electronics

electronics
/i lek tron"iks, ee'lek-/, n.
(used with a sing. v.) the science dealing with the development and application of devices and systems involving the flow of electrons in a vacuum, in gaseous media, and in semiconductors.
[1905-10; see ELECTRONIC, -ICS]

* * *

I
Branch of physics that deals with the emission, behaviour, and effects of electrons and with electronic devices.

The beginnings of electronics can be traced to experiments with electricity. In the 1880s Thomas Alva Edison and others observed the flow of current between elements in an evacuated glass tube. A two-electrode vacuum tube constructed by John A. Fleming (1849–1945) produced a useful output current. The Audion, invented by Lee De Forest (1907), was followed by further improvements. The invention of the transistor at Bell Labs (1947) initiated a progressive miniaturization of electronic components that by the mid 1980s resulted in high-density microprocessors, which in turn led to tremendous advances in computer technology and computer-based automated systems. See also semiconductor.
II
(as used in expressions)
Koninklijke Philips Electronics NV
Royal Philips Electronics NV

* * *

Introduction

      branch of physics and electrical engineering that deals with the emission, behaviour, and effects of electrons and with electronic devices.

      Electronics encompasses an exceptionally broad range of technology. The term originally was applied to the study of electron behaviour and movement, particularly as observed in the first electron tubes (electron tube). It came to be used in its broader sense with advances in knowledge about the fundamental nature of electrons and about the way in which the motion of these particles could be utilized. Today many scientific and technical disciplines deal with different aspects of electronics. Research in these fields has led to the development of such key devices as transistors (transistor), integrated circuits (integrated circuit), lasers (laser), and optical fibres (fibre optics). These in turn have made it possible to manufacture a wide array of electronic consumer, industrial, and military products. Indeed, it can be said that the world is in the midst of an electronic revolution at least as significant as the industrial revolution of the 19th century.

      This article reviews the historical development of electronics, highlighting major discoveries and advances. It also describes some key electronic functions and the manner in which various devices carry out these functions.

The history of electronics

The vacuum tube era
      Theoretical and experimental studies of electricity during the 18th and 19th centuries led to the development of the first electrical machines and the beginning of the widespread use of electricity. The history of electronics began to evolve separately from that of electricity late in the 19th century with the identification of the electron by the English physicist Sir Joseph John Thomson (Thomson, Sir J.J.) and the measurement of its electric charge by the American physicist Robert A. Millikan (Millikan, Robert Andrews) in 1909.

      At the time of Thomson's work, the American inventor Thomas A. Edison (Edison, Thomas Alva) had observed a bluish glow in some of his early lightbulbs under certain conditions and found that a current would flow from one electrode in the lamp to another if the second one ( anode) were made positively charged with respect to the first ( cathode). Work by Thomson and his students and by the English engineer John Ambrose Fleming (Fleming, Sir John Ambrose) revealed that this so-called Edison effect (thermionic emission) was the result of the emission of electrons from the cathode, the hot filament in the lamp. The motion of the electrons to the anode, a metal plate, constituted an electric current that would not exist if the anode were negatively charged.

      This discovery provided impetus for the development of electron tubes (electron tube), including an improved X-ray tube by the American engineer William D. Coolidge (Coolidge, William D) and Fleming's thermionic valve (a two-electrode vacuum tube) for use in radio receivers. The detection of a radio signal, which is a very high-frequency alternating current (AC), requires that the signal be rectified; i.e., the alternating current must be converted into a direct current (DC) by a device that conducts only when the signal has one polarity but not when it has the other—precisely what Fleming's valve (patented in 1904) did. Previously, radio signals were detected by various empirically developed devices such as the “cat whisker” detector, which was composed of a fine wire (the whisker) in delicate contact with the surface of a natural crystal of lead sulfide (galena) or some other semiconductor material. These devices were undependable, lacked sufficient sensitivity, and required constant adjustment of the whisker-to-crystal contact to produce the desired result. Yet these were the forerunners of today's solid-state devices. The fact that crystal rectifiers worked at all encouraged scientists to continue studying them and gradually to obtain the fundamental understanding of the electrical properties of semiconducting materials necessary to permit the invention of the transistor.

      In 1906 Lee De Forest (De Forest, Lee), an American engineer, developed a type of vacuum tube that was capable of amplifying radio signals. De Forest added a grid of fine wire between the cathode and anode of the two-electrode thermionic valve constructed by Fleming. The new device, which De Forest dubbed the Audion, was thus a three-electrode vacuum tube. In operation, the anode in such a vacuum tube is given a positive potential (positively biased) with respect to the cathode, while the grid is negatively biased. A large negative bias on the grid prevents any electrons emitted from the cathode from reaching the anode; however, because the grid is largely open space, a less negative bias permits some electrons to pass through it and reach the anode. Small variations in the grid potential can thus control large amounts of anode current.

      The vacuum tube permitted the development of radio broadcasting, long-distance telephony, television, and the first electronic digital computers. These early electronic computers were, in fact, the largest vacuum-tube systems ever built. Perhaps the best-known representative is the ENIAC (computer) (Electronic Numerical Integrator and Computer (ENIAC)), completed in 1946.

      The special requirements of the many different applications of vacuum tubes led to numerous improvements, enabling them to handle large amounts of power, operate at very high frequencies, have greater than average reliability, or be made very compact (the size of a thimble). The cathode-ray tube, originally developed for displaying electrical waveforms on a screen for engineering measurements, evolved into the television picture tube. Such tubes operate by forming the electrons emitted from the cathode into a thin beam that impinges on a fluorescent screen at the end of the tube. The screen emits light that can be viewed from outside the tube. Deflecting the electron beam causes patterns of light to be produced on the screen, creating the desired optical images.

      Other specialized types of vacuum tubes, developed or refined during World War II for military purposes, are still used today in microwave ovens and as extremely high-frequency transmitters aboard space satellites. Notwithstanding the remarkable success of solid-state devices in most electronic applications, there are certain specialized functions that only vacuum tubes can perform. These usually involve operation at extremes of power or frequency. Vacuum tubes continue to be used as display devices for television sets and computer monitors because other means of providing the function are more expensive, though even this situation is changing.

      Vacuum tubes are fragile and ultimately wear out in service. Failure occurs in normal usage either from the effects of repeated heating and cooling as equipment is switched on and off (thermal fatigue), which ultimately causes a physical fracture in some part of the interior structure of the tube, or from degradation of the properties of the cathode by residual gases in the tube. Vacuum tubes also take time (from a few seconds to several minutes) to “warm up” to operating temperature—an inconvenience at best and in some cases a serious limitation to their use. These shortcomings motivated scientists at Bell Laboratories to seek an alternative to the vacuum tube and led to the development of the transistor.

The semiconductor revolution
Invention of the transistor
      The invention of the transistor in 1947 by John Bardeen (Bardeen, John), Walter H. Brattain (Brattain, Walter H.), and William B. Shockley (Shockley, William B.) of the Bell research staff provided the first of a series of new devices with remarkable potential for expanding the utility of electronic equipment. Transistors, along with such subsequent developments as integrated circuits (integrated circuit), are made of crystalline solid materials called semiconductors (semiconductor), which have electrical properties that can be varied over an extremely wide range by the addition of minuscule quantities of other elements. The electric current in semiconductors is carried by electrons, which have a negative charge, and also by “holes,” analogous entities that carry a positive charge. The availability of two kinds of charge carriers in semiconductors is a valuable property exploited in many electronic devices made of such materials.

      Early transistors were produced using germanium as the semiconductor material, because methods of purifying it to the required degree had been developed during and shortly after World War II. Because the electrical properties of semiconductors are extremely sensitive to the slightest trace of certain other elements, only about one part per billion of such elements can be tolerated in material to be used for making semiconductor devices.

      During the late 1950s, research on the purification of silicon succeeded in producing material suitable for semiconductor devices, and new devices made of silicon were manufactured from about 1960. Silicon quickly became the preferred raw material, because it is much more abundant than germanium and thus less expensive. In addition, silicon retains its semiconducting properties at higher temperatures than does germanium. Silicon diodes can be operated at temperatures up to 200 °C (400 °F), whereas germanium diodes cannot be operated above 85 °C (185 °F). There was one other important property of silicon, not appreciated at the time but crucial to the development of low-cost transistors and integrated circuits: silicon, unlike germanium, forms a tenaciously adhering oxide film with excellent electrical insulating properties when it is heated to high temperatures in the presence of oxygen. This film is utilized as a mask to permit the desired impurities that modify the electrical properties of silicon to be introduced into it during manufacture of semiconductor devices. The mask pattern, formed by a photolithographic process, permits the creation of tiny transistors and other electronic components in the silicon.

Integrated circuits
      By 1960 vacuum tubes were rapidly being supplanted by transistors, because the latter had become less expensive, did not burn out in service, and were much smaller and more reliable. Computers (computer) employed hundreds of thousands of transistors each. This fact, together with the need for compact, lightweight electronic missile-guidance systems, led to the invention of the integrated circuit (IC) independently by Jack Kilby (Kilby, Jack) of Texas Instruments Incorporated in 1958 and by Jean Hoerni and Robert Noyce (Noyce, Robert) of Fairchild Semiconductor Corporation in 1959. Kilby is usually credited with having developed the concept of integrating device and circuit elements onto a single silicon chip, while Noyce is given credit for having conceived the method for integrating the separate elements.

      Early ICs contained about 10 individual components on a silicon chip 3 mm (0.12 inch) square. By 1970 the number was up to 1,000 on a chip of the same size at no increase in cost. Late in the following year the first microprocessor was introduced. The device contained all the arithmetic, logic, and control circuitry required to perform the functions of a computer's central processing unit (CPU). This type of large-scale IC was developed by a team at Intel Corporation, the same company that also introduced the memory IC in 1971. The stage was now set for the computerization of small electronic equipment.

      Until the microprocessor appeared on the scene, computers were essentially discrete pieces of equipment used primarily for data processing and scientific calculations. They ranged in size from minicomputers (computer), comparable in dimensions to a small filing cabinet, to mainframe systems that could fill a large room. The microprocessor enabled computer engineers to develop microcomputers (microcomputer)—systems about the size of a lunch box or smaller but with enough computing power to perform many kinds of business, industrial, and scientific tasks. Such systems made it possible to control a host of small instruments or devices (e.g., numerically controlled lathes and one-armed robotic devices for spot welding) by using standard components programmed to do a specific job. The very existence of computer hardware inside such devices is not apparent to the user.

      The large demand for microprocessors generated by these initial applications led to high-volume production and a dramatic reduction in cost. This in turn promoted the use of the devices in many other applications—for example, in household appliances and automobiles, for which electronic controls had previously been too expensive to consider. Continued advances in IC technology gave rise to very large-scale integration (VLSI), which substantially increased the circuit density of microprocessors. These technological advances, coupled with further cost reductions stemming from improved manufacturing methods, made feasible the mass production of personal computers (personal computer) for use in offices, schools, and homes.

      By the mid-1980s inexpensive microprocessors had stimulated computerization of an enormous variety of consumer products. Common examples included programmable microwave ovens and thermostats, clothes washers and dryers, self-tuning television sets and self-focusing cameras, videocassette recorders and video games, telephones and answering machines, musical instruments, watches, and security systems. Microelectronics also came to the fore in business, industry, government, and other sectors. Microprocessor-based equipment proliferated, ranging from automatic teller machines (ATMs) and point-of-sale terminals in retail stores to automated factory assembly systems and office workstations.

 By mid-1986 memory ICs with a capacity of 262,144 bits (binary digits) were available. In fact, Gordon E. Moore (Moore, Gordon E.), one of the founders of Intel, observed as early as 1965 that the complexity of ICs was approximately doubling every 18–24 months, which was still the case in 2000. This empirical “Moore's law” is widely used in forecasting the technological requirements for manufacturing future ICs (see figure—>).

Compound semiconductor materials
      Many semiconductor materials other than silicon and germanium exist, and they have different useful properties. Silicon carbide is a compound semiconductor, the only one composed of two elements from column IV of the periodic table (periodic law). It is particularly suited for making devices for specialized high-temperature applications. Other compounds formed by combining elements from column III of the periodic table—such as aluminum, gallium, and indium—with elements from column V—such as phosphorus, arsenic, and antimony—are of particular interest. These so-called III-V compounds are used to make semiconductor devices that emit light efficiently or that operate at exceptionally high frequencies.

      A remarkable characteristic of these compounds is that they can, in effect, be mixed together. One can produce gallium arsenide or substitute aluminum for some of the gallium or also substitute phosphorus for some of the arsenic. When this is done, the electrical and optical properties of the material are subtly changed in a continuous fashion in proportion to the amount of aluminum or phosphorus used.

      Except for silicon carbide, these compounds have the same crystal structure. This makes possible the gradation of composition, and thus the properties, of the semiconductor material within one continuous crystalline body. Modern material-processing techniques allow these compositional changes to be controlled accurately on an atomic scale.

      These characteristics are exploited in making semiconductor lasers that produce light of any given wavelength within a considerable range. Such lasers are used, for example, in compact disc players and as light sources for optical fibre communication.

Digital electronics
      Computers understand only two numbers, 0 and 1, and do all their arithmetic operations in this binary (binary number system) mode. Many electrical and electronic devices have two states: they are either off or on. A light switch is a familiar example, as are vacuum tubes and transistors. Because computers have been a major application for integrated circuits from their beginning, digital integrated circuits have become commonplace. It has thus become easy to design electronic systems that use digital language to control their functions and to communicate with other systems.

      A major advantage in using digital methods is that the accuracy of a stream of digital signals can be verified, and, if necessary, errors can be corrected. In contrast, signals that vary in proportion to, say, the sound of an orchestra can be corrupted by “noise,” which once present cannot be removed. An example is the sound from a phonograph record, which always contains some extraneous sound from the surface of the recording groove even when the record is new. The noise becomes more pronounced with wear. Contrast this with the sound from a digital compact disc recording. No sound is heard that was not present in the recording studio. The disc and the player contain error-correcting features that remove any incorrect pulses (perhaps arising from dust on the disc) from the information as it is read from the disc.

      As electronic systems become more complex, it is essential that errors produced by noise be removed; otherwise, the systems may malfunction. Many electronic systems are required to operate in electrically noisy environments, such as in an automobile. The only practical way to assure immunity from noise is to make such a system operate digitally. In principle it is possible to correct for any arbitrary number of errors, but in practice this may not be possible. The amount of extra information that must be handled to correct for large rates of error reduces the capacity of the system to handle the desired information, and so trade-offs are necessary.

      A consequence of the veritable explosion in the number and kinds of electronic systems has been a sharp growth in the electrical noise level of the environment. Any electrical system generates some noise, and all electronic systems are to some degree susceptible to disturbance from noise. The noise may be conducted along wires connected to the system, or it may be radiated through the air. Care is necessary in the design of systems to limit the amount of noise that is generated and to shield the system properly to protect it from external noise sources.

Optoelectronics
      A new direction in electronics employs photons (packets of light) instead of electrons. By common consent these new approaches are included in electronics, because the functions that are performed are, at least for the present, the same as those performed by electronic systems and because these functions usually are embedded in a largely electronic environment. This new direction is called optical electronics or optoelectronics.

      In 1966 it was proposed on theoretical grounds that glass fibres could be made with such high purity that light could travel through them for great distances. Such fibres were produced during the early 1970s. They contain a central core in which the light travels. The outer cladding is made of glass of a different chemical formulation and has a lower optical index of refraction. This difference in refractive index indicates that light travels faster in the cladding than it does in the core. Thus, if the light beam begins to move from the core into the cladding, its path is bent so as to move it back into the core. The light is constrained within the core even if the fibre is bent into a circle.

      The core of early optical fibres was of such a diameter (several micrometres [μm], or about one-tenth the diameter of a human hair) that the various rays of light in the core could travel in slightly different paths, the shortest directly down the axis and other longer paths wandering back and forth across the core. This limited the maximum distance that a pulse of light could travel without becoming unduly spread by the time it arrived at the receiving end of the fibre, with the central ray arriving first and others later. In a digital communications system, successive pulses can overlap one another and be indistinguishable at the receiving end. Such fibres are called multimode fibres, in reference to the various paths (or modes) that the light can follow.

      During the late 1970s, fibres were made with smaller core diameters in which the light was constrained to follow only one path. This occurs if the core has a diameter a little larger than the wavelength of the light traveling in it—i.e., about 10 to 15 μm (0.01 to 0.015 mm, or 0.0004 to 0.0006 inch). These single-mode fibres avoid the difficulty described above. By 1993 optical fibres capable of carrying light signals more than 215 km (135 miles) became available. Such distance records have become obsolete with the use of specialized fibres that incorporate integral amplifying features. Fibres employing these optical amplifiers carry light signals over transoceanic distances without the conventional pulse regeneration measures that were needed in the past.

      Optical fibres have several advantages over copper wires or coaxial cables. They can carry information at a much higher rate, they occupy less space (an important feature in large cities and in buildings), and they are quite insensitive to electrical noise. Moreover, it is virtually impossible to make unauthorized connections to them. Costs, initially high, have dropped to the point where most new installations of telephone circuits between switching centres and over longer distances consist of optical fibres.

      Given the fact that communication signals arrive at a central switching office in optical form, it has been attractive to consider switching them from one route to another by optical means rather than electrically, as is done today. The distances between central offices in most cases are substantially shorter than the distance light can travel within a fibre. Optical switching would make unnecessary the detection and regeneration of the light signals, steps that are currently required. Such optical central-office switches are ready for installation today and will further advance the dramatic changes wrought by the use of light waves rather than electrons.

      Another direction in optoelectronics builds in part on the foregoing developments but to a quite different end. A key problem in developing faster computers and faster integrated circuits to use in them is related to the time required for electrical signals to travel over wire interconnections. This is a difficulty both for the integrated circuits themselves and for the connections between them. Under the best circumstances, electrical signals can travel in a wire at about 90 percent of the speed of light. A more usual rate is 50 percent. Light travels about 30 cm (12 inches) in a billionth of a second. Modern computers operate at speeds of more than one billion operations per second. Thus, if two signals that start simultaneously from different sites are to arrive at their destination simultaneously, the paths they travel must not differ in length by more than a few centimetres.

      Two approaches can be envisioned. In one, all the integrated circuits are placed as close together as possible to minimize the distances that signals must travel. This creates a cooling problem, because the integrated circuits generate heat. In the other possible approach, all the paths for signals are made equal to the longest path. This requires the use of much more wire, because most paths are longer than they would otherwise be. All this wire takes space, which means that the integrated circuits have to be placed farther apart than is preferable.

      Ultimately, as computers operate even faster, neither approach will work, and a radically new technique must be used. Optical communication between integrated circuits is one possible answer. Light beams do not take up space or interfere with cooling air. If the communication is optical, then the computation might be done optically as well. Optical computation will require a radically different form of integrated circuit, which can in principle be made of gallium arsenide and related III-V compounds. These matters are currently under serious study in research laboratories.

Superconducting electronics
      Numerous metals completely lose their resistance to the flow of electric current at temperatures approaching absolute zero (0 K, −273 °C, or −460 °F) and become superconducting (superconductivity). Other equally dramatic changes in electrical properties occur as well. One of these is the Josephson effect, named for the British physicist Brian D. Josephson (Josephson, Brian D.), who predicted and then discovered the phenomenon in 1962. The Josephson effect governs the passage of current from one superconducting metal to another through a very thin insulating film between them (the Josephson junction) and the effects of small magnetic fields on this current.

      Josephson junction devices change from one electrical state to another in extraordinarily short times, offering the possibility of producing superconducting microcircuits that operate faster than any other kind known. Serious efforts have been made to construct a computer on this basis, but most of the projects have been either discontinued or sharply cut back because of technical difficulties. Interest in the approach has also waned because of increases in the speed of III-V semiconductor microcircuits.

      Josephson junctions have other uses in science. They make extremely sensitive detectors of small magnetic fields, for example. The voltage across a Josephson junction is known on theoretical grounds to be dependent only on the values of certain basic physical constants. Since these constants are known to great accuracy, Josephson junctions are now used to provide the absolute standard of voltage.

      Other important applications of Josephson junctions have to do with the metrology of very high-speed signals. Measurements of fast phenomena require the use of even faster measurement tools, which Josephson devices provide.

Flat-panel displays
      Display devices convey information in visible form from electronic devices to human viewers. Common examples are the faces on digital watches, numerical indicators on stereo equipment, and the picture tubes in television sets and computer monitors. Until recently the most versatile of these has been the picture tube, which can present numbers, letters, graphs, and both still and moving pictures. While picture tubes set a very high standard of performance and provide bright colour images, they are bulky, heavy, and expensive. Designers of television receivers have long desired a display device having the virtues of the picture tube but fewer of the disadvantages, so that a “picture on the wall” television set can be produced.

      New developments in flat-panel displays have made this possible. Such displays are advanced versions of the liquid crystal display familiar in digital watch faces. They are essentially two parallel sheets of thin glass having the facing sides coated with a transparent yet electrically conducting film such as indium tin oxide. The film layer nearer the viewer is patterned, while the other layer is not. The space between the films is filled with a fluid with unusual electrical and optical properties, so that, if an electrical field is established between the two thin films, the molecules of the fluid line up in such a way that the light-reflecting or light-transmitting properties of the assembly are radically changed. The electro-optical fluid is an electrical insulator, so very little electric current flows. Thus, almost no power is consumed, making the display well suited for use in battery-powered applications. All flat-panel displays have these characteristics in common, but the many different varieties exploit the electro-optical effects in numerous ways.

      Displays that produce images are patterned with myriads of tiny picture elements that can be electrically activated independently to produce patterns of light and dark or arbitrary forms. Superposed colour filters having arrays of elements corresponding to those in the display permit the formation of colour images of a quality rivaling that of colour cathode-ray tube displays. Such displays are used as viewing devices for television sets, computers, and video and digital cameras.

       colour displays capable of serving as television screens or computer displays are available in sizes of more than 35 cm (15 inches) on the diagonal, at costs nearly competitive with picture tubes. There is strong demand for them in laptop computers, where the thinness of a flat-panel display is essential. Such displays have more than three million separate elements in the picture array, each of which must have separate means for its control. The control electronics is integrated into the display, for otherwise the number of individual wires needed to connect with the rest of the circuitry would be prohibitive.

      A great amount of effort is being expended to increase the size and decrease the cost of flat-panel displays, because the potential market for them is clearly substantial. Much of the reduction in cost is obtained through experience in manufacturing, where low yields attributable to defects in the patterns have been a major problem.

The science of electronics

Valence electrons
      Since electronics is concerned with the control of the motion of electrons, one must keep in mind that electrons, being negatively charged, are attracted to positive charges and repelled by other negative charges. Thus, electrons in a vacuum tend to space themselves apart from one another and form a cloud, subject to the influences of other charges that may be present. An electric current is created by the motion of electrons, whether in a vacuum, in a wire, or in any other electrically conducting medium. In each of these cases, electrons move as a result of their attraction to positive charges or repulsion from negative ones.

      An atom consists of a nucleus of protons and neutrons around which electrons, equal in number to the protons in the nucleus, travel in orbits much like those of the planets around the Sun. Because of this equality in the number of positively and negatively charged constituent particles, the atom as a whole is electrically uncharged. When atoms are combined into certain solids called covalent solids (notably the elements of column IV of the periodic table), the valence electrons (valence electron) (outer electrons) are shared between neighbouring atoms, and the atoms thereby become bound together. This occurs not only in elemental solids, wherein all the atoms are of the same kind, but also in chemical compounds (e.g., the III-V compounds).

      Different materials vary greatly in their ability to conduct electricity, depending directly on the ease or difficulty of setting electrons free from their atoms. In insulating materials all the outermost electrons of the atoms are tightly bound in the chemical bonds between atoms and are not free to move. In metals (metal) there are more valence electrons than are required for bonding, and these excess electrons are freely available for electrical conduction.

      Most insulators and metals are crystalline (crystal) materials but are composed of a great many very small crystals. (In all crystals the atoms are positioned in a regularly spaced three-dimensional array.) Semiconducting solids for electronic applications, however, are prepared as single large crystals. The fact that the atoms in a semiconductor are arranged in a periodic, three-dimensional array of large size (large, that is, in comparison with an atom) makes the atoms appear nearly invisible to electrons moving within a crystal. The reasons for this behaviour are too complex to explain here, but this property allows electrons to be quite mobile in semiconductors.

Conduction in semiconductors
 In semiconductors such as silicon (which is used as the example here), each constituent atom has four outer electrons, each of which pairs with an electron from one of four neighbouring atoms to form the interatomic bonds. Completely pure silicon thus has essentially no electrons available at room temperature for electronic conduction, making it a very poor conductor. However, if an atom from column V of the periodic table, such as phosphorus, is substituted for an atom of silicon, four of its five outer electrons will be used for bonding, while the fifth will be free to move within the crystal (see figure—>). If the replacement atom comes from column III of the periodic table—say, boron (boron group element)—it will have only three outer electrons, one too few to complete the four interatomic bonds. The fact that the crystal would be electrically neutral were this bond complete means that, if an electron is missing, the vacancy will have a positive charge. A neighbouring electron can move into the vacancy, leaving another vacancy in the electron's former place. This vacancy, with its positive charge, is thus mobile and is called a “hole.” Holes (hole) in semiconductors move about as readily as electrons do, but, because they are positively charged, they move in directions opposite to the motion of electrons.

      Semiconductors whose principal charge carriers are electrons are called n-type (n standing for negative). If the charge carriers are mainly holes, the material is p-type (p for positive). The process of substituting elements for the silicon (in this example) is called doping, while the elements are referred to as dopants (dopant). The amount of dopant that is required in practical devices is very small, ranging from about 100 dopant atoms per million silicon atoms downward to 1 per billion.

Fabrication of semiconductors
 Dopants may be added to the silicon either during the crystal growth process or later. Growth of silicon crystals begins with the preparation of extremely pure polycrystalline silicon having fewer than 1 dopant atom per 10 billion silicon atoms. This silicon is melted in a quartz-lined furnace. The temperature of the molten silicon is reduced to just above the melting point (1,410 °C [2,570 °F]), and a small bar (the seed) of silicon in single-crystal form is introduced into the surface of the melt. The molten silicon freezes slowly onto the seed with a crystalline structure that is continuous with the structure of the seed. The seed is slowly withdrawn, usually while rotating, under carefully controlled conditions, and it brings with it a cylindrical ingot of silicon that is a single crystal throughout. This ingot may be up to 300 mm (12 inches) in diameter and weigh up to 100 kg (220 pounds). (See illustration—> of the Czochralski method of crystal pulling.)

      After growth the silicon crystal is ground to a smooth cylindrical shape and sliced into thin wafers approximately 0.6 mm (0.02 inch) thick using diamond tools. The surfaces of the wafers are polished flat by a series of successively finer abrasives until one side has a perfect mirror finish.

      The process of fabricating semiconductor devices (semiconductor device) is a complex series of more than 600 sequential steps, all of which must be done with utmost precision in an environment cleaner than a hospital operating room. The objective is to add the correct dopants to the silicon in the proper amounts in the right places and to connect the transistors thus produced with thin films of metals separated by other thin films of insulating materials. The scale of lateral dimensions in integrated circuits ranged down to 0.13 μm (0.000005 inch) in 2001 and continues to decrease year by year. A high-power semiconductor device for industrial use, on the other hand, may be so large as to require a slice of silicon measuring well over 125 mm (5 inches) in diameter.

      The importance of having a thorough, detailed understanding of all the physical effects related to materials, fabrication processes, and device structures cannot be overstated.

      The motion of electrons and holes in semiconductors is governed by the theory of quantum mechanics, which was developed during the 1920s and '30s as a much more comprehensive theory of the behaviour of all the elementary particles that make up matter. The electrical and optical effects observed in semiconductor materials, their interactions, and the effects of temperature on them are all understood in nearly complete detail. This understanding not only makes it possible to explain quantitatively what is observed in laboratory experiments but is essential for predicting how new processes and devices work.

      The research necessary to develop such a detailed theoretical and experimental body of knowledge was initiated during the late 1940s and has continued in industrial, university, and government laboratories ever since. It is now possible to design new semiconductor devices to perform in a completely predictable fashion by calculating their performance from theory and from their physical configuration, with the aid of computers.

      The fabrication processes used to make real devices are not as well understood, although much has been learned. Theoretical designs incorporate the assumptions that the materials are entirely pure, that dopants exist only in the proper amounts and distributions, and that the dimensions of structures have the intended values. These assumptions are true in practice only to a limited degree. Major efforts in universities and company laboratories are focused on better understanding these issues and on developing improved computer-based modeling and process-design methods. Large sums of money are spent to provide equipment and manufacturing environments that adequately control each process step and protect the material being processed from contamination.

Basic electronic functions
 Rectification, or conversion of alternating current (AC) to direct current (DC), is mentioned in the section The vacuum tube era (electronics). A diode, or two-terminal device, is required for this process. Semiconductor diodes consist of a crystal, part of which is n-type and part p-type. The boundary between the two parts is called a n (p-n junction) junction (see figure—>). As noted above, there is a population of holes on the p-type side of the junction and a population of electrons on the n-type side. If a negative voltage is applied to the p-type side, implying a positive voltage applied to the n-type side, the holes in the p-type region will be attracted away from the p-n junction, as will the electrons on the n-type side. A region on either side of the p-n junction will be depleted of charge carriers, thus becoming effectively an insulator. In this condition, called reverse bias, only a very small leakage current flows.

 If these voltages are reversed, however, creating a condition called forward bias, the positive voltage on the p-type side will repel holes across the p-n junction; the negative voltage also will repel the electrons on the n-type side. Both holes and electrons will cross the p-n junction in opposing directions, creating an electric current (see figure—>).

      Many details of the motion of holes and electrons have been omitted from this simple description, but the principle seems clear. The p-n junction in a semiconductor diode conducts current with one polarity of applied voltage but not with the other polarity. Typical small diodes will conduct about 0.1 ampere with roughly a 1.5- volt forward bias, and they will withstand 100 or more volts with negligible current flow in the reverse direction. Large industrial diodes can carry up to 5,000 amperes and can block several thousand volts.

Using n-p-n transistors
  A transistor is constructed with two p-n junctions parallel and very close to one another. A typical configuration is the n transistor (see figure—>), which has different levels of doping in the two n-type regions and other features that improve its efficiency; the n-p-n regions correspond to the source (or emitter), gate (or base), and drain (or collector) of the circuit. In normal operation, such as in an amplifier circuit (see figure—>), there are provisions (batteries in this case) for applying a small forward bias to the base-emitter junction and a larger reverse bias to the base-collector junction. Resistors (resistor) are arranged in series with each battery to establish steady-state operating conditions, and an AC (alternating current) signal source is contained in the base lead. When the AC signal source is switched off, the battery in the emitter-base circuit causes a small current to flow through the series resistor and the forward-biased emitter-base junction. This results in excess electrons being present in the p-type base region of the transistor. Many more of these electrons are attracted to the collector region by the strong reverse bias on the collector than are attracted to the base connection. In an average n-p-n transistor, more than 100 electrons pass from the emitter to the collector for each 1 that passes from the emitter to the base.

      When the AC signal source is switched on, the base current is increased and decreased alternately. The collector current varies in the same way but to a hundredfold larger extent; in effect, the signal has been amplified. The varying collector current through the collector series resistor causes a varying voltage drop, which may be used as the signal source for a subsequent amplifying circuit. This example employs an n-p-n transistor. With a p-n-p transistor, the action is similar except that holes are the primary charge carriers, and the voltages of the batteries and thus the direction of current are reversed.

Using MOSFETs
 Another important type of transistor developed by the early 1960s is the field-effect transistor, such as a metal-oxide-semiconductor field-effect transistor, or MOSFET (see figure—>). Another type, the junction field-effect transistor, works in a similar fashion but is much less frequently used. The MOSFET consists of two regions: (1) the source and (2) the drain of one conductivity type embedded in a body of the opposite conductivity type. The space between the source and the drain is covered by a thin layer of silicon dioxide formed by heating the silicon in an oxidizing atmosphere. A third part of the device, the gate, is a thin metal layer deposited on the silicon dioxide.

      There are several types of MOSFETs, including an n-channel type, so designated because, when it is in operation, the application of a positive voltage to the gate with respect to the p-type region causes a thin conducting region containing mostly electrons to form in the p-type region just beneath the gate. The gate voltage repels holes and attracts electrons from the p-type region, in which there are some electrons even though the principal charge carriers are holes. The thin layer of electron-rich material, the channel, connects the source and drain electrically and permits current to flow between them when the drain is biased positively with respect to the source. The amount of current is controlled by the gate voltage. Without gate voltage, no current flows, because the p-n junction around the drain region is reverse-biased and because no channel exists. MOSFETs are widely used in integrated circuits.

Coupling amplifiers
      The existence of more than one type of transistor gives the circuit designer additional freedom not available for vacuum tube circuits and allows many clever circuits to be constructed. This becomes apparent in the direct coupling of successive amplifier stages. There are many ways to couple a signal from one circuit to another. Each has its advantages and disadvantages. Consideration must be given to the voltage levels in the circuits. In cases where the voltage level at the collector of the first amplifier is different from that at the base of the second, a direct connection could not be used. A transformer could be employed for coupling, with its primary in the collector circuit of the first amplifier and its secondary in the base circuit of the second one. However, transformers often do not exhibit uniform behaviour over a wide range of frequencies, which can be a problem. Transformers also are expensive and bulky. Similarly, a capacitor (capacitance) could be inserted between the collector of the first amplifier and the base of the second. This works well for many applications, providing uniform coupling inexpensively over a wide frequency range. At low frequencies, capacitive coupling becomes ineffective, however.

 The use of a p second amplifier allows direct connection between the amplifiers (see figure—>). If properly designed, this arrangement provides useful amplifying properties from DC to quite high frequencies. Care is required to avoid any changes in the DC operating conditions of the first amplifier; such changes will cause an amplified change in the DC conditions of the second one. Changes in temperature, in particular, can cause changes in resistor values and changes in the amplification properties of transistors. These factors must be carefully taken into account. Judicious use of feedback from later parts of a circuit to earlier ones can be utilized to stabilize such circuits or to perform various other useful functions (see below Oscillation (electronics)). In negative feedback, the feedback signal is of a sense opposite to the signal present at the point in the circuit where the feedback signal is applied. While this has the effect of reducing the overall gain of the circuit, it also corrects numerous small distortions that may have occurred in the signal. For example, if the amplifier does not amplify large signals as much as small ones, the feedback from larger signals will be less, as will the reduction in gain, and the larger signals will be increased in the output of the circuit. Thus the distortion is reduced.

      If feedback is positive, the feedback signal reinforces the original one, and an amplifier can be made to oscillate, or generate an AC signal. Such signals are needed for many purposes and are created in numerous kinds of oscillator circuits. In a tunable oscillator, such as that required for a radio receiver, the parallel combination of an inductor and a capacitor is a tuned circuit: at one frequency, and only one, the inductive effects and the capacitive effects balance. At this frequency the voltage developed across the tuned circuit is a maximum. Positive feedback is provided by the inductor in the collector circuit, which is magnetically coupled to the inductor of the tuned circuit. The connections to these inductors are arranged so that, when the collector current increases, the voltage at the base also increases, thus causing the collector current to rise further. The action of the tuned circuit reverses this sequence after a time and causes the base voltage to start to fall. This reduces the collector current; the positive feedback then further reduces the base voltage, and so on.

      The circuit is in fact an amplifier whose output provides the input signal. The tuned circuit affects the feedback process in such a way that the circuit responds to an input signal at only one frequency—namely, the frequency to which the inductor and capacitor are tuned. The variable capacitor provides a way to adjust the frequency of oscillation. The output signal is obtained from the emitter resistor, through which the current rises and falls in synchrony with the collector current.

 Oscillators that produce a single, accurate frequency are often needed. Such an oscillator is used in electronic watches. Other circuits in the watch count the output signals from the oscillator to determine the passage of time. These oscillators use a quartz crystal instead of a tuned circuit to establish the operating frequency (see figure—>).

      Quartz has the useful properties of changing its dimensions slightly if an electric field is applied to it and, conversely, of producing a small electrical voltage when pressure is applied (the piezoelectric effect (piezoelectricity)). In a quartz-crystal oscillator a small plate of quartz is provided with metal electrodes on its faces. Just as a bell rings when struck, the quartz plate also “rings,” but at a very high frequency, and produces an AC voltage between the electrodes at this mechanically resonant frequency. When such a crystal is used in an oscillator, positive feedback provides energy to the quartz crystal to keep it ringing, and the oscillator output frequency is precisely controlled by the quartz crystal.

      Quartz is not the only crystalline material that exhibits a piezoelectric effect, but it is used in this application because its oscillation frequency can be quite insensitive to temperature changes. Quartz-controlled oscillators are able to produce output frequencies from about 10 kilohertz to more than 200 megahertz and, in carefully controlled environments, can have a precision of one part in 100 billion, though one part in 10 million is more common.

switching and timing

Using transistors
      Transistors in amplifier circuits are used as linear devices; i.e., the input signal and the larger output signal are nearly exact replicas of each other. Transistors and other semiconductor devices may also be used as switches. In such applications the base or gate of a transistor, depending on the type of transistor in use, is employed as a control element to switch on or off the current between the emitter and collector or the source and drain. The purpose may be as simple as lighting an indicator lamp, or it may be of a much more complex nature.

      An example of a moderately sophisticated application is in a backup, or “uninterruptible,” power source for a computer. Such equipment consists of a storage battery (battery) (which is normally kept charged by rectifying the power coming from the AC power line), a circuit for converting the battery power into AC, and the necessary control circuits. The control circuits monitor the voltage supplied from the power line. If this voltage varies significantly either upward or downward from its normal values, the control circuit causes the power supply lines to the computer to be switched from the incoming power line to an alternate source of AC derived from the battery.

      Batteries are usually low-voltage DC sources. Consequently, their energy has to be converted to AC and applied to a transformer so as to raise the voltage to the proper level for operating the computer. The conversion from DC to AC, known as inversion, is often done with high-power transistors operated as switches. The battery is connected to the primary coil of the transformer through the transistors, first in one polarity and then in the other, at a frequency identical to the normal power-line frequency—usually 50 or 60 hertz.

      The same result could in principle be obtained by operating the transistors as an oscillator powered by the battery and supplying a smoothly varying AC voltage to the transformer rather than the square pulses obtained via the switching process. This is a much less efficient procedure, however. A transistor operated as a switch is quite efficient, because in its “off” condition very little current flows at a relatively high voltage (a slight leakage through the reverse-biased collector junction), while in the “on” condition the collector-emitter voltage is very low, even though the current is large. In both conditions, the power lost is the product of the voltage and the current. Given this fact, the loss is small, because at any instant either the voltage or the current is small.

Using thyristors
 Thyristors (thyristor) are another important class of semiconductor devices used in switching applications. The simplest of these devices is the controlled rectifier (see figure—>), made of silicon. It may be regarded as two transistors connected to each other.

      The device will start to conduct if a suitable amount of gate current is applied, but otherwise it will not. The gate current is the equivalent of the base current for the n-p-n transistor; the resulting larger collector current is the base current for the p-n-p transistor. The p-n-p transistor has an unusually wide base region, so its gain is small, especially at low currents. Its collector current augments the initial gate current, however. This positive feedback increases the current levels throughout the thyristor, increasing the gain of the p-n-p transistor, and at a certain point the combined currents through the n-p-n and p-n-p transistors are sufficient to maintain conduction through the device even if the gate current is removed. The transistors drive each other into a saturated condition such that the thyristor conducts a large current with a very low voltage drop, typically about one volt. The device remains in this conducting state for an arbitrary period and cannot be turned off under control of the gate. Conduction will cease if the anode polarity becomes negative with respect to the cathode.

      Thyristors are thus well suited for operation in AC (alternating current) rather than DC circuits. They can be switched on during the appropriate half-cycle of voltage (anode positive) and will automatically switch off when the polarity reverses. A single thyristor can be used as a rectifier to produce a variable DC output from a fixed AC input. Adjustment of the DC output is made by modifying the time at which the gate current is applied after the AC voltage crosses zero and becomes the right polarity for conduction. Two thyristors connected in antiparallel (i.e., the anode of each is connected to the cathode of the other) form an AC switch, one thyristor being able to conduct on one half-cycle and the other on the alternate half-cycle. The amount of AC power delivered to the load may be adjusted to any level between zero and full power by appropriate timing of the gate signals to the two thyristors.

      Thyristors are designed to handle both small and large amounts of power; the largest ones can withstand up to 5,000 volts in the “off” state and can conduct up to 2,000 amperes in the “on” state. Such a device is contained in an enclosure approximately 150 mm (6 inches) in diameter and about 30 mm (1 inch) thick fitted with external air- or water-cooling means. The power loss in the thyristor in such cases may be as much as 4 kilowatts, but the total amount of power handled may be up to 1,000 times as large. The efficiency is thus very high.

      Other types of thyristors include those in which the gate is able to turn off the thyristor and those that can be switched on in either direction of current flow. The latter finds wide use in light-duty applications—for example, in variable-speed home appliances and light dimmers.

      Thyristors have many applications in industrial equipment where substantial amounts of power must be controlled electronically. These applications range from transmission of electric power over long distances, which is more efficient if done as DC rather than AC, to control of heating elements in furnaces and supplying power for electronic equipment. The very large thyristors mentioned earlier are employed in power conversion for DC transmission, both from AC to DC and vice versa.

Optoelectronic functions
      Some electronic applications depend on the interactions between light and semiconductor materials mentioned in the section Optoelectronics (electronics). Such applications include the conversion of sunlight to electricity in solar cells (solar cell). Most cells of this type consist of silicon diodes in specially designed enclosures to allow sunlight to illuminate them. Silicon is transparent to infrared light; this component of solar radiation passes through a solar cell without generating electricity. The waves of visible light, however, have enough energy to create hole-electron pairs (the mechanism that results in the absorption of the light). In the vicinity of the p-n junction, the holes are attracted toward the electrons on the n-type side, and the electrons are attracted to the holes on the p-type side. This constitutes a current that can be used to power small electrical appliances or to charge storage batteries.

      There are special thyristors available that use light instead of a gate signal to initiate conduction. They have application in high-voltage systems wherein many thyristors in series must be employed to withstand the voltage. The practical difficulties involved in providing gate signals to all these thyristors, each at a different electrical potential, are simplified by using optical fibres (which are electrical insulators) to conduct pulses of light to the thyristors. The interaction of the light with the silicon produces carriers just as in a solar cell; these carriers provide the gate signal to switch on the thyristors.

      Light-emitting diodes (light-emitting diode) (LEDs) are used in many electronic systems as visual indicators. They are made from III-V compounds related to gallium arsenide; the ones that generate red light are usually composed of gallium arsenide phosphide. The central brake light on the rear of automobiles is commonly an array of red LEDs. The red light in traffic signals is also an LED application. With the availability of brilliant, low-cost blue LEDs, it is now possible to make replacements for incandescent lightbulbs using a suitable mixture of coloured LEDs to provide the appropriate colour. These newer applications are driven by the need for greater reliability or electrical efficiency to justify the increase in cost.

      Laser diodes, also made of III-V compounds, are used in digital audio and video disc players to read the minuscule tracks molded into the disc and containing the digitally recorded information. Lasers are employed because laser light can be focused into an extremely tiny spot of great brightness. The light scattered from the markings on the disc is detected by semiconductor photodiodes.

Robert I. Scace

Additional Reading
Historical developments in electronics are outlined in Henry B.O. Davis, Electrical and Electronic Technologies: A Chronology of Events and Inventors to 1900 (1981), and Electrical and Electronic Technologies: A Chronology of Events and Inventors from 1940 to 1980 (1985); G.W.A. Dummer and E. Davies, Electronic Inventions and Discoveries: Electronics from Its Earliest Beginnings to the Present Day, 4th rev. (1997); and W.A. Atherton and Charles Sussking, From Compass to Computer: A History of Electrical and Electronics Engineering (1984).Fundamental principles and basic functions of electronics are presented in Paul Horowitz and Winfield Hill, The Art of Electronics, 2nd ed. (1989); S.W. Amos and M.R. James, Principles of Transistor Circuits: Introduction to the Design of Amplifiers, Receivers, and Digital Circuits, 9th ed. (2000), an elementary discussion of devices and circuits; Robert J. Matthys, Crystal Oscillator Circuits, rev. ed. (1992), an introductory textbook covering a wide range of oscillators; B. Jayant Baliga, Modern Power Devices (1987, reprinted 1992), a comprehensive textbook on devices for power frequency applications; Robert Boylestad and Louis Nashelsky, Electronic Devices and Circuit Theory, 7th ed. (1998); and James T. Humphries and Leslie P. Sheets, Industrial Electronics, 4th ed. (1993).Robert I. Scace

* * *


Universalium. 2010.

Игры ⚽ Нужно решить контрольную?

Look at other dictionaries:

  • Electronics — refers to the flow of charge (moving electrons) through nonmetal conductors (mainly semiconductors), whereas electrical refers to the flow of charge through metal conductors. For example, flow of charge through silicon, which is not a metal,… …   Wikipedia

  • electronics — e‧lec‧tron‧ics [ɪˌlekˈtrɒnɪks, ˌelɪk ǁ ˈtrɑː ] noun [uncountable] 1. the study of making equipment that works using electricity and computer chip S: • She studied electronics at Manchester University. 2. MANUFACTURING equipment that works using… …   Financial and business terms

  • electronics — n. 1. the branch of physics that deals with the behavior of electrons. Electronics is primarily concerned with phenomena other than simple conduction, such as emission of electrons, storage of electrical charge, the effects of electrical fields… …   The Collaborative International Dictionary of English

  • electronics — 1910, from ELECTRONIC (Cf. electronic); Cf. also ICS (Cf. ics). The science of how electrons behave in vacuums, gas, semi conductors, etc …   Etymology dictionary

  • electronics — [n] electronic devices camcorders, CD players, computer chips, computers, electronic components, integrated circuitry, radios, stereos, televisions, transistors, VCRs, video cameras; concepts 463,499 …   New thesaurus

  • electronics — ► PLURAL NOUN 1) (usu. treated as sing. ) the branch of physics and technology concerned with the behaviour and movement of electrons, especially in semiconductors and gases. 2) (treated as pl. ) circuits or devices using transistors, microchips …   English terms dictionary

  • electronics — [ē΄lek tran′iks; ē lek΄tran′iks, i lek΄tran′iks; ] also [, el΄ek tran′iks] n. the science that deals with the behavior and control of electrons in vacuums and gases, and with the use of electron tubes, photoelectric cells, transistors, etc. pl.n …   English World dictionary

  • electronics — [[t]ɪlektrɒ̱nɪks[/t]] 1) N UNCOUNT Electronics is the technology of using transistors and silicon chips, especially in devices such as radios, televisions, and computers. ...Europe s three main electronics companies. ...cheaper, better consumer… …   English dictionary

  • electronics — I (New American Roget s College Thesaurus) Science of electrons Nouns 1. electronics, electron physics, electronic engineering, solid state electronics; musical instrument digital interface, MIDI; electron theory of atoms or solids;… …   English dictionary for students

  • electronics — e|lec|tron|ics [ıˌlekˈtrɔnıks, ˌelık US ˈtra: ] n 1.) [U] the study or industry of making equipment, such as computers and televisions, that work electronically electronics company/industry/firm etc ▪ an electronics engineer 2.) [plural]… …   Dictionary of contemporary English

Share the article and excerpts

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