electron tube

electron tube
an electronic device that consists, typically, of a sealed glass bulb containing two or more electrodes: used to generate, amplify, and rectify electric oscillations and alternating currents. Also called electronic tube. Cf. gas tube, vacuum tube.
[1920-25]

* * *

Introduction
also called  vacuum tube 

      device usually consisting of a sealed glass or metal-ceramic enclosure that is used in electronic circuitry to control a flow of electrons. Among the common applications of vacuum tubes are amplification (amplifier) of a weak current, rectification (rectifier) of an alternating current (AC) to direct current (DC), generation of oscillating (oscillator) radio-frequency (RF) power for radio and radar, and creation of images on a television screen or computer monitor. Common types of electron tubes include magnetrons (magnetron), klystrons (klystron), gyrotrons, cathode-ray (cathode ray) tubes (such as the thyratron), photoelectric cells (photoelectric cell) (also known as phototubes), and neon and fluorescent lamps (fluorescent lamp).

      Until the late 1950s, vacuum tubes were used in virtually every kind of electronic device—computers, radios, transmitters, components of high-fidelity sound systems, and so on. After World War II the transistor was perfected, and solid-state devices (solid-state device) (based on semiconductors (semiconductor)) came to be used in all applications at low power and low frequency. The common conception at first was that solid-state technology would rapidly render the electron tube obsolete. Such has not been the case, however, for each technology has come to dominate a particular frequency and power range. The higher power levels (hundreds of watts) and frequencies (above 8 gigahertz [GHz]) are dominated by electron tubes and the lower levels by solid-state devices. High power levels have always been required for radio transmitters, radar systems, and implements of electronic warfare, and microwave communications systems may require power levels of hundreds of watts. Power in these cases is frequently provided by klystrons, magnetrons, and traveling-wave tubes. Extremely high average power levels—several megawatts at frequencies above 60 GHz—are achieved by gyrotrons; these are used primarily for deep-space radars, microwave weapons, and drivers for high-energy particle accelerators (particle accelerator).

      Vacuum tube technology continues to advance, because of a combination of device innovation, enhanced understanding through improved mathematical modeling and design, and the introduction of superior materials. The bandwidth over which electron tubes operate has more than doubled since 1990. The efficiency with which DC power is converted to RF power has increased up to 75 percent in some devices. New materials, such as diamond for dielectrics, pyrolitic graphite for collectors, and new rare-earth magnets for beam control, greatly improve the power handling and efficiency of modern electron tubes.

Principles of electron tubes
 An electron tube has two or more electrodes separated either by vacuum (in a vacuum tube) or by ionized gas at low pressure (in a gas tube). Its operation depends on the generation and transfer of electrons through the tube from one electrode to another. The source of electrons is the cathode, usually a metallic electrode that releases a stream of electrons (see figure—>) by one of several mechanisms described below. Once the electrons have been emitted, their movement is controlled by an electric field, a magnetic field, or both. An electric field is established by the application of a voltage between the electrodes in the tube, while a magnetic field may be produced outside the tube by an electromagnet or a permanent magnet. In its simplest form, an electron is attracted and accelerated by the positive electrode (a plate, or anode) and is repelled and slowed by the negative electrode (cathode). An electric field can be used to change the path of the electron stream, alter the number of flowing electrons (change the electric current), and modify their speed. The magnetic field serves primarily to control the movement of the electrons from one electrode to another.

Electron emission
      In its most general sense, the emission of electrons results from directing energy in the form of heat, atomic-scale collisions, or strong electric fields to the cathode in such a way that electrons within the material are given enough kinetic energy to escape the surface. The most widely used mechanism in vacuum tubes is thermionic emission, or electron emission by application of heat.

      The amount of energy needed to release electrons from a given material is known as its electronic work function. It follows that the ideal materials for cathodes are those that yield the lowest electronic work function. barium, strontium, and thorium are commonly used for cathodes because of their low electronic work functions, from 1.2 to 3.5 electron volts (electron volt) (eV). Newer experimental materials, such as scandate (an alloy of barium and scandium oxide), have been discovered with slightly lower electronic work functions.

      The anode, meanwhile, is usually made of a good conductor—such as iron, nickel, or carbon—that does not readily emit electrons at typical operating temperatures.

      When solids are heated to high temperatures—about 1,000 °C (1,800 °F) or higher—electrons can be emitted from the surface. (This phenomenon was first observed by the American inventor Thomas Alva Edison (Edison, Thomas Alva) in 1883 and is known as the Edison effect.) Thermionic emission is not thoroughly understood, but researchers have been able to describe it mathematically, using wave mechanics.

      The most popular models rest on the Richardson-Dushman equation, derived in the 1920s, and the Langmuir-Child equation, formulated shortly thereafter. The former states that the current per unit of area, J, is given by

      where k is Boltzman's constant, A is a constant of the material and its surface finish and is theoretically about 120 amperes per square centimetre per kelvin, T is the temperature of the solid, and W is its work function.

 As electrons are emitted by the application of heat, an electron cloud can form in front of the cathode. Such a cloud acts to repel low-energy electrons, which return to the cathode. This limiting mechanism is aptly referred to as the space-charge-limited operation. In a device such as the diode (see figure—>), the positive voltage applied to the anode attracts electrons from the cloud. The higher the voltage, the more electrons flow to the anode until the saturation voltage has been reached, at which point all the emitted electrons flow to the anode (known as the saturation current). In the space-charge-limited operation, the current density, J, is described by the Langmuir-Child law

      where Va is the anode voltage and d is the distance between the anode and the cathode. The key characteristics of thermionic emission, as observed and predicted by equations (1) and (2), are the temperature-limited region and the space-charge-limited region. Much research has been concerned with the transition between the regions and with decreasing the work function of the cathode materials.

      When a metal or dielectric is bombarded by ions or electrons, electrons within the material may acquire sufficient kinetic energy to be emitted from the surface. The bombarding electrons are called primary, and the emitted electrons are designated secondary. The amount of secondary emission depends on the properties of the material and the energy and angle of incidence of the primary electrons. Material properties are characterized by the secondary-emission ratio, defined as the number of secondary electrons emitted per primary electron. Typically, the maximum secondary-emission ratio lies between 0.5 and 1.5 for pure metals and occurs for incident electron energies between 200 and 1,000 eV. The approximate energy distribution of secondary electrons emitted from a pure metal is skewed in such a way that about 85 percent of them have energies less than 20 eV.

      Positive ion (cation) bombardment also can cause secondary emission, but it is much less efficient than electron bombardment, because only a small fraction of an ion's energy can be imparted to (much lighter) electrons.

      Electron emission is influenced by an electric field applied at the cathode. For very strong electric fields, the electron emission becomes independent of temperature because the potential barrier at the surface of the cathode is made extremely narrow and electrons tunnel through the barrier even when they have low kinetic energy. Electric field strength must be about a billion volts (volt) per metre in order to cause field emissions.

Electron motion in a vacuum
      Fundamental to all electron devices are the dynamics of charged particles under different electric and magnetic fields. The motion of an electron in a uniform field is given by a simple application of Isaac Newton (Newton, Sir Isaac)'s second law of motion, force = mass × acceleration, in which the force is exerted on the electron by an applied electric field E (measured in volts per metre). Mathematically, the equation of motion of an electron in a uniform field is given by

      in which e is the electron charge 1.60 × 10−19 coulombs, E denotes the field in volts per metre, m is the electron mass 9.107 × 10−31 kilogram, and dv/dt denotes the rate of change of velocity, which is the electron's acceleration.

      If a magnetic field is also present, the electron will experience a second force, but only when the electron is in motion. The force will then be proportional to the product of charge and the velocity component that is perpendicular to the electric field E and to the magnetic flux density B (measured in webers per square centimetre). The force will be directed perpendicular to both the electric field and the electron velocity. Thus, an electron traveling parallel to an electric field and at right angles to a uniform magnetic field will be deflected in a direction perpendicular to both magnetic and electric fields. Because the force is constantly perpendicular to the velocity, the electron will trace out a perfectly circular trajectory and will maintain that motion at a rate called the cyclotron frequency, ωc, given by e/mB. The circle traced out by the electron has a radius equal to mv/eB. This circular motion is exploited in many electron devices for generating or amplifying radio-frequency (RF) power.

      An electron traveling parallel to a uniform magnetic field is unaffected by that field, but any departure from parallelism gives rise to a perpendicular component of velocity and thus a force. This force gives the nearly parallel electron a helical motion about the direction of the magnetic field, keeping it from diverging far from the parallel path. The equation of motion in any of these instances is

      where v is the velocity of the electron in metres per second in the perpendicular direction to the plane of B and v, and θ is the angle between the directions of B and v. The magnetic flux density is expressed in webers per square centimetre (1 weber per centimetre2 = 104 gauss = 107/4π amperes per metre).

      Of interest, too, is the situation in which the magnetic and electric fields are perpendicular to each other. This configuration is used in beam-focusing devices as well as in a class of devices called magnetrons (see the section Magnetrons (electron tube)). In this case the motion of the electrons is a combination of translation and circular trajectories. The resultant trajectory is a cycloid.

      Equations (3) and (4) are sufficient to solve for the path and time of transit of electrons in an electron tube except that they require E and B to be known, and these may depend on the presence of electrons or ions. The currents in electron tubes are small enough in most cases that their effect on the magnetic field is usually negligible. The cumulative effect of the electron or ion charge (called space charge) on the electric field cannot always be neglected, however, and this introduces computational difficulty unless the geometry is simple. Furthermore, the electrode currents are so dependent on space charges that the performance characteristics of electron tubes are largely determined by these charges. The electric field with or without space charge can be determined by Gauss's theorem of electrostatics, which states how electric fields are associated with charges. Basically, the rate of change of E with distance is equal to ρ/ε0, in which ρ is the electric charge density in coulombs per metre, and ε0is the permittivity 8.95 × 10−12 farads per metre.

      The current per unit area, i, entering any surface—as that of an electrode in a tube—is the time rate of change of charge at that surface. This current is the sum of two components, one constituting the actual arrival of electrons at the electrode and the other resulting from the change of induced charge by any change of the electric field with time. Thus, i is the sum of ρv + ε0dE/dt, where v is the electron density and dE/dt is the time-varying electric field. At low frequencies of operation or under steady conditions, the second term is not important. The contrary is true at high frequencies. This equation and the one relating the electric fields to the charges are fundamental to all high-vacuum electron tube phenomena and are sufficient to obtain theoretical solutions.

Energy transfer
      The fundamental importance of a large class of electronic devices lies in their ability to amplify power. This power amplification results from the conversion of the energy stored in an external power supply to an output energy in the load circuit of the electron device. The mechanism that makes this conversion possible is the electron's change in kinetic energy as it is accelerated or decelerated by an electric field. Because energy is conserved, the RF field will increase (amplification) if the electrons lose kinetic energy, and, conversely, it will decrease if the electrons gain kinetic energy.

      When a modulated electron convection current flows in an electric field of the same modulation frequency, the power transfer, P, between the field and the electron is given by

      where lc is the electron convection current and E is the electric field. Both lc and E are complex quantities; substituting their values into equation (5) and separating the real and imaginary parts yields

      in which ϕl and ϕE are the phase angles of the modulated convection current and electric field, respectively. Insight into the meaning of equations (6) and (7) may be obtained by considering a physical picture. The negative electron flow (convection current) may be supposed to induce positive charges on the electrodes from which the E field emanates. If the phase is proper, meaning that the induced charges constructively add to the current associated with the modulated E field, the E field grows. Thus, in equations (6) and (7),

and
becomes zero. Conversely, if the phases are 180° apart,
goes to zero, and
and power is transferred from the field to the electron current. In practice, different methods are used to produce density modulation in an electron beam (see below).

Common tubes and their applications
      Many types of electron tubes are involved in RF electric power generation and amplification. Another class of electron tubes is employed for rectification and switching (thyratrons (thyratron) and ignitrons (ignitron)). Some vacuum and gas tubes are designed merely to illuminate a target, as in the case of a television tube. This discussion focuses on those electron tubes that serve as circuit elements, functioning as rectifiers (rectifier), microwave RF sources, and amplifiers (amplifier). Of these, the most important are the latter two types, because they constitute the technology of choice in a wide range of high-power microwave and millimetre-wave applications. Within this category the main varieties are klystrons, magnetrons, crossed-field amplifiers, traveling-wave tubes, gyrotrons, and free-electron lasers. Special applications have given impetus to the development of microwave power sources capable of generating tremendous amounts of power (up to billions of watts). These devices are called fast-wave tubes. Some of these and other significant vacuum tubes are delineated below, as are gas tubes employed for rectification and switching.

Klystrons (klystron)
 Devices of this kind are used as amplifiers and RF signal sources at microwave frequencies (e.g., in radio relay systems and for dielectric heating) and also as oscillators (e.g., in continuous-wave Doppler radar systems). The klystron is a linear beam device; that is, the electron flow is in a straight line focused by an axial magnetic field. The velocities of electrons emitted from the cathode are modulated to produce a density-modulated electron beam. The principle of operation involved here can be explained in terms of a two-cavity klystron amplifier, the basic structure of which is shown in the figure—>.

      The first grid next to the cathode controls the number of electrons in the electron beam and focuses the beam. The voltage between the cathode and the cavity resonators (the buncher and the catcher, which serve as reservoirs of electromagnetic oscillations) is the accelerating potential and is commonly referred to as the beam voltage. This voltage accelerates the DC electron beam to a high velocity before injecting it into the grids of the buncher cavity. The grids of the cavity enable the electrons to pass through, but they confine the magnetic fields within the cavity. The space between the grids is referred to as the interaction space, or gap. When the electrons traverse this space, they are subjected to RF potentials at a frequency determined by the resonant frequency of the buncher cavity and the input-signal frequency. The amplitude of the RF voltage between the grids is determined by the amplitude of the input signal. Electrons traversing the interaction space when the RF potential on grid 3 is positive with respect to grid 2 are accelerated by the field, while those crossing the gap one half-cycle later are decelerated. In this process essentially no energy is taken from the buncher cavity, since the average number of electrons slowed down is equal to the average number of electrons speeded up. The decelerated electrons give up energy to the fields inside the buncher cavity, while those that have been accelerated absorb energy from its fields.

      Upon leaving the interaction gap, the electrons enter a region called the drift, or bunching, space, in which the electrons that were speeded up overtake the slower-moving ones. This causes the electrons to bunch and results in the density modulation of the beam, with the electron bunches representing an RF current in the beam. The catcher is located at a point where the bunching is maximum. This cavity is tuned to the same frequency as the input frequency of the buncher cavity. The power output at the catcher is obtained by slowing down the electron bunches. If an alternating field exists at the cavity and grid 4 is positive with respect to grid 5, the electron bunches passing through the grids will be decelerated, and they will deliver energy to the output cavity. In this way the electron bunches induce an RF current on the walls of the catcher cavity identical to the RF current in the beam. At resonance the oscillation in the cavity builds up in proper phase to retard the electron bunches. The power of the RF output is equal to the difference in the kinetic energy of the electrons averaged before and after passing the interaction gap.

      The positive electrode, or collector, located beyond the catcher collects the electrons; it is designed to minimize secondary emission. (Such emission occurs because of the impact of electrons that reach the end wall.)

      The klystron amplifier described above can be converted into an oscillator by employing feedback from the output cavity to the input cavity in proper phase and of sufficient amplitude to overcome the losses in the system.

      The power levels of klystrons are achieved through the use of large beam voltages and currents. In simple terms, the output power P is given by P = efficiency × IE, where I and E are the beam current and voltage and the efficiency is how well the DC power supplied is converted to RF power. For klystrons the efficiency can be as high as 70 percent. By collecting the spent electron beam at a potential significantly below that of the cavities, even higher efficiency can be achieved—as much as another 10 to 15 percent.

      Klystrons are used in ultrahigh-frequency (UHF) television transmissions, which operate at power levels of less than 50 kilowatts. For ground-based communications, the range of power levels is from 1 to 20 kilowatts. Pulsed klystrons are primarily used in radar and in scientific and medical linear accelerators. Some applications employ more than two cavities to obtain higher gain and more bandwidth. The power gain of the klystron is dependent on the voltage and current as well as on the number of cavities used. The larger the number of cavities employed, the larger the gain that can be obtained. There is, however, a practical limit imposed by the onset of RF instability.

Magnetrons (magnetron)
 Magnetrons are primarily used to generate power at microwave frequencies for radar systems, microwave ovens, plasma screens, linear accelerators, and the creation of plasmas used for such applications as thin-film deposition and ionic etching. Within a magnetron electrons are constrained by the combined effect of a radial electrostatic field and an axial magnetic field. Magnetrons can be manufactured relatively inexpensively because they require so few parts—namely, a cathode, an anode, an antenna, and a magnet. A typical magnetron for microwave ovens, shown in the figure—> in cutaway view, is described below.

      The cylindrical anode structure contains a number of equally spaced cavity resonators with the anode surface adjacent to the cylindrical cathode. Permanent magnets are used to provide the necessary magnetic field, which is perpendicular to the electric field between the cathode and the anode. The power output is coupled through an antenna that runs from one of the cavities to a waveguide that channels the microwave radiation to the cooking chamber.

      As in other types of oscillators, the oscillation originates in random phenomena in the electron space charge and in the cavity resonators. The cavity oscillations produce electric fields that spread outward into the interaction space. Energy is transferred from the radial DC field to the RF field by electrons. The first orbit of an electron occurs when the RF field across the gap is in a direction to retard its motion. The resulting transfer of energy is from the electron to the tangential component of the RF field. After losing energy, the electron is accelerated again by the radial DC field and moves to the next cavity. The electron gives up most of its energy to the cavities before it finally terminates on the anode surface. There is a net delivery of energy to the cavity resonators because electrons that absorb energy from the RF field are quickly returned to the cathode. By contrast, the energy in the rotational component of motion of the electrons in the retarding RF field remains practically unaffected, and the electrons may orbit around the cathode many times.

      Magnetrons have a wide range of output powers—from those used in microwave ovens for cooking, which generate 600 to 1,000 watts, to special ones capable of generating pulsed power levels up to 1,000,000 watts. The DC-to-RF power-conversion efficiency typically ranges from 50 to 85 percent.

Crossed-field amplifiers
      Crossed-field amplifiers (CFA) share several characteristics with magnetrons. Both contain a cylindrical cathode coaxial with an RF structure, and each of these tubes constitutes a diode in which a magnetic field is established perpendicular to an electric field between the cathode and the anode. Another similarity is that their RF structure serves as the electron collector and must therefore be very rugged. The key difference is that CFAs use a delay line to slow down the RF, which thereby allows it to interact more efficiently with the electron stream. Thus, amplification occurs through most of one rotation of the electrons before the signal is extracted into an output waveguide. With this scheme CFAs are capable of achieving very high conversion efficiencies of more than 70 percent. Additionally, the output power of CFAs is obtained with relatively low beam voltage, two to three times lower than other devices at the same power level. The gain characteristic of CFAs is a highly nonlinear one and relatively low (one to two orders of magnitude lower) compared with other electron tubes. Bandwidths of CFAs are typically 10 to 20 percent. The advantages of the CFAs are their high efficiency, small size, and relatively low-voltage operation. They are capable of average power levels from 1 kilowatt at 10 GHz to 1 megawatt at 1 GHz.

Traveling-wave tubes
      These are generally used to amplify microwave signals over broad bandwidths. The main elements of a traveling-wave tube (TWT) are (1) an electron gun, (2) a focusing structure that keeps the electrons in a linear path, (3) an RF circuit that causes RF fields to interact with the electron beam, and (4) a collector with which to collect the electrons. There are two main types of TWTs, and these are differentiated by the RF structure. One uses a slow-wave circuit called a helix for propagating the RF wave for electron-RF field interaction, and the other employs a series of staggered cavities coupled to each other for wave propagation. Each type has different characteristics and finds its use in different applications. The helix TWT is distinct from other electron tubes, as it is the only one that does not use RF cavities. Because cavities have bandwidth limitations, the coupled-cavity TWT also is bandwidth-limited to typically 10 to 20 percent. The helix TWT, however, has no particular bandwidth limitations, and, for all practical purposes, an octave bandwidth (100 percent) is attainable.

 The basic helix TWT is shown schematically in the figure—>. The electron gun contains a cathode that emits electrons, and these are formed by the gun electrodes into a beam that is injected into the opening of the helix.

      Because space-charge forces tend to make the electrons diverge radially, a focusing structure is used to keep the beam at a desired diameter by causing diverging electrons to be sent toward the axis of the helix. In this manner the electron beam is maintained at the desired diameter all along the length of the helix. This is necessary because the electron-RF field interaction takes place continuously over the length of the helix within the helix diameter. In order to achieve this interaction, the diameter and pitch of the helix must be such that the RF wave traveling on the helix wire at the speed of light (about 300,000 km, or 186,000 miles, per second) is slowed down in its axial travel to be in synchrony with the velocity of the electrons in the beam. The axial phase velocity of the wave is approximated by multiplying the speed of light by the ratio of the pitch to the circumference of the helix. The axial phase velocity is relatively constant over a wide range of frequencies, and this characteristic provides for the large bandwidths of helix TWTs. For typical applications the electrons travel down the helix axis at about one-tenth the speed of light. The voltage required to impart this velocity to the electrons is on the order of 10,000 volts. The RF output power and frequency required determine the actual voltage and current to be used.

      The amplifying action of the TWT occurs via a continuous interaction between the axial component of the electric field wave traveling down the centre of the helix and the electron beam moving along the axis of the helix at the same time. The electrons are continually slowed down, and their energy is transferred to the wave along the helix. The electrons tend to bunch in regions where the RF field ahead is decelerating and the field behind is accelerating. The interaction between a bunched electron beam and a helix may be viewed in terms of induced currents. The bunches of electrons induce positive charges on the helix, and these charges move in phase with the bunches. If the phase is proper, this current adds to the current associated with the RF wave flowing in the helix and causes the wave to grow. The interaction is continuous along the length of the helix, which may be up to 25 cm (10 inches) in length. The wave amplitude growing on the helix, in turn, causes the electrons to bunch more, and the growing bunches of electrons result in a continuous exponential growth of the helix wave with distance. Typical gains are on the order of 4 decibels per centimetre, and overall gains are 40 to 60 decibels for helix tubes of practical sizes and applications. After the electron beam has exited the helix, the electrons are decelerated by a multistage collector. By this action a large fraction of the unused beam energy can be recovered via a power supply, which thus increases the overall efficiency of the TWT. The DC-to-RF conversion efficiency of TWTs, both helix and coupled-cavity, is similar and is in the range of 50 to 75 percent, depending on the power level and bandwidth.

      A special application of helix TWTs is their use as amplifiers in communications or scientific satellites and other spacecraft. The helix is ideal for this application because of its small size and weight, high efficiency, and low RF-distortion characteristics. TWTs in space have demonstrated very reliable operation, amassing tens of millions of hours of operation without failure.

Fast-wave electron tubes
      Conventional electron tubes are designed to produce electron-field interaction by slowing down the RF wave to about one-tenth the speed of light. The continuing trend toward high power (more than 1 megawatt at frequencies of 60 GHz and 100 kilowatts at frequencies of 200 GHz) requires vacuum electronic devices, which operate on a different principle from that of the conventional slow-wave electron tubes. The physics of the previously described electron tubes dictates that the size of their RF elements must be in the order of the wavelength of the signal being propagated. Consequently, dimensions and cross sections get extremely small at frequencies above 60 GHz, and the traditional type of tube cannot be made. A different way of creating the electron-field interaction is to allow the RF wave to propagate at essentially the speed of light by letting it pass, for example, through a section of a waveguide. Electrons used for energy transfer to the fast RF wave are bunched either by rippled magnetic fields or by RF fields that induce angular-velocity modulation. The bunched electrons give part of their energy to a properly phased microwave RF field. The advantage of fast-wave devices is that the RF circuits are large compared with the wavelength of a signal. Thus, such devices can be manufactured with large dimensions and still operate at exceedingly high frequencies—e.g., 100 GHz or higher. The fast-wave tubes typically operate at very high voltages to generate the high electron velocities required for resonance conditions, which thereby permits an energy exchange to take place. In fact, it is the resonance due to the electrons in a magnetic field that determines the frequency and not a cavity structure, as in a klystron. The high-voltage AC currents used are the main reason that fast-wave devices produce exceedingly high RF power levels, up to millions of watts at very high frequencies (more than 100 GHz).

 One major type of fast-wave electron tube is the gyrotron (see figure—>). Sometimes called the cyclotron resonance maser, this device can generate megawatts of pulsed RF power at millimetre and submillimetre wavelengths. Gyrotrons make use of an energy-transfer mechanism between an electron orbiting in a magnetic field and an electromagnetic field at the cyclotron frequency. The cyclotron frequency is inversely proportional to the mass of the electron and directly proportional to its velocity and to the strength of the magnetic field. At very high velocities (near the speed of light), the electron increases in mass (because of relativistic effects), and this increase lowers the cyclotron frequency. The interaction between the orbiting electron and the electromagnetic field is such that, if energy is given to the field, the electron loses some mass and the phase of the cyclotron wave changes. This results in a form of electron bunching analogous to the bunching in a klystron (see above Klystrons (electron tube)).

      In another major type of fast-wave tube, an electromagnetic wave (electromagnetic radiation) travels down a circular or rectangular waveguide and interacts with an undulating electron beam. The undulating motion of the electron beam is produced by a periodic magnetic field. The electrons bunch up as in the klystron process. When the bunches interact with the traveling wave, the electron (electron beam) energy is converted to RF energy and results in amplification. Beam voltages in these devices are on the order of 100 kilovolts, and, with electron currents of about 35 amperes, steady-state power levels of 300 watts or pulsed peak power levels of 200 kilowatts can be generated at millimetre wavelengths.

      Gyrotrons and other fast-wave tubes are used in certain high-frequency (35 to 94 GHz) radar applications, in communications systems, for plasma heating in some experimental thermonuclear fusion reactors, and in industrial materials processing.

Gas electron tubes
      In gas tubes the conductivity between the electrodes differs from that of a vacuum because of the presence of a small amount of gas. Common uses of such devices are rectification and switching (e.g., opening inductive energy-storage circuits, on-off modulations, and closing applications). Examples of gas tubes are the thyratron and the ignitron. Some thyratrons can handle up to 50 kilovolts, can switch thousands of amperes, and are capable of handling powers up to 40 megawatts. Thyratrons are used in radar pulse modulators, particle accelerators, and high-voltage medical equipment.

      The modern gas tube is typically a coaxial four-electrode device that contains hydrogen gas at a pressure of 50–400 millitorrs (0.000066–0.00053 atmosphere). A low-voltage discharge is initiated near the cathode by the electrons that it generates, and the hydrogen gas molecules are ionized by collisions with the electrons. The electrons released by the ionized hydrogen bombard the cathode, giving rise to secondary electrons. This secondary electron emission sustains the low-voltage discharge. Some primary and secondary electrons are accelerated from the cathode and undergo more collisions with the hydrogen gas molecules. The plasma formed near the cathode can be enlarged so that contact is made with the electrode serving as the anode, and the conduction plasma path is established. The resulting current can be interrupted by means of a control grid with small apertures that pinch off the flow of plasma.

Additional Reading
Curtis L. Hemenway, Richard W. Henry, and Martin Caulton, Physical Electronics, 2nd ed. (1967), treats the fundamental physics of electron tubes. James T. Coleman, Microwave Devices (1982), provides a general treatment of vacuum devices, including fast-wave tubes. A.S. Gilmour, Jr., Microwave Tubes (1986), is a comprehensive treatment of modern microwave electron tubes. Samuel Y. Liao, Microwave Electron-Tube Devices (1988), gives theoretical and experimental coverage of the basic and newer types of electron tubes. A.S. Gilmour, Jr., Principles of Traveling-Wave Tubes (1994), is a theoretical and experimental treatment of slow-wave electron devices.Estanislao Navarro Sosa

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Electron tube — can be used to describe either of two things: * Vacuum tube * Gas filled tube …   Wikipedia

  • electron tube — n. a sealed tube, usually of glass, containing a vacuum or a gas and electrodes that are connected to pinlike terminals that protrude from the base, used as a plug in component in an electronic device to control the flow of electrons …   English World dictionary

  • Electron Tube —   [engl.], Elektronenröhre …   Universal-Lexikon

  • electron tube — elektroninė lempa statusas T sritis fizika atitikmenys: angl. electron tube; electronic tube; electronic valve vok. Elektronenröhre, f; Katodenröhre, f rus. катодная лампа, f; электронная лампа, f pranc. tube électronique, m …   Fizikos terminų žodynas

  • electron-tube detector — lempinis detektorius statusas T sritis Standartizacija ir metrologija apibrėžtis Detektorius su elektronine lempa. atitikmenys: angl. electron tube detector; electron valve detector; vacuum tube detector vok. Röhrendemodulator, m; Röhrendetektor …   Penkiakalbis aiškinamasis metrologijos terminų žodynas

  • electron tube mixer — lempinis maišiklis statusas T sritis fizika atitikmenys: angl. electron tube mixer; tube mixer vok. Röhrenmischer, m; Röhrenmischstufe, f rus. ламповый смеситель, m pranc. mélangeur à tube, m …   Fizikos terminų žodynas

  • electron-tube detector — lempinis detektorius statusas T sritis fizika atitikmenys: angl. electron tube detector; vacuum tube detector vok. Röhrendemodulator, m; Röhrendetektor, m rus. ламповый детектор, m pranc. détecteur à lampe, m …   Fizikos terminų žodynas

  • electron tube tester — elektroninių lempų tikrintuvas statusas T sritis Standartizacija ir metrologija apibrėžtis Elektroninių lempų parametrų matuoklis. atitikmenys: angl. electron tube tester vok. Röhrenprüfgerät, n rus. прибор для испытания ламп, m pranc. lampemètre …   Penkiakalbis aiškinamasis metrologijos terminų žodynas

  • electron tube — noun electronic device consisting of a system of electrodes arranged in an evacuated glass or metal envelope • Syn: ↑tube, ↑vacuum tube, ↑thermionic vacuum tube, ↑thermionic tube, ↑thermionic valve • Hypernyms: ↑electronic device …   Useful english dictionary

  • electron tube — noun Physics an evacuated or gas filled tube in which a current of electrons flows between electrodes …   English new terms dictionary

Share the article and excerpts

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