Classical Electromagnetism

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 5

Classical electromagnetism

Classical electromagnetism or classical electrodynamics is a branch of theoretical physics that studies the
interactions between electric charges and currents using an extension of the classical Newtonian model. The
theory provides a description of electromagnetic phenomena whenever the relevant length scales and field
strengths are large enough that quantum mechanical effects are negligible. For small distances and low field
strengths, such interactions are better described by quantum electrodynamics.

Fundamental physical aspects of classical electrodynamics are presented in many texts, such as those by
Feynman, Leighton and Sands,[1] Griffiths,[2] Panofsky and Phillips,[3] and Jackson.[4]

Contents
History
Lorentz force
The electric field E
Electromagnetic waves
General field equations
Models
See also
References

History
The physical phenomena that electromagnetism describes have been studied as separate fields since antiquity.
For example, there were many advances in the field of optics centuries before light was understood to be an
electromagnetic wave. However, the theory of electromagnetism, as it is currently understood, grew out of
Michael Faraday's experiments suggesting an electromagnetic field and James Clerk Maxwell's use of
differential equations to describe it in his A Treatise on Electricity and Magnetism (1873). For a detailed
historical account, consult Pauli,[5] Whittaker,[6] Pais,[7] and Hunt.[8]

Lorentz force
The electromagnetic field exerts the following force (often called the Lorentz force) on charged particles:

where all boldfaced quantities are vectors: F is the force that a particle with charge q experiences, E is the
electric field at the location of the particle, v is the velocity of the particle, B is the magnetic field at the
location of the particle.

The above equation illustrates that the Lorentz force is the sum of two vectors. One is the cross product of the
velocity and magnetic field vectors. Based on the properties of the cross product, this produces a vector that is
perpendicular to both the velocity and magnetic field vectors. The other vector is in the same direction as the
electric field. The sum of these two vectors is the Lorentz force.

Therefore, in the absence of a magnetic field, the force is in the direction of the electric field, and the
magnitude of the force is dependent on the value of the charge and the intensity of the electric field. In the
absence of an electric field, the force is perpendicular to the velocity of the particle and the direction of the
magnetic field. If both electric and magnetic fields are present, the Lorentz force is the sum of both of these
vectors.

Although the equation appears to suggest that the electric and magnetic fields are independent, the equation
can be rewritten in term of four-current (instead of charge) and a single tensor that represents the combined
electromagnetic field ( )

The electric field E


The electric field E is defined such that, on a stationary charge:

where q0 is what is known as a test charge and F is the force on that charge. The size of the charge doesn't
really matter, as long as it is small enough not to influence the electric field by its mere presence. What is plain
from this definition, though, is that the unit of E is N/C (newtons per coulomb). This unit is equal to V/m
(volts per meter); see below.

In electrostatics, where charges are not moving, around a distribution of point charges, the forces determined
from Coulomb's law may be summed. The result after dividing by q0 is:

where n is the number of charges, qi is the amount of charge associated with the ith charge, ri is the position of
the ith charge, r is the position where the electric field is being determined, and ε0 is the electric constant.

If the field is instead produced by a continuous distribution of charge, the summation becomes an integral:

where is the charge density and is the vector that points from the volume element to the
point in space where E is being determined.

Both of the above equations are cumbersome, especially if one wants to determine E as a function of position.
A scalar function called the electric potential can help. Electric potential, also called voltage (the units for
which are the volt), is defined by the line integral

where φ(r) is the electric potential, and C is the path over which the integral is being taken.
Unfortunately, this definition has a caveat. From Maxwell's equations, it is clear that ∇ × E is not always zero,
and hence the scalar potential alone is insufficient to define the electric field exactly. As a result, one must add
a correction factor, which is generally done by subtracting the time derivative of the A vector potential
described below. Whenever the charges are quasistatic, however, this condition will be essentially met.

From the definition of charge, one can easily show that the electric potential of a point charge as a function of
position is:

where q is the point charge's charge, r is the position at which the potential is being determined, and ri is the
position of each point charge. The potential for a continuous distribution of charge is:

where is the charge density, and is the distance from the volume element to point in space
where φ is being determined.

The scalar φ will add to other potentials as a scalar. This makes it relatively easy to break complex problems
down in to simple parts and add their potentials. Taking the definition of φ backwards, we see that the electric
field is just the negative gradient (the del operator) of the potential. Or:

From this formula it is clear that E can be expressed in V/m (volts per meter).

Electromagnetic waves
A changing electromagnetic field propagates away from its origin in the form of a wave. These waves travel in
vacuum at the speed of light and exist in a wide spectrum of wavelengths. Examples of the dynamic fields of
electromagnetic radiation (in order of increasing frequency): radio waves, microwaves, light (infrared, visible
light and ultraviolet), x-rays and gamma rays. In the field of particle physics this electromagnetic radiation is
the manifestation of the electromagnetic interaction between charged particles.

General field equations


As simple and satisfying as Coulomb's equation may be, it is not entirely correct in the context of classical
electromagnetism. Problems arise because changes in charge distributions require a non-zero amount of time to
be "felt" elsewhere (required by special relativity).

For the fields of general charge distributions, the retarded potentials can be computed and differentiated
accordingly to yield Jefimenko's equations.

Retarded potentials can also be derived for point charges, and the equations are known as the Liénard–
Wiechert potentials. The scalar potential is:
where q is the point charge's charge and r is the position. rq and vq are the position and velocity of the charge,
respectively, as a function of retarded time. The vector potential is similar:

These can then be differentiated accordingly to obtain the complete field equations for a moving point particle.

Models
Branches of classical electromagnetism such as optics, electrical and electronic engineering consist of a
collection of relevant mathematical models of different degrees of simplification and idealization to enhance
the understanding of specific electrodynamics phenomena, cf.[9] An electrodynamics phenomenon is
determined by the particular fields, specific densities of electric charges and currents, and the particular
transmission medium. Since there are infinitely many of them, in modeling there is a need for some typical,
representative

(a) electrical charges and currents, e.g. moving pointlike charges and electric and magnetic
dipoles, electric currents in a conductor etc.;
(b) electromagnetic fields, e.g. voltages, the Liénard–Wiechert potentials, the monochromatic
plane waves, optical rays; radio waves, microwaves, infrared radiation, visible light,
ultraviolet radiation, X-rays, gamma rays etc.;
(c) transmission media, e.g. electronic components, antennas, electromagnetic waveguides,
flat mirrors, mirrors with curved surfaces convex lenses, concave lenses; resistors, inductors,
capacitors, switches; wires, electric and optical cables, transmission lines, integrated circuits
etc.;

all of which have only few variable characteristics. It worth mentioning that the exact representation of the
electromagnetic field is used in the analysis and design of antennas.

See also
Electromagnetism
Maxwell's equations
Weber electrodynamics
Wheeler–Feynman absorber theory
Leontovich boundary condition

References
1. Feynman, R. P., R .B. Leighton, and M. Sands, 1965, The Feynman Lectures on Physics, Vol.
II: the Electromagnetic Field, Addison-Wesley, Reading, Massachusetts
2. Griffiths, David J. (2013). Introduction to Electrodynamics (4th ed.). Boston, Mas.: Pearson.
ISBN 978-0321856562.
3. Panofsky, W. K., and M. Phillips, 1969, Classical Electricity and Magnetism, 2nd edition,
Addison-Wesley, Reading, Massachusetts
4. Jackson, John D. (1998). Classical Electrodynamics (3rd ed.). New York: Wiley. ISBN 978-0-
471-30932-1.
5. Pauli, W., 1958, Theory of Relativity, Pergamon, London
6. Whittaker, E. T., 1960, History of the Theories of the Aether and Electricity, Harper Torchbooks,
New York.
7. Pais, A., 1983, Subtle is the Lord: The Science and the Life of Albert Einstein, Oxford University
Press, Oxford
8. Bruce J. Hunt (1991) The Maxwellians
9. Peierls, Rudolf. Model-making in physics, Contemporary Physics, Volume 21 (1), January
1980, 3-17.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Classical_electromagnetism&oldid=993035343"

This page was last edited on 8 December 2020, at 13:10 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like