Maxwell's Equations
Maxwell's Equations
Maxwell's Equations
Maxwell's equations, or Maxwell–Heaviside equations, are a set of coupled partial differential equations that, together
with the Lorentz force law, form the foundation of classical electromagnetism, classical optics, electric and magnetic
circuits. The equations provide a mathematical model for electric, optical, and radio technologies, such as power
generation, electric motors, wireless communication, lenses, radar, etc. They describe how electric and magnetic fields
are generated by charges, currents, and changes of the fields.[note 1] The equations are named after the physicist and
mathematician James Clerk Maxwell, who, in 1861 and 1862, published an early form of the equations that included
the Lorentz force law. Maxwell first used the equations to propose that light is an electromagnetic phenomenon. The
modern form of the equations in their most common formulation is credited to Oliver Heaviside.[1]
Maxwell's equations may be combined to demonstrate how fluctuations in electromagnetic fields (waves) propagate at
a constant speed in vacuum, c (299 792 458 m/s).[2] Known as electromagnetic radiation, these waves occur at various
wavelengths to produce a spectrum of radiation from radio waves to gamma rays.
The publication of the equations marked the unification of a theory for previously separately described phenomena:
magnetism, electricity, light, and associated radiation. Since the mid-20th century, it has been understood that
Maxwell's equations do not give an exact description of electromagnetic phenomena, but are instead a classical limit
of the more precise theory of quantum electrodynamics.
Conceptual descriptions
Gauss's law
Gauss's law describes the relationship between an electric field and electric charges: an electric field points away from
positive charges and towards negative charges, and the net outflow of the electric field through a closed surface is
proportional to the enclosed charge, including bound charge due to polarization of material. The coefficient of the
proportion is the permittivity of free space.
Gauss's law for magnetism
Gauss's law for magnetism states that electric charges have no magnetic analogues, called magnetic monopoles; no
north or south magnetic poles exist in isolation.[3] Instead, the magnetic field of a material is attributed to a dipole, and
the net outflow of the magnetic field through a closed surface is zero. Magnetic dipoles may be represented as loops
of current or inseparable pairs of equal and opposite "magnetic charges". Precisely, the total magnetic flux through a
Gaussian surface is zero, and the magnetic field is a solenoidal vector field.[note 3]
Faraday's law
The Maxwell–Faraday version of Faraday's law of induction describes how a time-varying magnetic field corresponds
to curl of an electric field.[3] In integral form, it states that the work per unit charge required to move a charge around a
closed loop equals the rate of change of the magnetic flux through the enclosed surface.
The electromagnetic induction is the operating principle behind many electric generators: for example, a rotating bar
magnet creates a changing magnetic field and generates an electric field in a nearby wire.
The original law of Ampère states that magnetic fields relate to electric current. Maxwell's addition states that
magnetic fields also relate to changing electric fields, which Maxwell called displacement current. The integral form
states that electric and displacement currents are associated with a proportional magnetic field along any enclosing
curve.
Maxwell's addition to Ampère's law is important because the laws of Ampère and Gauss must otherwise be adjusted
for static fields.[4] As a consequence, it predicts that a rotating magnetic field occurs with a changing electric field.[3][5]
A further consequence is the existence of self-sustaining electromagnetic waves which travel through empty space.
The speed calculated for electromagnetic waves, which could be predicted from experiments on charges and
currents,[note 4] 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 1861, thereby
unifying the theories of electromagnetism and optics.
The differential and integral formulations are mathematically equivalent; both are useful. The integral formulation
relates fields within a region of space to fields on the boundary and can often be used to simplify and directly calculate
fields from symmetric distributions of charges and currents. On the other hand, the differential equations are purely
local and are a more natural starting point for calculating the fields in more complicated (less symmetric) situations,
for example using finite element analysis.[8]
Integral equations
In the integral equations,
Gauss's law
Maxwell–Faraday equation
(Faraday's law of induction)
Gauss's law
Maxwell–Faraday equation
(Faraday's law of induction)
The equations simplify slightly when a system of quantities is chosen in the speed of light, c, is used for
nondimensionalization, so that, for example, seconds and lightseconds are interchangeable, and c = 1.
Further changes are possible by absorbing factors of 4π. This process, called rationalization, affects whether
Coulomb's law or Gauss's law includes such a factor (see Heaviside–Lorentz units, used mainly in particle physics).
According to the (purely mathematical) Gauss divergence theorem, the electric flux through the boundary surface ∂Ω
can be rewritten as
By the Kelvin–Stokes theorem we can rewrite the line integrals of the fields around the closed boundary curve ∂Σ to an
integral of the "circulation of the fields" (i.e. their curls) over a surface it bounds, i.e.
Charge conservation
The invariance of charge can be derived as a corollary of Maxwell's equations. The left-hand side of the modified
Ampere's law has zero divergence by the div–curl identity. Expanding the divergence of the right-hand side,
interchanging derivatives, and applying Gauss's law gives:
i.e.,
In a region with no charges (ρ = 0) and no currents (J = 0), such as in a vacuum, Maxwell's equations reduce to:
Taking the curl ( ∇×) of the curl equations, and using the curl of the curl identity we obtain
The quantity has the dimension of (time/length)2. Defining , the equations above have the form
of the standard wave equations
Already during Maxwell's lifetime, it was found that the known values for and give , then
already known to be the speed of light in free space. This led him to propose that light and radio waves were
propagating electromagnetic waves, since amply confirmed. In the old SI system of units, the values of
and are defined constants, (which means that by definition
) that define the ampere and the metre. In the new SI system, only c keeps its defined
value, and the electron charge gets a defined value.
In materials with relative permittivity, εr, and relative permeability, μr, the phase velocity of light becomes
In addition, E and B are perpendicular to each other and to the direction of wave propagation, and are in phase with
each other. A sinusoidal plane wave is one special solution of these equations. Maxwell's equations explain how these
waves can physically propagate through space. The changing magnetic field creates a changing electric field through
Faraday's law. In turn, that electric field creates a changing magnetic field through Maxwell's addition to Ampère's law.
This perpetual cycle allows these waves, now known as electromagnetic radiation, to move through space at velocity c.
Macroscopic formulation
The above equations are the microscopic version of Maxwell's equations, expressing the electric and the magnetic
fields in terms of the (possibly atomic-level) charges and currents present. This is sometimes called the "general" form,
but the macroscopic version below is equally general, the difference being one of bookkeeping.
The microscopic version is sometimes called "Maxwell's equations in a vacuum": this refers to the fact that the material
medium is not built into the structure of the equations, but appears only in the charge and current terms. The
microscopic version was introduced by Lorentz, who tried to use it to derive the macroscopic properties of bulk matter
from its microscopic constituents.[12]: 5
"Maxwell's macroscopic equations", also known as Maxwell's equations in matter, are more similar to those that
Maxwell introduced himself.
Differential
Integral equations Differential equations
Name equations
(SI convention) (Gaussian convention)
(SI convention)
Gauss's law
Maxwell–Faraday equation
(Faraday's law of induction)
In the macroscopic equations, the influence of bound charge Qb and bound current Ib is incorporated into the
displacement field D and the magnetizing field H, while the equations depend only on the free charges Qf and free
currents If. This reflects a splitting of the total electric charge Q and current I (and their densities ρ and J) into free and
bound parts:
The cost of this splitting is that the additional fields D and H need to be determined through phenomenological
constituent equations relating these fields to the electric field E and the magnetic field B, together with the bound
charge and current.
See below for a detailed description of the differences between the microscopic equations, dealing with total charge
and current including material contributions, useful in air/vacuum;[note 6] and the macroscopic equations, dealing with
free charge and current, practical to use within materials.
When an electric field is applied to a dielectric material its molecules respond by forming microscopic electric dipoles
– their atomic nuclei move a tiny distance in the direction of the field, while their electrons move a tiny distance in the
opposite direction. This produces a macroscopic bound charge in the material even though all of the charges involved
are bound to individual molecules. For example, if every molecule responds the same, similar to that shown in the
figure, these tiny movements of charge combine to produce a layer of positive bound charge on one side of the
material and a layer of negative charge on the other side. The bound charge is most conveniently described in terms of
the polarization P of the material, its dipole moment per unit volume. If P is uniform, a macroscopic separation of
charge is produced only at the surfaces where P enters and leaves the material. For non-uniform P, a charge is also
produced in the bulk.[13]
Somewhat similarly, in all materials the constituent atoms exhibit magnetic moments that are intrinsically linked to the
angular momentum of the components of the atoms, 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 charge is traveling a large distance. These bound currents can be described using the
magnetization M.[14]
The very complicated and granular bound charges and bound currents, therefore, can be represented on the
macroscopic scale in terms of P and M, which average these charges and currents on a sufficiently large scale so as
not to see the granularity of individual atoms, but also sufficiently small that they vary with location in the material. As
such, Maxwell's macroscopic equations ignore many details on a fine scale that can be unimportant to understanding
matters on a gross scale by calculating fields that are averaged over some suitable volume.
where P is the polarization field and M is the magnetization field, which are defined in terms of microscopic bound
charges and bound currents respectively. The macroscopic bound charge density ρb and bound current density Jb in
terms of polarization P and magnetization M are then defined as
If we define the total, bound, and free charge and current density by
Constitutive relations
In order to apply 'Maxwell's macroscopic equations', it is necessary to specify the relations between displacement field
D and the electric field E, as well as the magnetizing field H and the magnetic field B. Equivalently, we have to specify
the dependence of the polarization P (hence the bound charge) and the magnetization M (hence the bound current) on
the applied electric and magnetic field. The equations specifying this response are called constitutive relations. For
real-world materials, the constitutive relations are rarely simple, except approximately, and usually determined by
experiment. See the main article on constitutive relations for a fuller description.[15]: 44–45
For materials without polarization and magnetization, the constitutive relations are (by definition)[9]: 2
In applications one also has to describe how the free currents and charge density behave in terms of E and B possibly
coupled to other physical quantities like pressure, and the mass, number density, and velocity of charge-carrying
particles. E.g., the original equations given by Maxwell (see History of Maxwell's equations) included Ohm's law in the
form
Alternative formulations
Following is a summary of some of the numerous other mathematical formalisms to write the microscopic Maxwell's
equations, with the columns separating the two homogeneous Maxwell equations from the two inhomogeneous ones
involving charge and current. Each formulation has versions directly in terms of the electric and magnetic fields, and
indirectly in terms of the electrical potential φ and the vector potential A. Potentials were introduced as a convenient
way to solve the homogeneous equations, but it was thought that all observable physics was contained in the electric
and magnetic fields (or relativistically, the Faraday tensor). The potentials play a central role in quantum mechanics,
however, and act quantum mechanically with observable consequences even when the electric and magnetic fields
vanish (Aharonov–Bohm effect).
Each table describes one formalism. See the main article for details of each formulation. SI units are used throughout.
Vector calculus
Formulation Homogeneous equations Inhomogeneous equations
Fields
Fields
space + time
Potentials
Fields
Relativistic formulations
The Maxwell equations can also be formulated on a spacetime-like Minkowski space where space and time are treated
on equal footing. The direct spacetime formulations make manifest that the Maxwell equations are relativistically
invariant. Because of this symmetry, the electric and magnetic fields are treated on equal footing and are recognized as
components of the Faraday tensor. This reduces the four Maxwell equations to two, which simplifies the equations,
although we can no longer use the familiar vector formulation. In fact the Maxwell equations in the space + time
formulation are not Galileo invariant and have Lorentz invariance as a hidden symmetry. This was a major source of
inspiration for the development of relativity theory. Indeed, even the formulation that treats space and time separately
is not a non-relativistic approximation and describes the same physics by simply renaming variables. For this reason
the relativistic invariant equations are usually called the Maxwell equations as well.
Fields
Minkowski space
Fields
any spacetime
Differential forms
Formulation Homogeneous equations Inhomogeneous equations
Fields
any spacetime
Solutions
Maxwell's equations are partial differential equations that relate the electric and magnetic fields to each other and to
the electric charges and currents. Often, the charges and currents are themselves dependent on the electric and
magnetic fields via the Lorentz force equation and the constitutive relations. These all form a set of coupled partial
differential equations which are often very difficult to solve: the solutions encompass all the diverse phenomena of
classical electromagnetism. Some general remarks follow.
As for any differential equation, boundary conditions[19][20][21] and initial conditions[22] are necessary for a unique
solution. For example, even with no charges and no currents anywhere in spacetime, there are the obvious solutions for
which E and B are zero or constant, but there are also non-trivial solutions corresponding to electromagnetic waves. In
some cases, Maxwell's equations are solved over the whole of space, and boundary conditions are given as asymptotic
limits at infinity.[23] In other cases, Maxwell's equations are solved in a finite region of space, with appropriate
conditions on the boundary of that region, for example an artificial absorbing boundary representing the rest of the
universe,[24][25] or periodic boundary conditions, or walls that isolate a small region from the outside world (as with a
waveguide or cavity resonator).[26]
Jefimenko's equations (or the closely related Liénard–Wiechert potentials) are the explicit solution to Maxwell's
equations for the electric and magnetic fields created by any given distribution of charges and currents. It assumes
specific initial conditions to obtain the so-called "retarded solution", where the only fields present are the ones created
by the charges. However, Jefimenko's equations are unhelpful in situations when the charges and currents are
themselves affected by the fields they create.
Numerical methods for differential equations can be used to compute approximate solutions of Maxwell's equations
when exact solutions are impossible. These include the finite element method and finite-difference time-domain
method.[19][21][27][28][29] For more details, see Computational electromagnetics.
Overdetermination of Maxwell's
equations
Maxwell's equations seem overdetermined, in that they involve six unknowns (the three components of E and B) but
eight equations (one for each of the two Gauss's laws, three vector components each for Faraday's and Ampere's
laws). (The currents and charges are not unknowns, being freely specifiable subject to charge conservation.) This is
related to a certain limited kind of redundancy in Maxwell's equations: It can be proven that any system satisfying
Faraday's law and Ampere's law automatically also satisfies the two Gauss's laws, as long as the system's initial
condition does, and assuming conservation of charge and the nonexistence of magnetic monopoles.[30][31] This
explanation was first introduced by Julius Adams Stratton in 1941.[32]
Although it is possible to simply ignore the two Gauss's laws in a numerical algorithm (apart from the initial
conditions), the imperfect precision of the calculations can lead to ever-increasing violations of those laws. By
introducing dummy variables characterizing these violations, the four equations become not overdetermined after all.
The resulting formulation can lead to more accurate algorithms that take all four laws into account.[33]
Both identities , which reduce eight equations to six independent ones, are the true
[34][35]
reason of overdetermination.
Equivalently, the overdetermination can be viewed as implying conservation of electric and magnetic charge, as they
are required in the derivation described above but implied by the two Gauss's laws.
For linear algebraic equations, one can make 'nice' rules to rewrite the equations and unknowns. The equations can be
linearly dependent. But in differential equations, and especially partial differential equations (PDEs), one needs
appropriate boundary conditions, which depend in not so obvious ways on the equations. Even more, if one rewrites
them in terms of vector and scalar potential, then the equations are underdetermined because of gauge fixing.
Variations
Popular variations on the Maxwell equations as a classical theory of electromagnetic fields are relatively scarce
because the standard equations have stood the test of time remarkably well.
Magnetic monopoles
Maxwell's equations posit that there is electric charge, but no magnetic charge (also called magnetic monopoles), in
the universe. Indeed, magnetic charge has never been observed, despite extensive searches,[note 7] and may not exist. If
they did exist, both Gauss's law for magnetism and Faraday's law would need to be modified, and the resulting four
equations would be fully symmetric under the interchange of electric and magnetic fields.[9]: 273–275
See also
Explanatory notes
References
Further reading
Other