electromagnetism

electromagnetism
/i lek'troh mag"ni tiz'euhm/, n.
1. the phenomena associated with electric and magnetic fields and their interactions with each other and with electric charges and currents.
2. Also, electromagnetics. the science that deals with these phenomena.
[1820-30; ELECTRO- + MAGNETISM]

* * *

Branch of physics that deals with the relationship between electricity and magnetism.

Their merger into one concept is tied to three historical events. Hans C. Orsted's accidental discovery in 1820 that magnetic fields are produced by electric currents spurred efforts to prove that magnetic fields can induce currents. Michael Faraday showed in 1831 that a changing magnetic field can induce a current in a circuit, and James Clerk Maxwell predicted that a changing electric field has an associated magnetic field. The technological revolution attributed to the development of electric power and modern communications can be traced to these three landmarks.

* * *

Introduction

      science of charge and of the forces and fields associated with charge. electricity and magnetism are two aspects of electromagnetism.

      Electricity and magnetism were long thought to be separate forces. It was not until the 19th century that they were finally treated as interrelated phenomena. In 1905 Albert Einstein's special theory of relativity established beyond a doubt that both are aspects of one common phenomenon. At a practical level, however, electric and magnetic forces behave quite differently and are described by different equations. Electric forces (Coulomb force) are produced by electric charges either at rest or in motion. Magnetic forces, on the other hand, are produced only by moving charges and act solely on charges in motion.

      Electric phenomena occur even in neutral matter because the forces act on the individual charged constituents. The electric force, in particular, is responsible for most of the physical and chemical properties of atoms and molecules. It is enormously strong compared with gravity (gravitation). For example, the absence of only one electron out of every billion molecules in two 70-kilogram (154-pound) persons standing two metres (two yards) apart would repel them with a 30,000-ton force. On a more familiar scale, electric phenomena are responsible for the lightning and thunder accompanying certain storms.

      Electric and magnetic forces can be detected in regions called electric and magnetic fields. These fields are fundamental in nature and can exist in space far from the charge or current that generated them. Remarkably, electric fields can produce magnetic fields and vice versa, independent of any external charge. A changing magnetic field produces an electric field, as the English physicist Michael Faraday discovered in work that forms the basis of electric power generation. Conversely, a changing electric field produces a magnetic field, as the Scottish physicist James Clerk Maxwell deduced. The mathematical equations formulated by Maxwell incorporated light and wave phenomena into electromagnetism. He showed that electric and magnetic fields travel together through space as waves of electromagnetic radiation, with the changing fields mutually sustaining each other. Examples of electromagnetic waves traveling through space independent of matter are radio and television waves, microwaves, infrared rays, visible light, ultraviolet light, X rays, and gamma rays. All of these waves travel at the same speed—namely, the velocity of light (roughly 300,000 kilometres, or 186,000 miles, per second). They differ from each other only in the frequency at which their electric and magnetic fields oscillate.

      Maxwell's equations still provide a complete and elegant description of electromagnetism down to, but not including, the subatomic scale. The interpretation of his work, however, was broadened in the 20th century. Einstein's special relativity theory merged electric and magnetic fields into one common field and limited the velocity of all matter to the velocity of electromagnetic radiation. During the late 1960s, physicists discovered that other forces in nature have fields with a mathematical structure similar to that of the electromagnetic field. These other forces are the nuclear force, responsible for the energy released in nuclear fusion, and the weak force, observed in the radioactive decay of unstable atomic nuclei. In particular, the weak and electromagnetic forces have been combined into a common force called the electroweak force. The goal of many physicists to unite all of the fundamental forces, including gravity, into one grand unified theory has not been attained to date.

      An important aspect of electromagnetism is the science of electricity, which is concerned with the behaviour of aggregates of charge, including the distribution of charge within matter and the motion of charge from place to place. Different types of materials are classified as either conductors or insulators on the basis of whether charges can move freely through their constituent matter. Electric current is the measure of the flow of charges; the laws governing currents in matter are important in technology, particularly in the production, distribution, and control of energy.

      The concept of voltage, like those of charge and current, is fundamental to the science of electricity. Voltage is a measure of the propensity of charge to flow from one place to another; positive charges generally tend to move from a region of high voltage to a region of lower voltage. A common problem in electricity is determining the relationship between voltage and current or charge in a given physical situation.

      This article seeks to provide a qualitative understanding of electromagnetism as well as a quantitative appreciation for the magnitudes associated with electromagnetic phenomena.

Fundamentals
      Everyday modern life is pervaded by electromagnetic phenomena. When a light bulb (lightbulb) is switched on, a current flows through a thin filament in the bulb; the current heats the filament to such a high temperature that it glows, illuminating its surroundings. Electric clocks and connections link simple devices of this kind into complex systems such as traffic lights that are timed and synchronized with the speed of vehicular flow. Radio and television sets receive information carried by electromagnetic waves traveling through space at the speed of light. To start an automobile, currents in an electric starter motor generate magnetic fields that rotate the motor shaft and drive engine pistons to compress an explosive mixture of gasoline and air; the spark initiating the combustion is an electric discharge, which makes up a momentary current flow.

Coulomb's law
      Many of these devices and phenomena are complex, but they derive from the same fundamental laws of electromagnetism. One of the most important of these is Coulomb's law, which describes the electric force (Coulomb force) between charged objects. Formulated by the 18th-century French physicist Charles-Augustin de Coulomb (Coulomb, Charles-Augustin de), it is analogous to Newton's law (Newton's law of gravitation) for the gravitational (gravitation) force. Both gravitational and electric forces decrease with the square of the distance between the objects, and both forces act along a line between them. In Coulomb's law, however, the magnitude and sign of the electric (electric charge) force are determined by the charge, rather than the mass, of an object. Thus, charge determines how electromagnetism influences the motion of charged objects. (Charge is a basic property of matter. Every constituent of matter has an electric charge with a value that can be positive, negative, or zero. For example, electrons are negatively charged, and atomic nuclei are positively charged. Most bulk matter has an equal amount of positive and negative charge and thus has zero net charge.)

      According to Coulomb, the electric force for charges at rest has the following properties:

      (1) Like charges repel each other; unlike charges attract. Thus, two negative charges repel one another, while a positive charge attracts a negative charge.

      (2) The attraction or repulsion acts along the line between the two charges.

      (3) The size of the force varies inversely as the square of the distance between the two charges. Therefore, if the distance between the two charges is doubled, the attraction or repulsion becomes weaker, decreasing to one-fourth of the original value. If the charges come 10 times closer, the size of the force increases by a factor of 100.

      (4) The size of the force is proportional to the value of each charge. The unit used to measure charge is the coulomb (C). If there were two positive charges, one of 0.1 coulomb and the second of 0.2 coulomb, they would repel each other with a force that depends on the product 0.2 × 0.1. If each of the charges were reduced by one-half, the repulsion would be reduced to one-quarter of its former value.

      Static cling is a practical example of the Coulomb force. In static cling, garments made of synthetic material collect a charge, especially in dry winter air. A plastic or rubber comb passed quickly through hair also becomes charged and will pick up bits of paper. The synthetic fabric and the comb are insulators; charge on these objects cannot move easily from one part of the object to another. Similarly, an office copy machine uses electric force to attract particles of ink to paper.

Principle of charge conservation
      Like Coulomb's law, the principle of charge conservation is a fundamental law of nature. According to this principle, the charge of an isolated system cannot change. If an additional positively charged particle appears within a system, a particle with a negative charge of the same magnitude will be created at the same time; thus, the principle of conservation of charge is maintained. In nature, a pair of oppositely charged particles is created when high-energy radiation interacts with matter; an electron and a positron are created in a process known as pair production.

      The smallest subdivision of the amount of charge that a particle can have is the charge of one proton, +1.602 × 10−19 coulomb. The electron has a charge of the same magnitude but opposite sign—i.e., −1.602 × 10−19 coulomb. An ordinary flashlight battery delivers a current that provides a total charge flow of approximately 5,000 coulomb, which corresponds to more than 1022 electrons, before it is exhausted.

       electric current is a measure of the flow of charge, as, for example, charge flowing through a wire. The size of the current is measured in amperes (ampere) and symbolized by i. An ampere of current represents the passage of one coulomb of charge per second, or 6.2 billion billion electrons (6.2 × 1018 electrons) per second. A current is positive when it is in the direction of the flow of positive charges; its direction is opposite to the flow of negative charges.

Electric fields and forces
      The force and conservation laws are only two aspects of electromagnetism, however. Electric and magnetic forces are caused by electromagnetic fields. The term field denotes a property of space, so that the field quantity has a numerical value at each point of space. These values may also vary with time. The value of the electric or magnetic field is a vector—i.e., a quantity having both magnitude and direction. The value of the electric field at a point in space, for example, equals the force that would be exerted on a unit charge at that position in space.

      Every charged object sets up an electric field in the surrounding space. A second charge “feels” the presence of this field. The second charge is either attracted toward the initial charge or repelled from it, depending on the signs of the charges. Of course, since the second charge also has an electric field, the first charge feels its presence and is either attracted or repelled by the second charge, too.

 The electric field from a charge is directed away from the charge when the charge is positive and toward the charge when it is negative. The electric field from a charge at rest is shown in Figure 1—> for various locations in space. The arrows point in the direction of the electric field, and the length of the arrows indicates the strength of the field at the midpoint of the arrows.

      If a positive charge were placed in the electric field, it would feel a force in the direction of the field. A negative charge would feel a force in the direction opposite the direction of the field.

      In calculations, it is often more convenient to deal directly with the electric field than with the charges; frequently, more is known about the field than about the distribution of charges in space. For example, the distribution of charges in conductors is generally unknown because the charges move freely within the conductor. In static situations, however, the electric field in a conductor in equilibrium has a definite value, zero, because any force on the charges inside the conductor redistributes them until the field vanishes. The unit of electric field is newtons per coulomb, or volts per metre.

      The electric potential is another useful field. It provides an alternative to the electric field in electrostatics problems. The potential is easier to use, however, because it is a single number, a scalar, instead of a vector. The difference in potential between two places measures the degree to which charges are influenced to move from one place to another. If the potential is the same at two places (i.e., if the places have the same voltage (volt)), charges will not be influenced to move from one place to the other. The potential on an object or at some point in space is measured in volts; it equals the electrostatic energy that a unit charge would have at that position. In a typical 12-volt car battery, the battery terminal that is marked with a + sign is at a potential 12 volts greater than the potential of the terminal marked with the − sign. When a wire, such as the filament of a car headlight, is connected between the + and the − terminals of the battery, charges move through the filament as an electric current and heat the filament; the hot filament radiates light.

Magnetic fields and forces
      The magnetic force influences only those charges that are already in motion. It is transmitted by the magnetic field. Both magnetic fields and magnetic forces are more complicated than electric fields and electric forces. The magnetic field does not point along the direction of the source of the field; instead, it points in a perpendicular direction. In addition, the magnetic force acts in a direction that is perpendicular to the direction of the field. In comparison, both the electric force and the electric field point directly toward or away from the charge.

   The present discussion will deal with simple situations in which the magnetic field is produced by a current of charge in a wire. Certain materials, such as copper, silver, and aluminum, are conductors that allow charge to flow freely from place to place. If an external influence establishes a current in a conductor, the current generates a magnetic field. For a long straight wire, the magnetic field has a direction that encircles the wire on a plane perpendicular to the wire. The strength of the magnetic field decreases with distance from the wire. The arrows in Figure 2—> represent the size and direction of the magnetic field for a current moving in the direction indicated. Figure 2A—> shows an end view with the current coming toward the reader, while Figure 2B—> provides a three-dimensional view of the magnetic field at one position along the wire.

 In subsequent figures, continuous lines will be used to represent the direction of electric and magnetic fields. These lines emphasize the important fact that electric fields begin on positive charges and end on negative charges, while magnetic fields do not have beginnings or ends and close on themselves. The magnetic field shown in Figure 2—> is unusually simple. Highly complex and useful magnetic fields can be generated by the proper choice of conductors to carry electric currents. Under development are thermonuclear fusion (nuclear fusion) reactors for obtaining energy from the fusion of light nuclei in the form of very hot plasmas of hydrogen isotopes. The plasmas have to be confined by magnetic fields (dubbed “magnetic bottles”) as no material container can withstand such high temperatures. Charged particles are also confined by magnetic fields in nature. Large numbers of charged particles, mostly protons and electrons, are trapped in huge bands around the Earth by its magnetic field. These bands are known as the Van Allen radiation belts. Disturbance of the Earth's confining magnetic field produces spectacular displays, the so-called northern lights, in which trapped charged particles are freed and crash through the atmosphere to Earth.

Interaction of a magnetic field with a charge
 How does the magnetic field interact with a charged object? If the charge is at rest, there is no interaction. If the charge moves, however, it is subjected to a force, the size of which increases in direct proportion with the velocity of the charge. The force has a direction that is perpendicular both to the direction of motion of the charge and to the direction of the magnetic field. There are two possible precisely opposite directions for such a force for a given direction of motion. This apparent ambiguity is resolved by the fact that one of the two directions applies to the force on a moving positive charge while the other direction applies to the force on a moving negative charge. Figure 3—> illustrates the directions of the magnetic force on positive charges and on negative charges as they move in a magnetic field that is perpendicular to the motion.

      Depending on the initial orientation of the particle velocity to the magnetic field, charges having a constant speed in a uniform magnetic field will follow a circular or helical path.

      Electric currents in wires are not the only source of magnetic fields. Naturally occurring minerals exhibit magnetic properties and have magnetic fields. These magnetic fields result from the motion of electrons in the atoms (atom) of the material. They also result from a property of electrons called the magnetic dipole moment, which is related to the intrinsic spin of individual electrons. In most materials, little or no field is observed outside the matter because of the random orientation of the various constituent atoms. In some materials such as iron, however, atoms within certain distances tend to become aligned in one particular direction.

      Magnets have numerous applications, ranging from use as toys and paper holders on home refrigerators to essential components in electric generators and machines that can accelerate particles to speeds approaching that of light. The practical application of magnetism in technology is greatly enhanced by using iron and other ferromagnetic materials with electric currents in devices like motors. These materials amplify the magnetic field produced by the currents and thereby create more powerful fields.

      While electric and magnetic effects are well separated in many phenomena and applications, they are coupled closely together when there are rapid time fluctuations. Faraday's law of induction describes how a time-varying magnetic field produces an electric field (see below Faraday's law of induction (electromagnetism)). Important practical applications include the electric generator and transformer. In a generator, the physical motion of a magnetic field produces electricity for power. In a transformer, electric power is converted from one voltage level to another by the magnetic field of one circuit inducing an electric current in another circuit.

      The existence of electromagnetic waves depends on the interaction between electric and magnetic fields. Maxwell postulated that a time-varying electric field produces a magnetic field. His theory predicted the existence of electromagnetic waves in which each time-varying field produces the other field. For example, radio waves are generated by electronic circuits known as oscillators that cause rapidly oscillating currents to flow in antennas; the rapidly varying magnetic field has an associated varying electric field. The result is the emission of radio waves into space (see electromagnetic radiation: Generation of electromagnetic radiation (electromagnetic radiation)).

      Many electromagnetic devices can be described by circuits consisting of conductors and other elements. These circuits may operate with a steady flow of current, as in a flashlight, or with time-varying currents. Important elements in circuits include sources of power called electromotive forces; resistors, which control the flow of current for a given voltage; capacitors, which store charge and energy temporarily; and inductors, which also store electrical energy for a limited time. Circuits with these elements can be described entirely with algebra. (For more complicated circuit elements such as transistors, see semiconductor device and integrated circuit).

      Two mathematical quantities associated with vector fields, like the electric field E and the magnetic field B, are useful for describing electromagnetic phenomena. They are the flux (electric flux) of such a field through a surface and the line integral of the field along a path. The flux of a field through a surface measures how much of the field penetrates through the surface; for every small section of the surface, the flux is proportional to the area of that section and depends also on the relative orientation of the section and the field. The line integral of a field along a path measures the degree to which the field is aligned with the path; for every small section of path, it is proportional to the length of that section and is also dependent on the alignment of the field with that section of path. When the field is perpendicular to the path, there is no contribution to the line integral. The fluxes of E and B through a surface and the line integrals of these fields along a path play an important role in electromagnetic theory. As examples, the flux of the electric field E through a closed surface measures the amount of charge contained within the surface; the flux of the magnetic field B through a closed surface is always zero because there are no magnetic monopoles (magnetic charges consisting of a single pole) to act as sources of the magnetic field in the way that charge is a source of the electric field.

Effects of varying magnetic fields
      The merger of electricity and magnetism from distinct phenomena into electromagnetism is tied to three closely related events. The first was Hans Christian Ørsted's accidental discovery of the influence of an electric current on a magnetic needle—namely, that magnetic fields are produced by electric currents. Ørsted's 1820 report of his observation spurred an intense effort by scientists to prove that magnetic fields can induce currents. The second event was Faraday's (Faraday, Michael) experimental proof that a changing magnetic field can induce a current in a circuit. The third was Maxwell's prediction that a changing electric field has an associated magnetic field. The technological revolution attributed to the development of electric power and radio communications can be traced to these three landmarks (see below).

Faraday's law of induction
      Faraday's discovery in 1831 of the phenomenon of magnetic induction is one of the great milestones in the quest toward understanding and exploiting nature. Stated simply, Faraday found that (1) a changing magnetic field in a circuit induces an electromotive force in the circuit; and (2) the magnitude of the electromotive force equals the rate at which the flux of the magnetic field through the circuit changes. The flux is a measure of how much field penetrates through the circuit. The electromotive force is measured in volts and is represented by the equation

      Here, Φ, the flux of the vector field B through the circuit, measures how much of the field passes through the circuit. To illustrate the meaning of flux, imagine how much water from a steady rain will pass through a circular ring of area A. When the ring is placed parallel to the path of the water drops, no water passes through the ring. The maximum rate at which drops of rain pass through the ring occurs when the surface is perpendicular to the motion of the drops. The rate of water drops crossing the surface is the flux of the vector field ρv through that surface, where ρ is the density of water drops and v represents the velocity of the water. Clearly, the angle between v and the surface is essential in determining the flux. To specify the orientation of the surface, a vector A is defined so that its magnitude is the surface area A in units of square metres and its direction is perpendicular to the surface. The rate at which raindrops pass through the surface is ρv cos θA, where θ is the angle between v and A. Using vector notation, the flux is ρv · A. For the magnetic field, the amount of flux through a small area represented by the vector dA is given by B · dA. For a circuit consisting of a single turn of wire, adding the contributions from the entire surface that is surrounded by the wire gives the magnetic flux Φ of equation (43—>). The rate of change of this flux is the induced electromotive force. The units of magnetic flux are webers (weber), with one weber equaling one tesla per square metre. Finally, the minus sign in equation (43—>) indicates the direction of the induced electromotive force and hence of any induced current. The magnetic flux through the circuit generated by the induced current is in whatever direction will keep the total flux in the circuit from changing. The minus sign in equation (43—>) is an example of Lenz's law for magnetic systems. This law, deduced by the Russian-born physicist Heinrich Friedrich Emil Lenz, states that “what happens is that which opposes any change in the system.”

 Faraday's law is valid regardless of the process that causes the magnetic flux to change. It may be that a magnet is moved closer to a circuit or that a circuit is moved closer to a magnet. Figure 4—> shows a magnet brought near a conducting ring and gives the direction of the induced current and field, thus illustrating both Faraday's and Lenz's laws. Another alternative is that the circuit may change in size in a fixed external magnetic field or, as in the case of alternating-current generation, that the circuit may be a coil of conducting wire rotating in a magnetic field so that the flux Φ varies sinusoidally in time.

      The magnetic flux Φ through a circuit has to be considered carefully in the application of Faraday's law given in equation (43—>). For example, if a circuit consists of a coil with five closely spaced turns and if ϕ is the magnetic flux through a single turn, then the value of Φ for the five-turn circuit that must be used in Faraday's law is Φ = 5ϕ. If the five turns are not the same size and closely spaced, the problem of determining Φ can be quite complex.

Self-inductance and mutual inductance
      The self-inductance of a circuit is used to describe the reaction of the circuit to a changing current in the circuit, while the mutual inductance with respect to a second circuit describes the reaction to a changing current in the second circuit. When a current i1 flows in circuit 1, i1 produces a magnetic field B1; the magnetic flux through circuit 1 due to current i1 is Φ11. Since B1 is proportional to i1, Φ11 is as well. The constant of proportionality is the self-inductance L1 of the circuit. It is defined by the equation

      As indicated earlier, the units of inductance are henrys. If a second circuit is present, some of the field B1 will pass through circuit 2 and there will be a magnetic flux Φ21 in circuit 2 due to the current i1. The mutual inductance M21 is given by

      The magnetic flux in circuit 1 due to a current in circuit 2 is given by Φ12 = M12i2. An important property of the mutual inductance is that M21 = M12. It is therefore sufficient to use the label M without subscripts for the mutual inductance of two circuits.

      The value of the mutual inductance of two circuits can range from +√L1L2 to −√L1L2, depending on the flux linkage between the circuits. If the two circuits are very far apart or if the field of one circuit provides no magnetic flux through the other circuit, the mutual inductance is zero. The maximum possible value of the mutual inductance of two circuits is approached as the two circuits produce B fields with increasingly similar spatial configurations.

      If the rate of change with respect to time is taken for the terms on both sides of equation (44—>), the result is dΦ11/dt = L1di1/dt. According to Faraday's law, dΦ11/dt is the negative of the induced electromotive force. The result is the equation frequently used for a single inductor in an AC circuit—i.e.,

      The phenomenon of self-induction was first recognized by the American scientist Joseph Henry (Henry, Joseph). He was able to generate large and spectacular electric arcs (electric arc) by interrupting the current in a large copper coil with many turns. While a steady current is flowing in a coil, the energy in the magnetic field is given by 1/2Li2. If both the inductance L and the current i are large, the amount of energy is also large. If the current is interrupted, as, for example, by opening a knife-blade switch, the current and therefore the magnetic flux through the coil drop quickly. Equation (46—>) describes the resulting electromotive force induced in the coil, and a large potential difference is developed between the two poles of the switch. The energy stored in the magnetic field of the coil is dissipated as heat and radiation in an electric arc across the space between the terminals of the switch. Due to advances in superconducting wires for electromagnets, it is possible to use large magnets with magnetic fields of several teslas for temporarily storing electric energy as energy in the magnetic field. This is done to accommodate short-term fluctuations in the consumption of electric power.

 A transformer is an example of a device that uses circuits with maximum mutual induction. Figure 5—> illustrates the configuration of a typical transformer. Here, coils of insulated conducting wire are wound around a ring of iron constructed of thin isolated laminations or sheets. The laminations minimize eddy currents (eddy current) in the iron. Eddy currents are circulatory currents induced in the metal by the changing magnetic field. These currents produce an undesirable by-product— heat in the iron. Energy loss in a transformer can be reduced by using thinner laminations, very “soft” (low-carbon) iron and wire with a larger cross section, or by winding the primary and secondary circuits with conductors that have very low resistance. Unfortunately, reducing the heat loss increases the cost of transformers. Transformers used to transmit and distribute power are commonly 98 to 99 percent efficient. While eddy currents are a problem in transformers, they are useful for heating objects in a vacuum. Eddy currents are induced in the object to be heated by surrounding a relatively nonconducting vacuum enclosure with a coil carrying a high-frequency alternating current.

      In a transformer, the iron ensures that nearly all the lines of B passing through one circuit also pass through the second circuit and that, in fact, essentially all the magnetic flux is confined to the iron. Each turn of the conducting coils has the same magnetic flux; thus, the total flux for each coil is proportional to the number of turns in the coil. As a result, if a source of sinusoidally varying electromotive force is connected to one coil, the electromotive force in the second coil is given by

      Thus, depending on the ratio of N2 to N1, the transformer can be either a step-up or a step-down device for alternating voltages. For many reasons, including safety, generation and consumption of electric power occur at relatively low voltages. Step-up transformers are used to obtain high voltages before electric power is transmitted, since for a given amount of power, the current in the transmission lines is much smaller. This minimizes energy lost by resistive heating of the conductors.

      Faraday's law constitutes the basis for the power industry and for the transformation of mechanical energy into electric energy. In 1821, a decade before his discovery of magnetic induction, Faraday conducted experiments with electric wires rotating around compass needles. This earlier work, in which a wire carrying a current rotated around a magnetized needle and a magnetic needle was made to rotate around a wire carrying an electric current, provided the groundwork for the development of the electric motor.

Effects of varying electric fields
      Maxwell's (Maxwell, James Clerk) prediction that a changing electric field generates a magnetic field was a masterstroke of pure theory. The Maxwell equations for the electromagnetic field unified all that was hitherto known about electricity and magnetism and predicted the existence of an electromagnetic phenomenon that can travel as waves with the velocity of 1/√ε0μ0 in a vacuum. That velocity, which is based on constants obtained from purely electric measurements, corresponds to the speed of light. Consequently, Maxwell concluded that light itself was an electromagnetic phenomenon. Later, Einstein's special relativity theory postulated that the value of the speed of light is independent of the motion of the source of the light. Since then, the speed of light has been measured with increasing accuracy. In 1983 it was defined to be exactly 299,792,458 metres per second. Together with the cesium clock, which has been used to define the second, the speed of light serves as the new standard for length.

  The circuit in Figure 6—> is an example of a magnetic field generated by a changing electric field. A capacitor with parallel plates is charged at a constant rate by a steady current flowing through the long, straight leads in Figure 6A—>.

 Special Comp—>   The objective is to apply Ampère's circuital law (Ampère's law) for magnetic fields to the path P, which goes around the wire in Figure 6A—>. This law (named in honour of the French physicist André-Marie Ampère) can be derived from the Biot and Savart equation for the magnetic field produced by a current (equation [34—>]). Using vector calculus notation, Ampère's law states that the integral ∮B · dl along a closed path surrounding the current i is equal to μ0i. (An integral is essentially a sum, and, in this case, ∮B · dl is the sum of B cos θdl taken for a small length of the path until the complete loop is included. At each segment of the path dl, θ is the angle between the field B and dl. ) The current i in Ampère's law is the total flux of the current density J through any surface surrounded by the closed path. In Figure 6A—>, the closed path is labeled P, and a surface S1 is surrounded by path P. All the current density through S1 lies within the conducting wire. The total flux of the current density is the current i flowing through the wire. The result for surface S1 reflects the value of the magnetic field around the wire in the region of the path P. In Figure 6B—>, path P is the same but the surface S2 passes between the two plates of the capacitor. The value of the total flux of the current density through the surface should also be i. There is, however, clearly no motion of charge at all through the surface S2. The dilemma is that the value of the integral ∮B · dl for the path P cannot be both μ0i and zero.

      Maxwell's resolution of this dilemma was his conclusion that there must be some other kind of current density, called the displacement current Jd, for which the total flux through the surface S2 would be the same as the current i through the surface S1. Jd would take, for the surface S2, the place of the current density J associated with the movement of charge, since J is clearly zero due to the lack of charges between the plates of the capacitor. What happens between the plates while the current i is flowing? Because the amount of charge on the capacitor increases with time, the electric field between the plates increases with time too. If the current stops, there is an electric field between the plates as long as the plates are charged, but there is no magnetic field around the wire. Maxwell decided that the new type of current density was associated with the changing of the electric field. He found that

      where D = ε0E and E is the electric field between the plates. In situations where matter is present, the field D in equation (48—>) is modified to include polarization effects; the result is D = ε0E + P. The field D is measured in coulombs per square metre. Adding the displacement current to Ampère's law represented Maxwell's prediction that a changing electric field also could be a source of the magnetic field B. Following Maxwell's predictions of electromagnetic waves, the German physicist Heinrich Hertz (Hertz, Heinrich) initiated the era of radio communications in 1887 by generating and detecting electromagnetic waves.

      Using vector calculus notation, the four equations of Maxwell's (Maxwell's equations) theory of electromagnetism are

      where D = ε0E + P, and H = B/μ0M. The first equation is based on Coulomb's (Coulomb's law) inverse square law for the force between two charges; it is a form of Gauss's law, which relates the flux of the electric field through a closed surface to the total charge enclosed by the surface. The second equation is based on the fact that apparently no magnetic monopoles exist in nature; if they did, they would be point sources of magnetic field. The third is a statement of Faraday's law of magnetic induction, which reveals that a changing magnetic field generates an electric field. The fourth is Ampère's law as extended by Maxwell to include the displacement current discussed above; it associates a magnetic field to a changing electric field as well as to an electric current.

      Maxwell's four equations represent a complete description of the classical theory of electromagnetism. His discovery that light is an electromagnetic wave meant that optics could be understood as part of electromagnetism. Only in microscopic situations is it necessary to modify Maxwell's equations to include quantum effects. That modification, known as quantum electrodynamics (QED), accounts for certain atomic properties to a degree of precision exceeding one part in 100 million.

      Sometimes it is necessary to shield apparatus from external electromagnetic fields. For a static electric field, this is a simple matter; the apparatus is surrounded by a shield made of a good conductor (e.g., copper). Shielding apparatus from a steady magnetic field is more difficult because materials with infinite magnetic permeability μ do not exist; for example, a hollow shield made of soft iron will reduce the magnetic field inside to a considerable extent but not completely. As discussed earlier, it is sometimes possible to superpose a field in the opposite direction to produce a very low field region and then to use additional material with a high μ for shielding. In the case of electromagnetic waves, the penetration of the waves in matter varies, depending on the frequency of the radiation and the electric conductivity of the medium. The skin depth δ (which is the distance in the conducting medium traversed for an amplitude decrease of 1/e, about 1/3) is given by

      At high frequency, the skin depth is small. Therefore, to transmit electronic messages through seawater, for example, a very low frequency must be used to get a reasonable fraction of the signal far below the surface.

      A metal shield can have some holes in it and still be effective. For instance, a typical microwave oven has a frequency of 2.5 gigahertz, which corresponds to a wavelength of about 12 centimetres for the electromagnetic wave (electromagnetic radiation) inside the oven. The metal shield on the door has small holes about two millimetres in diameter; the shield works because the wavelength of the microwave radiation is much greater than the size of the holes. On the other hand, the same shield is not effective with radiation of a much shorter wavelength. Visible light passes through the holes in the shield, as evidenced by the fact that it is possible to see inside a microwave oven when the door is closed.

Edwin Kashy Sharon Bertsch McGrayne

Historical survey
      Electric and magnetic forces have been known since antiquity, but they were regarded as separate phenomena for centuries. Magnetism was studied experimentally at least as early as the 13th century; the properties of the magnetic compass undoubtedly aroused interest in the phenomenon. Systematic investigations of electricity were delayed until the invention of practical devices for producing electric charge and currents. As soon as inexpensive, easy-to-use sources of electricity became available, scientists produced a wealth of experimental data and theoretical insights. As technology advanced, they studied, in turn, magnetism and electrostatics, electric currents and conduction, electrochemistry, magnetic and electric induction, the interrelationship between electricity and magnetism, and finally the fundamental nature of electric charge.

Early observations and applications
      The ancient (ancient Greek civilization) Greeks knew about the attractive force of both magnetite and rubbed amber. Magnetite, a magnetic oxide of iron mentioned in Greek texts as early as 800 BC, was mined in the province of Magnesia in Thessaly. Thales of Miletus, who lived nearby, may have been the first Greek to study magnetic forces. He apparently knew that magnetite attracts iron and that rubbing amber (a fossil tree resin that the Greeks called ēlektron) would make it attract such lightweight objects as feathers. According to Lucretius, the Roman author of the philosophical poem De rerum natura (“On the Nature of Things”) in the 1st century BC, the term magnet was derived from the province of Magnesia. Pliny the Elder, however, attributes it to the supposed discoverer of the mineral, the shepherd Magnes, “the nails of whose shoes and the tip of whose staff stuck fast in a magnetic field while he pastured his flocks.”

      The oldest practical application of magnetism was the magnetic compass, but its origin remains unknown. Some historians believe it was used in China as far back as the 26th century BC; others contend that it was invented by the Italians or Arabs and introduced to the Chinese during the 13th century AD. The earliest extant European reference is by Alexander Neckam (Neckam, Alexander) (d. 1217) of England.

      The first experiments with magnetism are attributed to Petrus Peregrinus de Maricourt, a French crusader and engineer. In his oft-cited Epistola de magnete (1269; “Letter on the Magnet”), Peregrinus describes having placed a thin iron rectangle on different parts of a spherically shaped piece of magnetite (or lodestone) and marked the lines along which it set itself. The lines formed a set of meridians of longitude passing through two points at opposite ends of the stone, in much the same way as the lines of longitude on the Earth's surface intersect at the North and South poles. By analogy, Peregrinus called the points the poles of the magnet. He further noted that, when a magnet is cut into pieces, each piece still has two poles. He also observed that unlike poles attract each other and that a strong magnet can reverse the polarity of a weaker one.

Emergence of the modern sciences of electricity and magnetism
      The founder of the modern sciences of electricity and magnetism was William Gilbert (Gilbert, William), physician to both Elizabeth I and James I of England. Gilbert spent 17 years experimenting with magnetism and, to a lesser extent, electricity. He assembled the results of his experiments and all of the available knowledge on magnetism in the treatise De Magnete, Magneticisque Corporibus, et de Magno Magnete Tellure (“On the Magnet, Magnetic Bodies, and the Great Magnet of the Earth”), published in 1600. As suggested by the title, Gilbert described the Earth as a huge magnet (geomagnetic field). He introduced the term electric for the force between two objects charged by friction and showed that frictional electricity occurs in many common materials. He also noted one of the primary distinctions between magnetism and electricity: the force between magnetic objects tends to align the objects relative to each other and is affected only slightly by most intervening objects, while the force between electrified objects is primarily a force of attraction or repulsion between the objects and is grossly affected by intervening matter. Gilbert attributed the electrification of a body by friction to the removal of a fluid, or “humour,” (humour) which then left an “effluvium,” or atmosphere, around the body. The language is quaint, but, if the “humour” is renamed “charge” and the “effluvium” renamed “electric field,” Gilbert's notions closely approach modern ideas.

Pioneering efforts
      During the 17th and early 18th centuries, as better sources of charge were developed, the study of electric effects became increasingly popular. The first machine to generate an electric spark was built in 1663 by Otto von Guericke (Guericke, Otto von), a German physicist and engineer. Guericke's electric generator consisted of a sulfur globe mounted on an iron shaft. The globe could be turned with one hand and rubbed with the other. Electrified by friction, the sphere alternately attracted and repulsed light objects from the floor.

      Stephen Gray, a British chemist, is credited with discovering that electricity can flow (1729). He found that corks stuck in the ends of glass tubes become electrified when the tubes are rubbed. He also transmitted electricity approximately 150 metres through a hemp thread supported by silk cords and, in another demonstration, sent electricity even farther through metal wire. Gray concluded that electricity flowed everywhere.

      From the mid-18th through the early 19th centuries, scientists believed that electricity was composed of fluid. In 1733 Charles François de Cisternay DuFay, a French chemist, announced that electricity consisted of two fluids: “vitreous” (from the Latin for “glass”), or positive, electricity; and “resinous,” or negative, electricity. When DuFay electrified a glass rod, it attracted nearby bits of cork. Yet, if the rod touched the pieces of cork, the cork fragments were repelled and also repelled one another. DuFay accounted for this phenomenon by explaining that, in general, matter was neutral because it contained equal quantities of both fluids; if, however, friction separated the fluids in a substance and left it imbalanced, the substance would attract or repel other matter.

Invention of the Leyden jar
      In 1745 a cheap and convenient source of electric sparks was invented by Pieter van Musschenbroek (Musschenbroek, Pieter van), a physicist and mathematician in Leiden, Neth. Later called the Leyden jar, it was the first device that could store large amounts of electric charge. (E. Georg von Kleist (Kleist, E Georg von), a German cleric, independently developed the idea for such a device, but did not investigate it as thoroughly as did Musschenbroek.) The Leyden jar devised by the latter consisted of a glass vial that was partially filled with water and contained a thick conducting wire capable of storing a substantial amount of charge. One end of this wire protruded through the cork that sealed the opening of the vial. The Leyden jar was charged by bringing this exposed end of the conducting wire into contact with a friction device that generated static electricity.

      Within a year after the appearance of Musschenbroek's device, William Watson, an English physician and scientist, constructed a more sophisticated version of the Leyden jar; he coated the inside and outside of the container with metal foil to improve its capacity to store charge. Watson transmitted an electric spark from his device through a wire strung across the River Thames at Westminster Bridge in 1747.

      The Leyden jar revolutionized the study of electrostatics. Soon “electricians” were earning their living all over Europe demonstrating electricity with Leyden jars. Typically, they killed birds and animals with electric shock or sent charges through wires over rivers and lakes. In 1746 the abbé Jean-Antoine Nollet, a physicist who popularized science in France, discharged a Leyden jar in front of King Louis XV by sending current through a chain of 180 Royal Guards. In another demonstration, Nollet used wire made of iron to connect a row of Carthusian monks more than a kilometre long; when a Leyden jar was discharged, the white-robed monks reportedly leapt simultaneously into the air.

      In the United States, Benjamin Franklin (Franklin, Benjamin) sold his printing house, newspaper, and almanac to spend his time conducting electricity experiments. In 1752 Franklin proved that lightning was an example of electric conduction by flying a silk kite during a thunderstorm. He collected electric charge from a cloud by means of wet twine attached to a key and thence to a Leyden jar. He then used the accumulated charge (charge conservation) from the lightning to perform electric experiments. Franklin enunciated the law now known as the conservation of charge (the net sum of the charges within an isolated region is always constant). Like Watson, he disagreed with DuFay's two-fluid theory. Franklin argued that electricity consisted of two states of one fluid, which is present in everything. A substance containing an unusually large amount of the fluid would be “plus,” or positively charged. Matter with less than a normal amount of fluid would be “minus,” or negatively charged. Franklin's one-fluid theory, which dominated the study of electricity for 100 years, is essentially correct because most currents are the result of moving electrons. At the same time, however, fundamental particles have both negative and positive charges and, in this sense, DuFay's two-fluid picture is correct.

      Joseph Priestley (Priestley, Joseph), an English physicist, summarized all available data on electricity in his book History and Present State of Electricity (1767). He repeated one of Franklin's experiments, in which the latter had dropped small corks into a highly electrified metal container and found that they were neither attracted nor repelled. The lack of any charge on the inside of the container caused Priestley to recall Newton's law that there is no gravitational force on the inside of a hollow sphere. From this, Priestley inferred that the law of force between electric charges must be the same as the law for gravitational force—i.e., that the force between masses diminishes with the inverse square of the distance between the masses. Although they were expressed in qualitative and descriptive terms, Priestley's laws are still valid today. Their mathematics was clarified and developed extensively between 1767 and the mid-19th century as electricity and magnetism became precise, quantitative sciences.

Formulation of the quantitative laws of electrostatics and magnetostatics
      Charles-Augustin de Coulomb (Coulomb, Charles-Augustin de) established electricity as a mathematical science during the latter half of the 18th century. He transformed Priestley's descriptive observations into the basic quantitative laws of electrostatics and magnetostatics. He also developed the mathematical theory of electric force and invented the torsion balance that was to be used in electricity experiments for the next 100 years. Coulomb used the balance to measure the force between magnetic poles and between electric charges at varying distances. In 1785 he announced his quantitative proof that electric and magnetic forces vary, like gravitation, inversely as the square of the distance (see above Fundamentals (electromagnetism)). Thus, according to Coulomb's law, if the distance between two charged masses is doubled, the electric force (Coulomb force) between them is reduced to a fourth. (The English physicist Henry Cavendish (Cavendish, Henry), as well as John Robison of Scotland, had made quantitative determinations of this principle before Coulomb, but they had not published their work.)

      The mathematicians Siméon-Denis Poisson (Poisson, Siméon-Denis) of France and Carl Friedrich Gauss (Gauss, Carl Friedrich) of Germany extended Coulomb's work during the 18th and early 19th centuries. Poisson's equation (published in 1813) and the law of charge conservation contain in two lines virtually all the laws of electrostatics. The theory of magnetostatics, which is the study of steady-state magnetic fields, also was developed from Coulomb's law. Magnetostatics uses the concept of a magnetic potential analogous to the electric potential (i.e., magnetic poles are postulated with properties analogous to electric charges).

      Michael Faraday built upon Priestley's work and conducted an experiment that verified quite accurately the inverse square law. Faraday's experiment involving the use of a metal ice pail and a gold-leaf electroscope was the first precise quantitative experiment on electric charge. In Faraday's time, the gold-leaf electroscope was used to indicate the electric state of a body. This type of apparatus consists of two thin leaves of gold hanging from an insulated metal rod that is mounted inside a metal box. When the rod is charged, the leaves repel each other and the deflection indicates the size of the charge. Faraday began his experiment by charging a metal ball suspended on an insulating silk thread. He then connected the gold-leaf electroscope to a metal ice pail resting on an insulating block and lowered the charged ball into the pail. The electroscope reading increased as the ball was lowered into the pail and reached a steady value once the ball was within the pail. When the ball was withdrawn without touching the pail, the electroscope reading fell to zero. Yet, when the ball touched the bottom of the pail, the reading remained at its steady value. On removal, the ball was found to be completely discharged. Faraday concluded that the electric charge produced on the outside of the pail, when the ball was inside but not in contact with it, was exactly equal to the initial charge on the ball. He then inserted into the pail other objects, such as a set of concentric pails separated from one another with various insulating materials like sulfur. In each case, the electroscope reading was the same once the ball was completely within the pail. From this, Faraday concluded that the total charge of the system was an invariable quantity equal to the initial charge of the ball. The present-day belief that conservation is a fundamental property of charge rests not only on the experiments of Franklin and Faraday but also on its complete agreement with all observations in electric engineering, quantum electrodynamics, and experimental electricity. With Faraday's work, the theory of electrostatics was complete.

Foundations of electrochemistry and electrodynamics
Development of the battery
      The invention of the battery in 1800 made possible for the first time major advances in the theories of electric current and electrochemistry. Both science and technology developed rapidly as a direct result, leading some to call the 19th century the age of electricity.

      The development of the battery was the accidental result of biological experiments conducted by Luigi Galvani (Galvani, Luigi). Galvani, a professor of anatomy at the Bologna Academy of Science, was interested in electricity in fish and other animals. One day he noticed that electric sparks from an electrostatic machine caused muscular contractions in a dissected frog that lay nearby. At first, Galvani assumed that the phenomenon was the result of atmospheric electricity because similar effects could be observed during lightning storms. Later, he discovered that whenever a piece of metal connected the muscle and nerve of the frog, the muscle contracted. Although Galvani realized that some metals appeared to be more effective than others in producing this effect, he concluded incorrectly that the metal was transporting a fluid, which he identified with animal electricity, from the nerve to the muscle. Galvani's observations, published in 1791, aroused considerable controversy and speculation.

      Alessandro Volta (Volta, Conte Alessandro), a physicist at the nearby University of Pavia, had been studying how electricity stimulates the senses of touch, taste, and sight. When Volta put a metal coin on top of his tongue and another coin of a different metal under his tongue and connected their surfaces with a wire, the coins tasted salty. Like Galvani, Volta assumed that he was working with animal electricity until 1796 when he discovered that he could also produce a current when he substituted a piece of cardboard soaked in brine for his tongue. Volta correctly conjectured that the effect was caused by the contact between metal and a moist body. Around 1800 he constructed what is now known as a voltaic pile consisting of layers of silver, moist cardboard, and zinc, repeated in that order, beginning and ending with a different metal. When he joined the silver and the zinc with a wire, electricity flowed continuously through the wire. Volta confirmed that the effects of his pile were equivalent in every way to those of static electricity. Within 20 years, galvanism, as electricity produced by a chemical reaction was then called, became unequivocally linked to static electricity. More important, Volta's invention provided the first source of continuous electric current. This rudimentary form of battery produced a smaller voltage than the Leyden jar, but it was easier to use because it could supply a steady current and did not have to be recharged.

      The controversy between Galvani, who mistakenly thought that electricity originated in the animal's nerve, and Volta, who realized that it came from the metal, divided scientists into two camps. Galvani was supported by Alexander von Humboldt (Humboldt, Alexander von) in Germany, while Volta was backed by Coulomb and other French physicists.

      Within six weeks of Volta's report, two English scientists, William Nicholson (Nicholson, William) and Anthony Carlisle, used a chemical battery to discover electrolysis (the process in which an electric current produces a chemical reaction) and initiate the science of electrochemistry. In their experiment the two employed a voltaic pile to liberate hydrogen and oxygen from water. They attached each end of the pile to brass wires and placed the opposite ends of the wires into salt water. The salt made the water a conductor. Hydrogen gas accumulated at the end of one wire; the end of the other wire was oxidized. Nicholson and Carlisle discovered that the amount of hydrogen and oxygen set free by the current was proportional to the amount of current used. By 1809 the English chemist Humphry Davy (Davy, Sir Humphry, Baronet) had used a stronger battery to free for the first time several very active metals—sodium, potassium, calcium, strontium, barium, and magnesium—from their liquid compounds. Faraday, who was Davy's assistant at the time, studied electrolysis quantitatively and showed that the amount of energy needed to separate a gram of a substance from its compound is closely related to the atomic weight of the substance. Electrolysis became a method of measuring electric current; and the quantity of charge that releases a gram atomic weight of a simple element is now called a faraday in his honour.

      Once scientists were able to produce currents with a battery, they could study the flow of electricity quantitatively. Because of the battery, the German physicist Georg Simon Ohm (Ohm, Georg Simon) was able experimentally in 1827 to quantify precisely a problem that Cavendish could only investigate qualitatively some 50 years earlier—namely, the ability of a material to conduct electricity. The result of this work— Ohm's law—explains how the resistance to the flow of charge depends on the type of conductor and on its length and diameter. According to Ohm's formulation, the current flow through a conductor is directly proportional to the potential difference, or voltage, and inversely proportional to the resistance—that is, i = V/R. Thus, doubling the length of an electric wire doubles its resistance, while doubling the cross-sectional area of the wire reduces the resistance by a half. Ohm's law is probably the most widely used equation in electric design.

Experimental and theoretical studies of electromagnetic (electromagnetism) phenomena
      One of the great turning points in the development of the physical sciences was Hans Christian Ørsted's (Ørsted, Hans Christian) announcement in 1820 that electric currents produce magnetic effects. (Ørsted made his discovery while lecturing to a class of physics students. He placed by chance a wire carrying current near a compass needle and was surprised to see the needle swing at right angles to the wire.) Ørsted's fortuitous discovery proved that electricity and magnetism are linked. His finding, together with Faraday's subsequent discovery that a changing magnetic field produces an electric current in a nearby circuit, formed the basis of both James Clerk Maxwell's unified theory of electromagnetism and most of modern electrotechnology.

      Once Ørsted's experiment had revealed that electric currents have magnetic effects, scientists realized that there must be magnetic forces between the currents. They began studying the forces immediately. A French physicist, François Arago (Arago, François), observed in 1820 that an electric current will orient unmagnetized iron filings in a circle around the wire. That same year, another French physicist, André-Marie Ampère (Ampère, André-Marie), developed Ørsted's observations in quantitative terms. Ampère showed that two parallel wires carrying electric currents attract and repel each other like magnets. If the currents flow in the same direction, the wires attract each other; if they flow in opposite directions, the wires repel each other. From this experiment, Ampère was able to express the right-hand rule for the direction of the force on a current in a magnetic field. He also established experimentally and quantitatively the laws of magnetic force between electric currents. He suggested that internal electric currents are responsible for permanent magnets and for highly magnetizable materials like iron. With Arago, he demonstrated that steel needles become more strongly magnetic inside a coil carrying an electric current. Experiments on small coils showed that, at large distances, the forces between two such coils are similar to those between two small bar magnets and, moreover, that one coil can be replaced by a bar magnet of suitable size without changing the forces. The magnetic moment of this equivalent magnet was determined by the dimensions of the coil, its number of turns, and the current flowing around it.

 William Sturgeon (Sturgeon, William) of England and Joseph Henry (Henry, Joseph) of the United States used Ørsted's discovery to develop electromagnets (electromagnet) during the 1820s. Sturgeon wrapped 18 turns of bare copper wire around a U-shaped iron bar. When he turned on the current, the bar became an electromagnet capable of lifting 20 times its weight. When the current was turned off, the bar was no longer magnetized. Henry repeated Sturgeon's work in 1829, using insulated wire to prevent short-circuiting. Using hundreds of turns, Henry created an electromagnet that could lift more than one ton of iron.

      Ørsted's experiment showing that electricity could produce magnetic effects raised the opposite question as well: Could magnetism induce an electric current in another circuit? The French physicist Augustin-Jean Fresnel (Fresnel, Augustin-Jean) argued that since a steel bar inside a metallic helix can be magnetized by passing a current through the helix, the bar magnet in turn should create a current in an enveloping helix. In the following decade many ingenious experiments were devised, but the expectation that a steady current would be induced in a coil near the magnet resulted in experimenters either accidentally missing or not appreciating any transient electric effects caused by the magnet.

Faraday's discovery of electric induction
 Faraday, the greatest experimentalist in electricity and magnetism of the 19th century and one of the greatest experimental physicists of all time, worked on and off for 10 years trying to prove that a magnet could induce (Faraday's law of induction) electricity. In 1831 he finally succeeded by using two coils of wire wound around opposite sides of a ring of soft iron (Figure 7—>). The first coil was attached to a battery; when a current passed through the coil, the iron ring became magnetized. A wire from the second coil was extended to a compass needle a metre away, far enough so that it was not affected directly by any current in the first circuit. When the first circuit was turned on, Faraday observed a momentary deflection of the compass needle and its immediate return to its original position. When the primary current was switched off, a similar deflection of the compass needle occurred but in the opposite direction. Building on this observation in other experiments, Faraday showed that changes in the magnetic field around the first coil are responsible for inducing the current in the second coil. He also demonstrated that an electric current can be induced by moving a magnet, by turning an electromagnet on and off, and even by moving an electric wire in the Earth's magnetic field. Within a few months, Faraday built the first, albeit primitive, electric generator.

      Henry had discovered electric induction quite independently in 1830, but his results were not published until after he had received news of Faraday's 1831 work, nor did he develop the discovery as fully as Faraday. In his paper of July 1832, Henry reported and correctly interpreted self-induction. He had produced large electric arcs from a long helical conductor when it was disconnected from a battery. When he had opened the circuit, the rapid decrease in the current had caused a large voltage between the battery terminal and the wire. As the wire lead was pulled away from the battery, the current continued to flow for a short time in the form of a bright arc between the battery terminal and the wire.

      Faraday's thinking was permeated by the concept of electric and magnetic lines of force (force, line of). He visualized that magnets, electric charges, and electric currents produce lines of force. When he placed a thin card covered with iron filings on a magnet, he could see the filings form chains from one end of the magnet to the other. He believed that these lines showed the directions of the forces and that electric current would have the same lines of force. The tension they build explains the attraction and repulsion of magnets and electric charges. Faraday had visualized magnetic curves as early as 1831 while working on his induction experiments; he wrote in his notes, “By magnetic curves I mean lines of magnetic forces which would be depicted by iron filings.” Faraday opposed the prevailing idea that induction occurred “at a distance”; instead, he held that induction occurs along curved lines of force because of the action of contiguous particles. Later, he explained that electricity and magnetism are transmitted through a medium that is the site of electric or magnetic “fields,” which make all substances magnetic to some extent.

      Faraday was not the only researcher laying the groundwork for a synthesis between electricity, magnetism, and other areas of physics. On the continent of Europe, primarily in Germany, scientists were making mathematical connections between electricity, magnetism, and optics. The work of the physicists Franz Ernst Neumann, Wilhelm Eduard Weber, and H.F.E. Lenz belongs to this period. At the same time, Helmholtz and the English physicists William Thomson (later Lord Kelvin) and James Prescott Joule (Joule, James Prescott) were clarifying the relationship between electricity and other forms of energy. Joule investigated the quantitative relationship between electric currents and heat during the 1840s and formulated the theory of the heating effects that accompany the flow of electricity in conductors. Helmholtz, Thomson, Henry, Gustav Kirchhoff (Kirchhoff, Gustav Robert), and Sir George Gabriel Stokes also extended the theory of the conduction and propagation of electric effects in conductors. In 1856 Weber (Weber, Wilhelm Eduard) and his German colleague, Rudolf Kohlrausch (Kohlrausch, Friedrich Wilhelm Georg), determined the ratio of electric and magnetic units and found that it has the same dimensions as light and that it is almost exactly equal to its velocity. In 1857 Kirchhoff used this finding to demonstrate that electric disturbances propagate on a highly conductive wire with the speed of light.

Maxwell's (Maxwell, James Clerk) unified theory of electromagnetism
      The final steps in synthesizing electricity and magnetism into one coherent theory were made by Maxwell. He was deeply influenced by Faraday's work, having begun his study of the phenomena by translating Faraday's experimental findings into mathematics. (Faraday was self-taught and had never mastered mathematics.) In 1856 Maxwell developed the theory that the energy of the electromagnetic field is in the space around the conductors as well as in the conductors themselves. By 1864 he had formulated his own electromagnetic theory of light, predicting that both light and radio waves are electric and magnetic phenomena. While Faraday had discovered that changes in magnetic fields produce electric fields (electric field), Maxwell added the converse: changes in electric fields produce magnetic fields even in the absence of electric currents. Maxwell predicted that electromagnetic disturbances traveling through empty space have electric and magnetic fields at right angles to each other and that both fields are perpendicular to the direction of the wave. He concluded that the waves move at a uniform speed equal to the speed of light and that light is one form of electromagnetic wave. Their elegance notwithstanding, Maxwell's radical ideas were accepted by few outside England until 1886, when the German physicist Heinrich Hertz verified the existence of electromagnetic waves traveling at the speed of light; the waves he discovered are known now as radio waves.

      Maxwell's four field equations represent the pinnacle of classical electromagnetic theory. Subsequent developments in the theory have been concerned either with the relationship between electromagnetism and the atomic structure of matter or with the practical and theoretical consequences of Maxwell's equations. His formulation has withstood the revolutions of relativity and quantum mechanics. His equations are appropriate for distances as small as 10-10 centimetres—100 times smaller than the size of an atom. The fusion of electromagnetic theory and quantum theory, known as quantum electrodynamics, is required only for smaller distances.

      While the mainstream of theoretical activity concerning electric and magnetic phenomena during the 19th century was devoted to showing how they are interrelated, some scientists made use of them to discover new properties of materials and heat. Weber developed Ampère's suggestion that there are internal circulating currents of molecular size in metals. He explained how a substance loses its magnetic properties when the molecular magnets point in random directions. Under the action of an external force, they may turn to point in the direction of the force; when all point in this direction, the maximum possible degree of magnetization is reached, a phenomenon known as magnetic saturation. In 1895 Pierre Curie of France discovered that a ferromagnetic substance has a specific temperature above which it ceases to be magnetic. Finally, superconductivity was discovered in 1900 by the German physicist Heike Kammerlingh-Onnes. In superconductivity electric conductors lose all resistance at very low temperatures.

Discovery of the electron and its ramifications
      Although little of major importance was added to electromagnetic theory in the 19th century after Maxwell, the discovery of the electron in 1898 opened up an entirely new area of study: the nature of electric charge and of matter itself. The discovery of the electron grew out of studies of electric currents in vacuum tubes (electron tube). Heinrich Geissler (Geissler, Heinrich), a glassblower who assisted the German physicist Julius Plücker (Plücker, Julius), improved the vacuum tube in 1854. Four years later, Plücker sealed two electrodes inside the tube, evacuated the air, and forced electric currents between the electrodes; he attributed the green glow that appeared on the wall of the tube to rays emanating from the cathode (cathode ray). From then until the end of the century, the properties of cathode-ray discharges were studied intensively. The work of the English physicist Sir William Crookes (Crookes, Sir William) in 1879 indicated that the luminescence was a property of the electric current itself. Crookes concluded that the rays were composed of electrified charged particles. In 1898 another English physicist, Sir J.J. Thomson (Thomson, Sir J.J.), identified a cathode ray as a stream of negatively charged particles, each having a mass 1/1836 smaller than that of a hydrogen ion. Thomson's discovery established the particulate nature of charge; his particles were later dubbed electrons.

      Following the discovery of the electron, electromagnetic theory became an integral part of the theories of the atomic, subatomic, and subnuclear structure of matter. This shift in focus occurred as the result of an impasse between electromagnetic theory and statistical mechanics over attempts to understand radiation from hot bodies. thermal radiation had been investigated in Germany by the physicist Wilhelm Wien (Wien, Wilhelm) between 1890 and 1900. Wien had virtually exhausted the resources of thermodynamics in dealing with this problem. Two British scientists, Lord Rayleigh (Rayleigh, John William Strutt, 3rd Baron) (John William Strutt) and Sir James Hopwood Jeans (Jeans, Sir James), had by 1900 applied the newly developed science of statistical mechanics to the same problem. They obtained results that, though in agreement with Wien's thermodynamic conclusions (as distinct from his speculative extensions of thermodynamics), only partially agreed with experimental observations. The German physicist Max Planck (Planck, Max) attempted to combine the statistical approach with a thermodynamic approach. By concentrating on the necessity of fitting together the experimental data, he was led to the formulation of an empirical law that satisfied Wien's thermodynamic criteria and accommodated the experimental data. When Planck interpreted this law in terms of Rayleigh's statistical concepts, he concluded that radiation of frequency ν exists only in quanta (quantum) of energy. Planck's result, including the introduction of the new universal constant h in 1900, marked the foundation of quantum mechanics and initiated a profound change in physical theory (see atom: Bohr's shell model (atom)).

      By 1900 it was apparent that Thomson's electrons were a universal constituent of matter and, thus, that matter is essentially electric in nature. As a result, in the early years of the 20th century, many physicists attempted to construct theories of the electromagnetic properties of metals, insulators, and magnetic materials in terms of electrons. In 1909 the Dutch physicist Hendrik Antoon Lorentz (Lorentz, Hendrik Antoon) succeeded in doing so in The Theory of Electrons and Its Applications to the Phenomena of Light and Radiant Heat; his work has since been modified by quantum theory.

      The other major conceptual advance in electromagnetic theory was the special theory of relativity. In Maxwell's time, a mechanistic view of the universe held sway. Sound was interpreted as an undulatory motion of the air, while light and other electromagnetic waves were regarded as undulatory motions of an intangible medium called ether. The question arose as to whether the velocity of light measured by an observer moving relative to ether would be affected by his motion. Albert Abraham Michelson (Michelson, A.A.) and Edward W. Morley of the United States had demonstrated in 1887 that light in a vacuum on Earth travels at a constant speed which is independent of the direction of the light relative to the direction of the Earth's motion through the ether. Lorentz and Henri Poincaré (Poincaré, Henri), a French physicist, showed between 1900 and 1904 that the conclusions of Michelson and Morley were consistent with Maxwell's equations. On this basis, Lorentz and Poincaré developed a theory of relativity in which the absolute motion of a body relative to a hypothetical ether is no longer significant. Poincaré named the theory the principle of relativity in a lecture at the St. Louis Exposition in September 1904. Planck gave the first formulation of relativistic dynamics two years later. The most general formulation of the special theory of relativity, however, was put forth by Einstein (Einstein, Albert) in 1905, and the theory of relativity is usually associated with his name. Einstein postulated that the speed of light is a constant, independent of the motion of the source of the light, and showed how the Newtonian laws of mechanics would have to be modified. While Maxwell had synthesized electricity and magnetism into one theory, he had regarded them as essentially two interdependent phenomena; Einstein showed that they are two aspects of the same phenomenon.

      Maxwell's equations, the special theory of relativity, the discovery of the electronic structure of matter, and the formulation of quantum mechanics all occurred before 1930. The quantum electrodynamics theory, developed between 1945 and 1955, subsequently resolved some minute discrepancies in the calculations of certain atomic properties. For example, the accuracy with which it is now possible to calculate one of the numbers describing the magnetic moment of the electron is comparable to measuring the distance between New York City and Los Angeles to within the thickness of a human hair. As a result, quantum electrodynamics is the most complete and precise theory of any physical phenomenon. The remarkable correspondence between theory and observation makes it unique among human endeavours.

Development of electromagnetic (electromagnetism) technology (technology, history of)
      Electromagnetic technology began with Faraday's discovery of induction in 1831 (see above). His demonstration that a changing magnetic field induces an electric current in a nearby circuit showed that mechanical energy (energy conversion) can be converted to electric energy. It provided the foundation for electric power generation, leading directly to the invention of the dynamo (electric generator) and the electric motor. Faraday's finding also proved crucial for lighting and heating systems.

      The early electric industry was dominated by the problem of generating electricity on a large scale. Within a year of Faraday's discovery, a small hand-turned generator in which a magnet revolved around coils was demonstrated in Paris. In 1833 there appeared an English model that featured the modern arrangement of rotating the coils in the field of a fixed magnet. By 1850 generators were manufactured commercially in several countries. Permanent magnets were used to produce the magnetic field in generators until the principle of the self-excited generator was discovered in 1866. (A self-excited generator has stronger magnetic fields because it uses electromagnets powered by the generator itself.) In 1870 Zénobe Théophile Gramme (Gramme, Zénobe-Théophile), a Belgian manufacturer, built the first practical generator capable of producing a continuous current. It was soon found that the magnetic field is more effective if the coil windings are embedded in slots in the rotating iron armature. The slotted armature, still in use today, was invented in 1880 by the Swedish engineer Jonas Wenström. Faraday's 1831 discovery of the principle of the AC transformer was not put to practical use until the late 1880s when the heated debate over the merits of direct-current and alternating-current systems for power transmission was settled in favour of the latter.

      At first, the only serious consideration for electric (electric arc) power was arc (arc lamp) lighting, in which a brilliant light is emitted by an electric spark between two electrodes. The arc lamp was too powerful for domestic use, however, and so it was limited to large installations like lighthouses, train stations, and department stores. Commercial development of an incandescent filament lamp, first invented in the 1840s, was delayed until a filament could be made that would heat to incandescence without melting and until a satisfactory vacuum tube could be built. The mercury pump, invented in 1865, provided an adequate vacuum, and a satisfactory carbon filament was developed independently by the English physicist Sir Joseph Wilson Swan and the American inventor Thomas A. Edison (Edison, Thomas Alva) during the late 1870s. By 1880 both had applied for patents for their incandescent lamps, and the ensuing litigation between the two men was resolved by the formation of a joint company in 1883. Thanks to the incandescent lamp, electric lighting became an accepted part of urban life by 1900. Since then, the tungsten filament lamp, introduced during the early 1900s, has become the principal form of electric lamp, though more efficient fluorescent gas discharge lamps have found widespread use as well.

      Electricity took on a new importance with the development of the electric motor. This machine, which converts electric energy to mechanical energy, has become an integral component of a wide assortment of devices ranging from kitchen appliances and office equipment to industrial robots and rapid-transit vehicles. Although the principle of the electric motor was devised by Faraday (Faraday, Michael) in 1821, no commercially significant unit was produced until 1873. In fact, the first important AC motor, built by the Serbian-American inventor Nikola Tesla (Tesla, Nikola), was not demonstrated in the United States until 1888. Tesla began producing his motors in association with the Westinghouse Electric Company a few years after DC motors had been installed in trains in Germany and Ireland. By the end of the 19th century, the electric motor had taken a recognizably modern form. Subsequent improvements have rarely involved radically new ideas; however, the introduction of better designs and new bearing, armature, magnetic, and contact materials has resulted in the manufacture of smaller, cheaper, and more efficient and reliable motors.

      The modern communications industry is among the most spectacular products of electricity. telegraph systems using wires and simple electrochemical or electromechanical receivers proliferated in western Europe and the United States during the 1840s. An operable cable was installed under the English Channel in 1865, and a pair of transatlantic cables were successfully laid a year later. By 1872 almost all of the major cities of the world were linked by telegraph.

      Alexander Graham Bell (Bell, Alexander Graham) patented the first practical telephone in the United States in 1876, and the first public telephone services were operating within a few years. In 1895 the British physicist Sir Ernest Rutherford (Rutherford, Ernest, Baron Rutherford of Nelson, of Cambridge) advanced Hertz's scientific investigations of radio waves and transmitted radio signals for more than one kilometre. Guglielmo Marconi (Marconi, Guglielmo), an Italian physicist and inventor, established wireless communications across the Atlantic employing radio waves of approximately 300- to 3,000-metre wavelength in 1901. Broadcast (broadcasting) radio transmissions were established during the 1920s.

      Telephone transmissions by radio waves, the electric recording and reproduction of sound, and television were made possible by the development of the triode tube. This three-electrode tube, invented by the American engineer Lee De Forest (De Forest, Lee), permitted for the first time the amplification of electric signals. Known as the Audion, this device played a pivotal role in the early development of the electronics industry.

      The first telephone transmission via radio signals was made from Arlington, Va., to the Eiffel Tower in Paris in 1915; and a commercial radio telephone service between New York City and London was begun in 1927. Besides such efforts, most of the major developmental work of this period was tied to the radio and phonograph entertainment industries and the sound film industry. Rapid progress was made toward transmitting moving pictures (motion-picture technology), especially in Great Britain; just before World War II, the British Broadcasting Corporation inaugurated the first public television service. Today, many regions of the electromagnetic spectrum are used for communications, including microwaves in the frequency range of approximately 7 × 109 hertz for satellite communication links and infrared light at a frequency of about 3 × 1014 hertz for optical fibre communications systems.

      Until 1939 the electronics industry was almost exclusively concerned with communications and broadcast entertainment. Scientists and engineers in Britain, Germany, France, and the United States did initiate research on radar systems capable of aircraft detection and antiaircraft fire-control during the 1930s, however, and this marked the beginning of a new direction for electronics. During World War II and after, the electronics industry made strides paralleled only by those of the chemical industry. Television became commonplace; and a broad array of new devices and systems, most notably the electronic digital computer, emerged.

      The electronic revolution of the last half of the 20th century has been made possible in large part by the invention of the transistor (1947) and such subsequent developments as the integrated circuit. (For detailed coverage of these and other major advances, see electronics.) This miniaturization and integration of circuit elements has led to a remarkable diminution in the size and cost of electronic equipment and an equally impressive increase in its reliability.

Frank Neville H. Robinson Edwin Kashy Sharon Bertsch McGrayne

Additional Reading
Richard P. Feynman, Robert B. Leighton, and Matthew Sands, The Feynman Lectures on Physics, vol. 2, The Electromagnetic Field (1964, reprinted 1977), is highly recommended for its lucid discussion of fundamentals. The same is true for Edward M. Purcell, Electricity and Magnetism, 2nd ed. (1985). John R. Reitz, Frederick J. Milford, and Robert W. Christy, Foundations of Electromagnetic Theory, 3rd ed. (1979), is a fine, compact, college-level text using vector calculus.J.L. Heilbron, Electricity in the 17th and 18th Centuries: A Study of Early Modern Physics (1979), provides a readable survey of significant developments, as does Edmund Whittaker, A History of the Theories of Aether and Electricity, rev. and enlarged ed., 2 vol. (1951–53, reprinted 1973). Charles Singer and T.I. Williams (eds.), A History of Technology, 8 vol. (1954–84), begins in the prehistoric period and concludes around 1950.Edwin Kashy Sharon Bertsch McGrayne

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • Electromagnetism — Electricity · …   Wikipedia

  • electromagnetism — ELECTROMAGNETÍSM s.n. Ramură a electricităţii care studiază câmpul magnetic produs de sarcinile electrice în mişcare, acţiunea câmpului magnetic asupra curentului electric, undele electrice şi cele magnetice, fenomenul inducţiei electromagnetice… …   Dicționar Român

  • electromagnetism — 1828; see ELECTRO (Cf. electro ) + MAGNETISM (Cf. magnetism) …   Etymology dictionary

  • electromagnetism — [ē lek΄trōmag′nə tiz΄əm, i lek΄trōmag′nə tiz΄əm ē lek΄trəmag′nə tiz΄əm, i lek΄trəmag′nə tiz΄əm] n. 1. magnetism produced by an electric current 2. a) the interaction between a magnetic field and an electric field b) the phenomena produced by such …   English World dictionary

  • electromagnetism — elektromagnetizmas statusas T sritis automatika atitikmenys: angl. electromagnetism vok. Elektromagnetismus, m rus. электромагнетизм, m pranc. électromagnétisme, m …   Automatikos terminų žodynas

  • electromagnetism — elektromagnetizmas statusas T sritis fizika atitikmenys: angl. electromagnetics; electromagnetism vok. Elektromagnetismus, m rus. электромагнетизм, m pranc. électromagnétisme, m …   Fizikos terminų žodynas

  • electromagnetism — electromagnetic ► ADJECTIVE ▪ relating to the interrelation of electric currents or fields and magnetic fields. DERIVATIVES electromagnetically adverb electromagnetism noun …   English terms dictionary

  • electromagnetism — noun Date: 1828 1. magnetism developed by a current of electricity 2. a. a fundamental physical force that is responsible for interactions between charged particles which occur because of their charge and for the emission and absorption of… …   New Collegiate Dictionary

  • electromagnetism — noun /ɪˌlɛk.tɹəʊˈmæɡ.nəˌtɪzm̩/ a) A unified fundamental force that combines the aspects of electricity and magnetism and is one of the four fundamental forces. (technically it can be unified with weak nuclear to form electroweak) Its gauge boson… …   Wiktionary

  • electromagnetism — elec·tro·mag·ne·tism i .lek trō mag nə .tiz əm n 1) magnetism developed by a current of electricity 2) physics dealing with the relations between electricity and magnetism * * * elec·tro·mag·net·ism (e lek″tro magґnə tiz əm) 1.… …   Medical dictionary

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”