time, but the increase drops to zero (because Ampere's law is satisfied) the
moment the front reaches the top and bottom of the rectangle. The simplest way to get this
behavior is to have an electric field of strength E, perpendicular to the magnetic field,
everywhere there is a magnetic field, so the electric field also spreads outwards at speed v.
(Note that, unlike the magnetic field, the electric field must point the same way on both sides
of the current sheet, otherwise its net flux through the rectangle would be zero.) 

The electric and magnetic field propagating to the right from a current sheet in the plane 0xy,
current in the negative x-direction, suddenly switched on. (To the left of the sheet, there is
similar propagation, but with reversed magnetic field.)

Magnetic Field

Electric Field

x
After time t , then, the electric field flux through the rectangular contour

 (in the yz-plane in the diagram above) will be just field x area
= E.2.vtL , and the rate of change will be 2EvL . (It's spreading both ways, hence the 2). 

Therefore ε0E.2.vL = -LI , the electric field is downwards and of strength E = I/(2ε0v ). 

Since B = µ0I/2, this implies: 

B = µ0ε0vE.

But we have another equation linking the field strengths of the electric and magnetic fields,
Maxwell's third equation: 

We can apply this equation to a rectangular contour with sides parallel to the E field, one side
being within vt of the current sheet, the other more distant, so the only contribution to the
integral is EL from the first side, which we take to have length L.  (This contour is all on one
side of the current sheet.) The area of the rectangle the magnetic flux is passing through will
be increasing at a rate Lv (square meters per second) as the magnetic field spreads outwards. 

It follows that 

E = vB.
Putting this together with the result of the fourth equation, 

B = µ0ε0vE, 

we deduce

v2 = 1/µ0ε0 

Substituting the defined value of µ0, and the experimentally measured value of ε0, we find
that the electric and magnetic fields spread outwards from the switched-on current sheet at a
speed of 3 x 108 meters per second. 

To understand how this relates to wave propagation, imagine now that shortly after the
current is switched on, the value of the current is suddenly doubled.  Repeating the argument
above for this more complicated situation, we find the following scenario:

The electric and magnetic field propagating to the right from a current sheet in the plane 0xy,
current in the negative x-direction, suddenly switched on, then the value of the current is
suddenly doubled.

Magnetic Field

Electric Field

x
We could have ramped up the field in a series of steps — and the profile of the
magnetic and electric fields would, effectively, be a graph of how the current built up over
time.

The next step is to imagine an electric current in the sheet that’s oscillating like a sine wave
as a function of time: the magnetic and electric fields will evidently be sine waves too!   In
fact, this is how electromagnetic waves are generated.  Of course, there’s no such thing as an
infinite current sheet, an antenna has an oscillating current going up and down a wire.  But
the mechanism is essentially the same:  the only difference is that the geometry of the waves
is complicated.  Far away, they’ll look like expanding spheres, a three-dimensional version of
the ripples on a pond when a stone falls in, instead of propagating planes.  But at large
distances a small fraction of these expanding spheres w, and that looks like a series of
planes.  The picture above of how the electric and magnetic fields relate to each other and to
the direction of propagation of the wave is correct.

This is how Maxwell discovered a speed equal to the speed of light from a purely theoretical
argument based on experimental determinations of forces between currents in wires and
forces between electrostatic charges. This of course led to the realization that light is an
electromagnetic wave, and that there must be other such waves with different wavelengths.
Hertz detected other waves, of much longer wavelengths, experimentally, and this led
directly to radio, tv, radar, cellphones,  etc. 

You might also like