You are on page 1of 45

Electromagnetism

Electromagnetism

Electricity Magnetism

[show]Electrostatics [show]Magnetostatics [show]Electrodynamics [show]Electrical Network [show]Covariant formulation [show]Scientists


vde

Electromagnetism is one of the four fundamental interactions of nature, along with strong interaction, weak interaction and gravitation. It is the force that causes the interaction between electrically charged particles; the areas in which this happens are called electromagnetic fields. Electromagnetism is the force responsible for practically all the phenomena encountered in daily life (with the exception of gravity). Ordinary matter takes its form as a result of intermolecular forces between individual molecules in matter. Electromagnetism is also the force which holds electrons and protons together inside atoms, which are the building blocks of molecules. This governs the processes involved in chemistry, which arise from interactions between the electrons orbiting atoms. The force of electromagnetism is manifested both in electric fields and magnetic fields; both are simply different aspects of electromagnetism, and hence are intrinsically related to each other. Thus, a changing electric field generates a magnetic field; conversely a changing magnetic field generates an electric field. This effect is called electromagnetic induction, and is the basis of operation for electrical generators, induction motors, and transformers. Mathematically speaking, magnetic fields and electric fields are convertible with relative motion as a four vector. Electric fields are the cause of several common phenomena, such as electric potential (such as the voltage of a battery) and electric current (such as the flow of electricity through a flashlight). Magnetic fields are the cause of the force associated with magnets. In quantum electrodynamics, electromagnetic interactions between charged particles can be calculated using the method of Feynman diagrams, in which we picture messenger particles

called virtual photons being exchanged between charged particles. This method can be derived from the field picture through perturbation theory.

Permeability (electromagnetism)
From Wikipedia, the free encyclopedia Jump to: navigation, search Magnetic Circuits Conventional Magnetic Circuits

Magnetomotive force Magnetic flux

Magnetic reluctance Phasor Magnetic Circuits Complex reluctance Z Related Concepts

Magnetic permeability Gyrator-capacitor model variables


Magnetic impedance zM Effective resistance rM Magnetic inductivity LM Magnetic capacitivity CM

This box: view talk edit

Simplified comparison of permeabilities for: ferromagnets (f), paramagnets(p), free space(0) and diamagnets (d) In electromagnetism, permeability is the measure of the ability of a material to support the formation of a magnetic field within itself. In other words, it is the degree of magnetization that a material obtains in response to an applied magnetic field. Magnetic permeability is

typically represented by the Greek letter . The term was coined in September, 1885 by Oliver Heaviside. The reciprocal of magnetic permeability is magnetic reluctivity. In SI units, permeability is measured in the henry per metre (H m-1), or newton per ampere squared (N A-2). The permeability constant (0), also known as the magnetic constant or the permeability of free space, is a measure of the amount of resistance encountered when forming a magnetic field in a classical vacuum. The magnetic constant has the exact (defined)[1] value 0 = 4107 1.2566370614...106 Hm-1 or NA-2).

Contents
[hide]

1 Explanation 2 Relative permeability 3 Diamagnetism 4 Paramagnetism 5 Values for some common materials 6 Complex permeability 7 See also 8 References 9 External links

[edit] Explanation
In electromagnetism, the auxiliary magnetic field H represents how a magnetic field B influences the organization of magnetic dipoles in a given medium, including dipole migration and magnetic dipole reorientation. Its relation to permeability is

where the permeability is a scalar if the medium is isotropic or a second rank tensor for an anisotropic medium. In general, permeability is not a constant, as it can vary with the position in the medium, the frequency of the field applied, humidity, temperature, and other parameters. In a nonlinear medium, the permeability can depend on the strength of the magnetic field. Permeability as a function of frequency can take on real or complex values. In ferromagnetic materials, the relationship between B and H exhibits both non-linearity and hysteresis: B is not a single-valued function of H[2], but depends also on the history of the material. For these materials it is sometimes useful to consider the incremental permeability defined as .

This definition is useful in local linearizations of non-linear material behavior, for example in a Newton-Raphson iterative solution scheme that computes the changing saturation of a magnetic circuit. Permeability is the inductance per unit length. In SI units, permeability is measured in henries per metre (Hm-1 = J/(A2m) = N A-2). The auxiliary magnetic field H has dimensions current per unit length and is measured in units of amperes per metre (A m-1). The product H thus has dimensions inductance times current per unit area (HA/m2). But inductance is magnetic flux per unit current, so the product has dimensions magnetic flux per unit area. This is just the magnetic field B, which is measured in webers (volt-seconds) per square-metre (Vs/m2), or teslas (T). B is related to the Lorentz force on a moving charge q: . The charge q is given in coulombs (C), the velocity v in m/s, so that the force F is in newtons (N):

H is related to the magnetic dipole density. A magnetic dipole is a closed circulation of electric current. The dipole moment has dimensions current times area, units ampere square-metre (Am2), and magnitude equal to the current around the loop times the area of the loop.[3] The H field at a distance from a dipole has magnitude proportional to the dipole moment divided by distance cubed[4], which has dimensions current per unit length.

[edit] Relative permeability


Relative permeability, sometimes denoted by the symbol r, is the ratio of the permeability of a specific medium to the permeability of free space given by the magnetic constant :

. In terms of relative permeability, the magnetic susceptibility is: m = r 1. m, a dimensionless quantity, is sometimes called volumetric or bulk susceptibility, to distinguish it from p (magnetic mass or specific susceptibility) and M (molar or molar mass susceptibility).

[edit] Diamagnetism
Main article: Diamagnetism Diamagnetism is the property of an object which causes it to create a magnetic field in opposition of an externally applied magnetic field, thus causing a repulsive effect. Specifically, an external magnetic field alters the orbital velocity of electrons around their nuclei, thus changing the magnetic dipole moment in the direction opposing the external field. Diamagnets are materials with a magnetic permeability less than 0 (a relative permeability less than 1). Consequently, diamagnetism is a form of magnetism that is only exhibited by a substance in the presence of an externally applied magnetic field. It is generally a quite weak effect in most materials, although superconductors exhibit a strong effect.

[edit] Paramagnetism
Main article: Paramagnetism Paramagnetism is a form of magnetism which occurs only in the presence of an externally applied magnetic field. Paramagnetic materials are attracted to magnetic fields, hence have a relative magnetic permeability greater than one (or, equivalently, a positive magnetic susceptibility). The magnetic moment induced by the applied field is linear in the field strength and rather weak. It typically requires a sensitive analytical balance to detect the effect. Unlike ferromagnets, paramagnets do not retain any magnetization in the absence of an externally applied magnetic field, because thermal motion causes the spins to become randomly oriented without it. Thus the total magnetization will drop to zero when the applied field is removed. Even in the presence of the field there is only a small induced magnetization because only a small fraction of the spins will be oriented by the field. This fraction is proportional to the field strength and this explains the linear dependency. The attraction experienced by ferromagnets is non-linear and much stronger, so that it is easily observed, for instance, in magnets on one's refrigerator.

Why bands occur in materials


See also: Conduction band, Valence band, and Bandgap The electrons of a single isolated atom occupy atomic orbitals, which form a discrete set of energy levels. If several atoms are brought together into a molecule, their atomic orbitals split, as in a coupled oscillation. This produces a number of molecular orbitals proportional to the number of atoms. When a large number of atoms (of order 1020 or more) are brought together to form a solid, the number of orbitals becomes exceedingly large, and the difference in energy between them becomes very small. Thus, in solids the levels form continuous bands of energy rather than the discrete energy levels of the atoms in isolation. However, some intervals of energy contain no orbitals, no matter how many atoms are aggregated, forming band gaps.

Within an energy band, energy levels are so numerous as to be a near continuum. First, the separation between energy levels in a solid is comparable with the energy that electrons constantly exchange with phonons (atomic vibrations). Second, it is comparable with the energy uncertainty due to the Heisenberg uncertainty principle, for reasonably long intervals of time. As a result, the separation between energy levels is of no consequence. Several approaches to finding band structure are discussed below.

Basic concepts

Figure 1: Simplified diagram of the electronic band structure of metals, semiconductors, and insulators.

Figure 2: First Brillouin zone of FCC lattice showing symmetry labels

Bulk band structure for Si,Ge,GaAs and InAs generated with tight binding model. Note that Si and Ge are indirect while GaAs and InAs are direct band gap materials. Any solid has a large number of bands. In theory, it can be said to have infinitely many bands (just as an atom has infinitely many energy levels). However, all but a few lie at energies so high that any electron that reaches those energies escapes from the solid. These bands are usually disregarded. Bands have different widths, based upon the properties of the atomic orbitals from which they arise. Also, allowed bands may overlap, producing (for practical purposes) a single large band. Figure 1 shows a simplified picture of the bands in a solid that allows the three major types of materials to be identified: metals, semiconductors and insulators. Metals contain a band that is partly empty and partly filled regardless of temperature. Therefore they have very high conductivity. The lowermost, almost fully occupied band in an insulator or semiconductor, is called the valence band by analogy with the valence electrons of individual atoms. The uppermost, almost unoccupied band is called the conduction band because only when electrons are excited to the conduction band can current flow in these materials. The difference between insulators and semiconductors is only that the forbidden band gap between the valence band and conduction band is larger in an insulator, so that fewer electrons are found there and the electrical conductivity is lower. Because one of the main mechanisms for electrons to be excited to the conduction band is due to thermal energy, the conductivity of semiconductors is strongly dependent on the temperature of the material.

This band gap is one of the most useful aspects of the band structure, as it strongly influences the electrical and optical properties of the material. Electrons can transfer from one band to the other by means of carrier generation and recombination processes. The band gap and defect states created in the band gap by doping can be used to create semiconductor devices such as solar cells, diodes, transistors, laser diodes, and others.

A semiconductor is a material that has an electrical conductivity due to flowing electrons (as opposed to ionic conductivity) which is intermediate in magnitude between that of a conductor and an insulator. This means roughly in the range 103 to 108 siemens per centimeter. Devices made from semiconductor materials are the foundation of modern electronics, including radio, computers, telephones, and many other devices. Semiconductor devices include the various types of transistor, solar cells, many kinds of diodes including the light-emitting diode, the silicon controlled rectifier, and digital and analog integrated circuits. Similarly, semiconductor solar photovoltaic panels directly convert light energy into electrical energy. In a metallic conductor, current is carried by the flow of electrons. In semiconductors, current is often schematized as being carried either by the flow of electrons or by the flow of positively charged "holes" in the electron structure of the material. Actually, however, in both cases only electron movements are involved. Common semiconducting materials are crystalline solids but amorphous and liquid semiconductors are known. These include hydrogenated amorphous silicon and mixtures of arsenic, selenium and tellurium in a variety of proportions. Such compounds share with better known semiconductors intermediate conductivity and a rapid variation of conductivity with temperature, as well as occasional negative resistance. Such disordered materials lack the rigid crystalline structure of conventional semiconductors such as silicon and are generally used in thin film structures, which are less demanding for as concerns the electronic quality of the material and thus are relatively insensitive to impurities and radiation damage. Organic semiconductors, that is, organic materials with properties resembling conventional semiconductors, are also known. Silicon is used to create most semiconductors commercially. Dozens of other materials are used, including germanium, gallium arsenide, and silicon carbide. A pure semiconductor is often called an intrinsic semiconductor. The electronic properties and the conductivity of a semiconductor can be changed in a controlled manner by adding very small quantities of other elements, called dopants, to the intrinsic material. In crystalline silicon typically this is achieved by adding impurities of boron or phosphorus to the melt and then allowing the melt to solidify into the crystal. This process is called "doping".[1]

Basic Semiconductor Physics

Materials can be catagorised into conductors , semiconductors or insulators by their ability to conduct electricity. It is a popular belief that insulators do not conduct electricity because their valence electrons are not free to wander throughout the material. In fact they are free to move around, however, in an insulator there are as many electrons as there are energy levels for them to occupy. If an electron swaps place with another electron no change is made since electrons are indistinguishable. There are higher energy levels, but to promote the electrons to these energy levels requires more energy than is usually practical.

Introduction
Metals conduct electricity easily because the energy levels between the conduction and valence band are closely spaced or there are more energy levels available than there are electrons to fill them so very little energy is required to find new energies for electrons to occupy. The resistivity of a material is a measure of how difficult it is for a current to flow. Semiconductors have a resistivity between 10-4< <108 Ohms m although these are rough limits. The band theory of materials explains qualitatively the difference between these types of materials. Electrons occupy energy levels from the lowest energies upwards. However, some energy levels are forbidden because of the wave like properties of atoms in the material. The allowed energy levels tend to form bands. The highest filled level at T=0 K is known as the valence band . Electrons in the valence band do not participate in the conduction process. The first unfilled level above the valence band is known as the conduction band . In metals, there is no forbidden gap; the conduction band and the valence band overlap, allowing free electrons to participate in the conduction process. Insulators have an energy gap that is far greater than the thermal energy of the electron, while semiconductor materials the energy gap is typically around 1eV. The diagram below shows the differences in metals, semiconductors and insulators in terms of the how the energy bands are separated Elemental semiconductors are semiconductors where each atom is of the same type such as Ge, Si. These atoms are bound together by covalent bonds, so that each atom shares an electron with its nearest neighbour, forming strong bonds. Compound semiconductors are made of two or more elements. Common examples are GaAs or InP. These compound

semiconductors belong to the III-V semiconductors so called because first and second elements can be found in group III and group V of the periodic table respectively. In compound semiconductors, the difference in electro-negativity leads to a combination of covalent and ionic bounding. Ternary semiconductors are formed by the addition of a small quantity of a third element to the mixture, for example Al x Ga 1-x As. The subscript x refers to the alloy content of the material, what proportion of the material is added and what proportion is replace by the alloy material. The addition of alloys to semiconductors can be extended to include quaternary materials such as Ga x In (1-x) As y P (1-y) or GaInNAs and even quinternary materials and even quinternary materials such as GaInNAsSb. Once again, the subscripts denote the proportion elements that constitute the mixture of elements. Alloying semiconductors in this way allows the energy gap and lattice spacing of the crystal to be chosen to suit the application. Intrinsic semiconductors are essentially pure semiconductor material. The semiconductor material structure should contain no impurity atoms. Elemental and compound semiconductors can be intrinsic semiconductors. At room temperature, the thermal energy of the atoms may allow a small number of the electrons to participate in the conduction process. Unlike metals, where the resistance of semiconductor material decreases with temperature. For semiconductors, as the temperature increases, the thermal energy of the valence electrons increases, allowing more of them to breach the energy gap into the conduction band. When an electron gains enough energy to escape the electrostatic attraction of its parent atom, it leaves behind a vacancy which may be filled be another electron. The vacancy produced can be thought of as a second carrier of positive charge. It is known as a hole . As electrons flow through the semiconductor, holes flow in the opposite direction. If there are n free electrons in an intrinsic semiconductor, then there must also be n holes. Holes and electrons created in this way are known as intrinsic charge carriers. The carrier concentration or charge density defines the number of charge carriers per unit volume. This relationship can be expressed as n=p where n is the number of electrons and p the number of holes per unit volume. The variation in the energy gap between different semiconductor materials means that the intrinsic carrier concentration at a given temperature also varies. An extrinsic semiconductor consists can be formed from an intrinsic semiconductor by added impurity atoms to the crystal in a process known as doping . To take the most simple example, consider Silicon. Since Silicon belongs to group IV of the periodic table, it has four valence electrons. In the crystal form, each atom shares an electron with a neighbouring atom. In this state it is an intrinsic semiconductor. B, Al, In, Ga all have three electrons in the valence band. When a small proportion of these atoms, (less than 1 in 10 6 ), is incorporated into the crystal the dopant atom has an insufficient number of bonds to share bonds with the surrounding Silicon atoms. One of the Silicon atoms

has a vacancy for an electron. It creates an a hole that contributes to the conduction process at all temperatures. Dopents that create holes in this manner are known as acceptors. This type of extrinsic semiconductor is known as p-type as it create positive charge carriers. Elements that belong to group V of the periodic table such as As, P, Sb have an extra electron in the valence band. When added as a dopant to intrinsic Silicon, the dopant atom contributes an additional electron to the crystal. Dopants that add electrons to the crystal are known as donors and the semiconductor material is said to be n-type. Doping of compound semiconductors is slightly more complicated. The effect of the dopant atom depends the site occupied by the atom on the lattice . In III-V semiconductors, atoms from group II act as a acceptors when occupying the site of a group III atom, while atoms in group VI act as donors when they replace atoms from group V. Dopant atoms from group IV have the property that they can act as acceptors or donor depending on whether they occupy the site of group III or group V atoms respectively. Such impurities are known as amphoteric impurities

Intrinsic

p-type

n-type

Schematic diagram showing the only the valence electron shell to illustrate intrinsic, p-type and n-type semiconductors. Often, we are interested in transitions that occur near the bottom of the conduction band minimum to valence band maximum; In this case, the is useful to draw the bandstructure energy as a function of position, setting the wavevector k =0. In this representation of the energy bands, the donors and acceptors form levels in the energy gap region. At T=0 K, any free carriers from donors and acceptors are bound to their atoms. So there is no conduction. For non-zero temperatures, the sites can be thermally ionised, releasing carriers in the bands so conduction can occur.

These shallow level impurities are known as hydrogenic impurities . For donor atoms, an electron orbits a lattice site, while for acceptors a hole orbits around a lattice site with residual negative charge. The energy required to ionise these carrier is much less than the binding energy of the hydrogen atom since the effective mass is smaller and the radius of the carrier orbit larger than that of the hydrogen atom. A table of common impurities and their activation energy is to be found in the reference section.

(1)

A rough estimate for the temperature of ionisation is at room temperature. Initially when the temperature is low , excitation from donors and acceptors can be the only source of carriers: in this range the conductivity is extrinsic. In this regime, the doping of the semiconductor determine whether the semiconductor is n-type or p-type. At higher enough temperatures, direct thermal excitation from the valence band to the conduction band swaps the extrinsic density. There is then an equal number of electrons and holes; the conductivity is intrinsic with distinction between n and p.

Band Structure and Effective Mass


The basic description of a semiconductor is its bandstructure, i.e. the variation of energy E with wave-vector k . The most important bands are: Valence band - the last filled energy level at T=0 K

Conduction band - the First unfilled energy level at T=0 K The valence band maximum is at k =0, is known as the gamma point . Where the conduction-band minimum also occurs at k =0, the semiconductor is said to be a direct band semiconductor. At non-zero k =0, the semiconductor is an indirect-band semiconductor . In addition to these two main conduction bands other bands may also be present. In III-V semiconductors, Ge and Si there are 3 valence bands with maxima at k =0. These are the light-hole , heavy-hole and spin-orbit split-off band . The bands in a semiconductor material are approximated by parabolic functions of k close to the bandedges. Conduction-band:

(2)

Valence-band: (3) The expression for the effective mass is found from the dynamics of a wave-packet, which represents a localised particle. The wave packet is a modulation envelope, with a carrier-wave running through it. The packet is made up of a small spread of frequencies around a central value 0 ; these are superimposed on each other. The wave packet moves at the group velocity v g

(4) If an electric field E f is applied, so that the wave packet moves a distance dx in time dt. The change in energy of the wave packet is

(5) This change corresponds to a change dk of the central k value k 0 ; it is given by

(6) We convert this to a time derivative:

(7) But

, so (8)

The equation for the acceleration, can be calculated from (9)

(10)

Substituting for

from our first major result,

(11) Comparing these forms, we see,

(12) The dynamics of the holes is more complicated. It is necessary to consider one unfilled state in the otherwise filled valence band. The result is that the hole mass acts like a particle with positive charge + e and mass m h given by

(13)

The Fermi Level and Intrinsic Semiconductors


Electrons are Fermions , and thus follow Fermi-Dirac distribution function

(14) where is the Fermi Energy often denoted E f or chemical potential in semiconductor physics is the energy at which there would be a fifty percent chance of finding an electron, if all energy levels were allowed. In order to apply the statistics, we need the density of states in the conduction and valence bands. These are derived from the basic principle that the density of states is constant in k-space. In the conduction band the density of states is given by:

(15)

and the valence band,

(16)

where E is measured from the top of the valence band.

(17) The density of electrons in the conduction band is

(18) In the valence band, the probability of a hole is

(19) and can be approximated

(20) A similar calculation yields the hole density

(21) Calculation of the Fermi level given the carrier concentration is useful in the calculation of laser gain, but since the function is not invertable, there is no analytical method for achieving this. However numerous approximations have been forumulated to calculate the Fermi level.

The value of depends on N a and N d . However can be eliminated between (18) and (21) to give the important relation

(22) where N c and N v are the prefactors in (18) and (21) .

(23)

As stated, (22) holds for all T and independent of the values of N a and N d . In the intrinsic region, the extrinsic density is negligible, and then n=p since each electron excited to the conduction band leaves a hole behind it. In the intrinsic region, therefore

(24)

If we substitute into (24) the values of n and p from (18) and (21)

(25) This gives the value of in the intrinsic region, simple manipulation leads to

(26)

That is, is displaced from the middle of the band gap by a temperature dependent term that depends on the ratio of the effective masses.

The Fermi Level and Extrinsic Semiconductors


What happens to with temperature when donors and acceptors are present? The charge neutrality condition governs the numbers of carriers.

(27)

where N a - and N d + are the number of ionised acceptor and donor sites. As the sketch shows, the probability of finding an electron on a donor site. The number of sites that are ionised is:

(28) A similar argument shows that

(29) The four terms in (27) are given in terms of (18) , (29) , (21) and (28) respectively, so can in fact be determined from (19). The general case has to be dealt with numerically; We take the case of n-type doping but with some counter-doping: and (30) at T=0, N a electrons move off donor sites to occupy the acceptor sites. Thus

(31)

The donor sites are partially occupied. This is only possible at T=0 if the Fermi-level is at the donor-site energy:

(32) This will not change for very low temperatures, into (18) gives for (33) It is seen that (29) is definitely a low-temperature result. For p-type doping, the result corresponding to (33) is for (34) The important technical region in the n-type material is the temperature range in which all the donors are ionised and the extrinsic electron density is higher than the intrinsic density. Full ionisation means: , so substitution of the value of

(35) Since N a electrons are required for occupation of the acceptor sites. Comparison of (24) and (7) gives

(36) The corresponding results for p-type doping are

(37)

(38) Note that in this technical region if the counter doping is negligible, , (35) and (37) simplify to or

(39)

(40) which is what we tell the engineers.

Transistor
From Wikipedia, the free encyclopedia Jump to: navigation, search For other uses, see Transistor (disambiguation).

Assorted discrete transistors. Packages in order from top to bottom: TO-3, TO-126, TO-92, SOT-23

A transistor is a semiconductor device used to amplify and switch electronic signals. It is made of a solid piece of semiconductor material, with at least three terminals for connection to an external circuit. A voltage or current applied to one pair of the transistor's terminals changes the current flowing through another pair of terminals. Because the controlled (output) power can be much more than the controlling (input) power, the transistor provides amplification of a signal. Today, some transistors are packaged individually, but many more are found embedded in integrated circuits. The transistor is the fundamental building block of modern electronic devices, and is ubiquitous in modern electronic systems. Following its release in the early 1950s the transistor revolutionised the field of electronics, and paved the way for smaller and cheaper radios, calculators, and computers, amongst other things.

Contents
[hide]

1 History 2 Importance o 2.1 Usage 3 Simplified operation o 3.1 Transistor as a switch o 3.2 Transistor as an amplifier 4 Comparison with vacuum tubes o 4.1 Advantages o 4.2 Limitations 5 Types o 5.1 Bipolar junction transistor o 5.2 Field-effect transistor o 5.3 Other transistor types o 5.4 Part numbers 5.4.1 Other schemes 5.4.2 Naming problems 6 Construction o 6.1 Semiconductor material o 6.2 Packaging 7 See also 8 References 9 Further reading 10 External links o 10.1 Datasheets
o

10.2 Patents

[edit] History
Main article: History of the transistor

A replica of the first working transistor.

Physicist Julius Edgar Lilienfeld filed the first patent for a transistor in Canada in 1925, describing a device similar to a Field Effect Transistor or "FET".[1] However, Lilienfeld did not publish any research articles about his devices,[citation needed] nor did his patent cite any examples of devices actually constructed. In 1934, German inventor Oskar Heil patented a similar device.[2] From 1942 Herbert Matar experimented with so-called Duodiodes while working on a detector for a Doppler RADAR system. The duodiodes built by him had two separate but very close metal contacts on the semiconductor substrate. He discovered effects that could not be explained by two independently operating diodes and thus formed the basic idea for the later point contact transistor. In 1947, John Bardeen and Walter Brattain at AT&T's Bell Labs in the United States observed that when electrical contacts were applied to a crystal of germanium, the output power was larger than the input. Solid State Physics Group leader William Shockley saw the potential in this, and over the next few months worked to greatly expand the knowledge of semiconductors. The term transistor was coined by John R. Pierce.[3] According to physicist/historian Robert Arns, legal papers from the Bell Labs patent show that William Shockley and Gerald Pearson had built operational versions from Lilienfeld's patents, yet they never referenced this work in any of their later research papers or historical articles.[4] The name 'transistor' is a portmanteau of the term 'transfer resistor'.[5] The first silicon transistor was produced by Texas Instruments in 1954.[6] This was the work of Gordon Teal, an expert in growing crystals of high purity, who had previously worked at Bell Labs.[7] The first MOS transistor actually built was by Kahng and Atalla at Bell Labs in 1960.[8]

[edit] Importance
The transistor is the key active component in practically all modern electronics, and is considered by many to be one of the greatest inventions of the twentieth century.[9] Its importance in today's society rests on its ability to be mass produced using a highly automated process (semiconductor device fabrication) that achieves astonishingly low pertransistor costs. Although several companies each produce over a billion individually packaged (known as discrete) transistors every year,[10] the vast majority of transistors now produced are in integrated circuits (often shortened to IC, microchips or simply chips), along with diodes, resistors, capacitors and other electronic components, to produce complete electronic circuits. A logic gate consists of up to about twenty transistors whereas an advanced microprocessor, as of 2009, can use as many as 2.3 billion transistors (MOSFETs).[11] "About 60 million transistors were built this year [2002] ... for [each] man, woman, and child on Earth."[12]

The transistor's low cost, flexibility, and reliability have made it a ubiquitous device. Transistorized mechatronic circuits have replaced electromechanical devices in controlling appliances and machinery. It is often easier and cheaper to use a standard microcontroller and write a computer program to carry out a control function than to design an equivalent mechanical control function.

[edit] Usage
The bipolar junction transistor, or BJT, was the most commonly used transistor in the 1960s and 70s. Even after MOSFETs became widely available, the BJT remained the transistor of choice for many analog circuits such as simple amplifiers because of their greater linearity and ease of manufacture. Desirable properties of MOSFETs, such as their utility in low-power devices, usually in the CMOS configuration, allowed them to capture nearly all market share for digital circuits; more recently MOSFETs have captured most analog and power applications as well, including modern clocked analog circuits, voltage regulators, amplifiers, power transmitters, motor drivers, etc.

[edit] Simplified operation

Simple circuit to show the labels of a bipolar transistor. The essential usefulness of a transistor comes from its ability to use a small signal applied between one pair of its terminals to control a much larger signal at another pair of terminals. This property is called gain. A transistor can control its output in proportion to the input signal; that is, it can act as an amplifier. Alternatively, the transistor can be used to turn current on or off in a circuit as an electrically controlled switch, where the amount of current is determined by other circuit elements. The two types of transistors have slight differences in how they are used in a circuit. A bipolar transistor has terminals labeled base, collector, and emitter. A small current at the base terminal (that is, flowing from the base to the emitter) can control or switch a much larger current between the collector and emitter terminals. For a field-effect transistor, the terminals are labeled gate, source, and drain, and a voltage at the gate can control a current between source and drain.

The image to the right represents a typical bipolar transistor in a circuit. Charge will flow between emitter and collector terminals depending on the current in the base. Since internally the base and emitter connections behave like a semiconductor diode, a voltage drop develops between base and emitter while the base current exists. The amount of this voltage depends on the material the transistor is made from, and is referred to as VBE.

[edit] Transistor as a switch

BJT used as an electronic switch, in grounded-emitter configuration. Transistors are commonly used as electronic switches, for both high power applications including switched-mode power supplies and low power applications such as logic gates. In a grounded-emitter transistor circuit, such as the light-switch circuit shown, as the base voltage rises the base and collector current rise exponentially, and the collector voltage drops because of the collector load resistor. The relevant equations: VRC = ICE RC, the voltage across the load (the lamp with resistance RC) VRC + VCE = VCC, the supply voltage shown as 6V If VCE could fall to 0 (perfect closed switch) then Ic could go no higher than VCC / RC, even with higher base voltage and current. The transistor is then said to be saturated. Hence, values of input voltage can be chosen such that the output is either completely off,[13] or completely on. The transistor is acting as a switch, and this type of operation is common in digital circuits where only "on" and "off" values are relevant.

[edit] Transistor as an amplifier

Amplifier circuit, standard common-emitter configuration. The common-emitter amplifier is designed so that a small change in voltage in (Vin) changes the small current through the base of the transistor and the transistor's current amplification combined with the properties of the circuit mean that small swings in Vin produce large changes in Vout.

Various configurations of single transistor amplifier are possible, with some providing current gain, some voltage gain, and some both. From mobile phones to televisions, vast numbers of products include amplifiers for sound reproduction, radio transmission, and signal processing. The first discrete transistor audio amplifiers barely supplied a few hundred milliwatts, but power and audio fidelity gradually increased as better transistors became available and amplifier architecture evolved. Modern transistor audio amplifiers of up to a few hundred watts are common and relatively inexpensive.

[edit] Comparison with vacuum tubes


Prior to the development of transistors, vacuum (electron) tubes (or in the UK "thermionic valves" or just "valves") were the main active components in electronic equipment.

[edit] Advantages
The key advantages that have allowed transistors to replace their vacuum tube predecessors in most applications are

Small size and minimal weight, allowing the development of miniaturized electronic devices. Highly automated manufacturing processes, resulting in low per-unit cost. Lower possible operating voltages, making transistors suitable for small, batterypowered applications. No warm-up period for cathode heaters required after power application. Lower power dissipation and generally greater energy efficiency. Higher reliability and greater physical ruggedness. Extremely long life. Some transistorized devices have been in service for more than 50 years. Complementary devices available, facilitating the design of complementarysymmetry circuits, something not possible with vacuum tubes. Insensitivity to mechanical shock and vibration, thus avoiding the problem of microphonics in audio applications.

[edit] Limitations

Silicon transistors do not operate at voltages higher than about 1,000 volts (SiC devices can be operated as high as 3,000 volts). In contrast, electron tubes have been developed that can be operated at tens of thousands of volts. High power, high frequency operation, such as that used in over-the-air television broadcasting, is better achieved in electron tubes due to improved electron mobility in a vacuum. Silicon transistors are much more vulnerable than electron tubes to an electromagnetic pulse generated by a high-altitude nuclear explosion.

[edit] Types

PNP

P-channel

NPN

Nchannel

BJT

JFET

BJT and JFET symbols

P-channel

Nchannel

JFET

MOSFET enh

MOSFET dep

JFET and IGFET symbols Transistors are categorized by


Semiconductor material: germanium, silicon, gallium arsenide, silicon carbide, etc. Structure: BJT, JFET, IGFET (MOSFET), IGBT, "other types" Polarity: NPN, PNP (BJTs); N-channel, P-channel (FETs) Maximum power rating: low, medium, high Maximum operating frequency: low, medium, high, radio frequency (RF), microwave (The maximum effective frequency of a transistor is denoted by the term fT, an abbreviation for "frequency of transition". The frequency of transition is the frequency at which the transistor yields unity gain). Application: switch, general purpose, audio, high voltage, super-beta, matched pair Physical packaging: through hole metal, through hole plastic, surface mount, ball grid array, power modules

Amplification factor hfe (transistor beta)[14]

The p-n Junction Diode


Contents

Introduction Conduction in Solids Doped Semiconductors Current Flow in Semiconductors The P-N Junction Junction Diode Behaviour The Diode Equation Summary References

Introduction
The most basic property of a junction diode is that it conducts an electric current in one direction and blocks it in the other. This behaviour arises from the electrical characteristics of a junction, called a p-n junction. fabricated within a semiconductor crystal. The most commonly used semiconductor material is silicon. The junction diode is useful in a wide variety of applications including the rectification of ac signals (producing dc from ac), the detection of radio signals, the conversion of solar power to electricity, and in the generation and detection of light. It also finds use in a variety of electronic circuits as a switch, as a voltage reference or even as a tunable capacitor. The p-n junction is also the basic building block of a host of other electronic devices, of which the most well-known is the junction transistor. For this reason, a study of the properties and behaviour of the p-n junction is important. In this chapter, the conduction of electricity in solids is reviewed first and the conduction properties of semiconductors are then explained. The construction of the p-n junction of an ideal diode is described and an explanation of the operation of the device is presented.

Conduction in solids
All matter consists of atoms. Each atom has electrons orbiting the nucleus. The nucleus contains the same amount of positive charge as the negative charge possessed by the orbiting electrons. The ability of any material to conduct electricity depends primarily on the behaviour of the electrons in the outer orbits. Therefore, it is necessary to review briefly some aspects of solid-state physics. This subject will be dealt with in more detail later in the course.

Conductors
In a metallic conductor such as copper, the atoms are arranged in a regular array called a crystal lattice. The electrons in the outer orbits of each metal atom are only loosely bound to the nucleus. These electrons are not closely associated with any particular atom and are free to move through the crystal lattice. Once an electron has left its orbit around a particular atom, that atom is left with an excess positive charge. The electron-deficient atom is called a positive ion. The electron that is now free to move is called a free electron. The free electrons in a conductor can be visualized as a cloud of electrons surrounding fixed positive ions as shown in Figure 1. At normal temperatures, the ions possess energy and vibrate. Collisions between vibrating ions and free electrons cause the electrons to move in a random manner. Over a long period of time, the net motion of these free electrons is zero. If an electric field is applied to the conductor, the free electrons will acquire additional energy and will tend to move in the direction dictated by the field. There will be a resulting net motion of free electrons. The net motion of charge carriers constitutes an electric current.

Insulators
In an insulator, nearly all electrons are very tightly bound to their respective atoms. There are practically no electrons that are able to move under the influence of an applied electric field. Therefore, an insulator cannot conduct any appreciable electric current under normal conditions.

Semiconductors
A semiconductor, such as silicon, has properties somewhere between those of a conductor and an insulator. The ability of a semiconductor to conduct electricity can be changed dramatically by adding small numbers of a different element to the semiconductor crystal. This process is called doping. Early experiments showed that an electric current through a semiconductor was carried by the flow of positive charges as well as negative charges (electrons).

Doped Semiconductors
A semiconductor crystal is called n-type if the addition of an impurity element results in a large number if free electrons (negative charge carriers) available for conduction. Each impurity atom is called a donor atom since it donates an electron. The electron is free to move and can contribute to an electric current. The positive ion left behind is fixed and cannot take part in conduction (see Figure 2).

A semiconductor crystal can be made p-type by doping it with a different element so that there are a large number of positive charge carriers available for conduction. The positive charge carriers actually correspond to vacancies or deficiencies of electrons in the bonds holding the atoms in the crystal lattice. The positive charges are called holes. These holes can move through the lattice as illustrated in one dimension in Figures 3(a) and 3(b). The dotted lines represent the crystal lattice. Note that the movement of a hole is due to the movement of a bound electron from one bond to another. It is not due to the motion of free electrons. In a p-type semiconductor, most of the mobile charge carriers are holes. A hole moving away from its host impurity atom is equivalent to the atom gaining or accepting an electron into its bonding structure. The host atom gains an excess negative charge and is then called an acceptor ion. This situation is illustrated in Figure 4. Note again that the ions are locked in the crystal lattice and therefore reperesent fixed charges and cannot contribute to current. On the other hand, the holes are mobile charge carriers and can contribute to current flow. Even in a highly doped p-type semiconductor there will always be some free electrons. This very small number of free electrons have been omitted in Figure 4 for clarity. Similarly, n-type semiconductors always contain some holes. The predominant mobile charge carriers are called majority carriers, whilst those in the minority are called minority carriers. For example, the majority carriers in ntype material are free electrons. The terms majority carriers and minority carriers have meaning only if the type of semiconductor (n- or p-) is specified. A pure or undoped semiconductor is said to be intrinsic. Such material has equal numbers of holes and free electrons. These carriers are produced as a result of thermal agitation of the atoms, even at room temperature. Some bound electrons can acquire sufficient energy to escape from their atoms, becoming free electrons and leaving holes behind., This process of producing hole-electron pairs is called thermal generation. It is possible for a free electron and a hole to come near each other in the course of their random wandering through the crystal. The free electron can then occupy the vacant position represented by the hole. The hole and electron are said to recombine. There is then no mobile charge carrier at that point. The rate of recombination depends upon the number of carriers present. Thermal generation and recombination occurs in both doped and undoped semiconductor material. When a semiconductor material is in thermal

equilibrium, the rate of generation of hole-electron pairs equals the rate of recombination. The density or concentration of both holes and electrons then remains constant.

Current flow in Semiconductors


An electric current can flow through a semiconductor as a result of the movement of holes and/or free electrons. There are two important processes that account for current flow in semiconductors. These processes are called drift and diffusion.

Drift
Applying an electric field across a semiconductor will cause holes and free electrons to drift through the crystal in the directions shown in Figure 5. The total current is equal to the sum of hole current (to the right) and electron current (tpo the left).

Diffusion
A drop of ink in a glass of water diffuses through the water until it is evenly distributed. The same process, called diffusion, occurs with semiconductors. For example, if some extra free electrons are introduced into a p-type semiconductor, the free electrons will redistribute themselves so that the concentration is more uniform. In the example shown in Figure 6, the free electrons will tend to move to the right. This net motion of charge carriers constitutes a diffusion current. (In Figure 6 the holes and acceptor ions are omitted for clarity.) In this example, the free electrons move away from the region of highest concentration. The higher the localized concentration, the greater will be the rate at which electrons move away. The same process applies to holes in an n-type semiconductor. Note that when a few minority carriers (such as the electrons in Figure 6) are diffusing through a sample, they will encounter a large number of majority carriers. Some recombination will occur. A number of both types of carrier will be lost.

The p-n Junction


Imagine that a p-type block of silicon can be placed in perfect contact with an ntype block. Free electrons from the n-type region will diffuse across the junction to the p-type side where they will recombine with some of the many holes in the p-type material. Similarly, holes will diffuse across the junction in the opposite direction and recombine, as shown in Figure 7.

The recombination of free electrons and holes in the vicinity of the junction leaves a narrow region on either side of the junction that contains no mobile charge. This narrow region which has been depleted of mibile charge is called the depletion layer. It extends into both the p-type and n-type regions as shown in Figure 8(a). Note that the diffusion of holes from the p-type side of the depletion layer leaves behind some uncovered fixed negative charges (the acceptor ions). Similarly, fixed positive charges (donor ions) are uncovered on the n-type side of the depletion layer. There is then a separation of charges: negative fixed charges on the p-type side of the depletion layer and positive fixed charges on the n-type side. This separation of charges causes an electric field to extend across the depletion layer. A potential difference must therefore exist across the depletion layer. The variation of potential with distance is shown in Figure 8(b). The uncovered charges give rise to a built-in potential of V_i_ volts. For a typical silicon p-n junction, V_i_ &ap; 0.6 to 0.7 volts. It varies with doping levels and temperature. The significance of this built-in potential is that it opposes the flow of holes and electrons across the junction. For this reason, the built-in potential is called a potential barrier or potential hill. In practice, a p-n junction is formed within a single crystal rather than simply joining two pieces together. Electrical contacts on either side of the crystal enable connection to an external circuit. The resulting device is called a junction diode.

Junction Diode Behaviour


The most important property of a junction diode is its ability to pass an electric current in one direction only. If the diode is connected to a simple circuit consisting of a battery and a resistor, the battery can be connected in either of two ways as shown in Figures 9(a) and 9(b). When the p-type region of the p-n junction is connected to the positive terminal of the battery, current will flow. The diode is said to be under forward bias. However, when the battery terminals are reversed, the p-n junction almost completely blocks the current flow. This is called reverse bias. If the diode is not connected at all, it is said to be open-circuited and of course no current can flow through the diode.

Forward bias
The application of a forward bias voltage V to a junction diode reduces the builtin potential from V_i_ to V_i_ - V, as shown in Figure 10. The reduction in the built-in potential is due to the applied voltage forcing more electrons into the n-type region and more holes into the p-type region, thus covering some of the fixed charges and narrowing the depletion layer. Since the

total uncovered charge is reduced, the built-in potential must be lower. Remembering that the built-in potential opposes the flow of majority carriers across the junction, a reduction in that potential makes it easier for holes in the ptype region to cross the junction and for electrons in the n-type region to cross the junction in the opposite direction. As the forward bias voltage is increased, the current through the junction becomes greater. When the applied voltage V approaches V_i_, the potential hill is almost removed. There is then little opposition to the flow of carriers across the junction and a large current can flow through the diode. The variation of diode current with voltage under forward bias is shown in the first quadrant of a typical junction diode current-voltage characteristic shown in Figure 11.

Reverse Bias
The application of a reverse voltage V_R_ extracts holes from the p-type region and free electrons from the n-type region and so uncovers more bound charges near the junction, as shown in Figure 12. The depletion layer therefore widens and the height of the potential hill is increased to (V_i_ + V_R_ ) volts. Majority carriers are thereby firther inhibited from crossing the junction. As the reverse voltage is increased, the current is reduced to almost zero. However, a very small reverse current does flow. This reverse saturation current depends only on the thermal generation of holes and electrons near the junction, not on the height of the potential barrier. In practice, this reverse saturation current is quite small but it increases with increasing temperature.

Junction Breakdown
The large increase in reverse current evident in Figure 11 is the result of junction breakdown. It occurs when the reverse voltage reaches a critical value.

The diode equation


A complete analysis of the abrupt p-n junction shows that that current I varies exponentially with applied voltage V. The exact relationship between current and voltage is given by the diode equation I = I_S_ (e^qV/kT^ - 1 ) where I_S_ is the saturation current, qis the electronic charge, k is Boltzmann's constant and T is temperature (\(deK). The diode equation applies for both forward and reverse bias. At extremes of high forward bias and large reverse voltages the behaviour of practical devices deviate from the above diode law.

The diode equation is a very important description of diode behaviour and is the basis for the mathematical description of the behaviour of many other electronic devices which employ p-n junctions. Exercise: Derive an expression for the slope of the current-voltage characteristic of a forward-biased junction diode, and state how this slope can be related to the resistance of the diode.

Summary
1. Semiconductors contain two types of mobile charge carriers, holes and electrons. The holes are positively charged, the electrons negatively charged. 2. A semiconductor may be doped with donor impurities (n-type doping) so that it contains mobile charges which are primarily electrons. 3. A semiconductor may be doped with acceptor impurities (p-type doping) so that it contains mobile charges which are mainly holes. 4. There are two important mechanisms for current flow in a semiconductor: 5. o diffusion of carriers as a result of a concentration gradient; and o drift of carriers in an electric field. 6. At equilibrium, a built-in potential or potential barrier of V_i_ volts is developed across a p-n junction. 7. With the application of a forward bias voltage V, the built-in potential is reduced to V_i_ - V, and current flows through the diode. 8. With the application of a reverse bias voltage V_R_, the height of the potential barrier is increased to V_i_ + V_R_ and little current can flow. 9. The total diode current I is related to the applied voltage V by 10. I = I_S_ ( e^{ q V / k T }^ - 1 ) where I_S_ is the reverse saturation current.

References
A.J. Diefenderfer, Principles of Electronic Instrumentation, Saunders, 1979 (Chapter 5). H.V. Malmstadt et al., Electronic Measurements for Scientists, Benjamin, Menlo Park, 1974 (Section 2-1).

What is a diode?
Tale of a Diode Outline of a Diode FAQ

Tale of a Diode
Below we have compiled a simple explanation of the history and principles behind diodes so that you can have a background of what we think is common knowledge. If you feel it's necessary, give it a glance to refresh your memory. "Well I already knew that!" is what some of you are thinking, and if that is case, feel free to skip this section. 1. Before the vacuum tube... Rectifying properties and the Edison effect were discovered in the two-pole vacuum tube 1884. But, actually eight years before that, the rectifying effect of selenium was discovered in 1876. As you can see, the history of using the properties of semiconductors to create diodes that possess rectifying effects is very old. It may be hard to believe, but the history of semiconductors dates back before that of the vacuum tube! 2. From germanium to silicon The first primitive diodes, like the selenium rectifier or crystal detectors, used iron pyrites and galena and other natural copper oxides (polycrystalline semiconductors). After that, as refining techniques advanced, we entered the age where highly sensitive single crystalline semiconductors could be produced with reliability-the age of germanium and silicon. Later we learned that germanium has a low resistance to heat, and so nowadays, most all semiconductors are made with silicon. 3. Rectification from the pn junction The diode element consists of a structure called the pn junction. The terminal attached to the p-type semiconductor is called the anode, and the terminal attached to the n-type semiconductor is called the cathode. Current is allowed to flow from the anode to the cathode, but almost completely prevented from flowing in the reverse direction. This phenomenon is called rectification, and, put simply, it converts alternating current to a unidirectional current. Diode model

Diode schematic electrical symbol

4. In other words, a diode is a valve!

If you intuitively conjure up an image for the effects of a diode, you might call it a "valve"-a valve for electrical current. If you think of electrical current as flowing water, the anode could be considered the upstream side and the cathode the downstream side. Water flows from upstream to downstream (or shall I say, the electrical current does), but the "valve" prevents it from flowing from downstream to upstream. This is the principle operation of a diode.

The valve is open and electricity flows (forward direction)

The valve is closed and electricity does not flow (reverse direction)

5. The many types of junctions

The current types of junctions in today's diodes can be divided into two main classifications: the pn junction and the Schottky barrier junction. The first one is a semiconductorto-semiconductor junction, and this type of junction can be further classified into diffusion type junctions and mesatype junctions. The latter one uses the effects caused between a semiconductor and a metal and therefore is not actually called a junction in terms of diodes. However, to make things easier to understand, it will be considered a junction here. Currently, the Schottky barrier diode is known for its low power requirements and high speeds, and ROHM is making great advances with its Schottky barrier diode series. A diode has two electrodes: the anode and the cathode. The anode is the (+) terminal and the cathode is the (-) terminal. The characteristics of the diode when current is

6. Forward bias characteristics and reverse bias characteristics

flowing from the anode to the cathode are called the forward bias characteristics and VF and IF are examples of these characteristics. Conversely, if a (-) voltage is applied to the anode and a (+) voltage is applied to the cathode, current is prevented from flowing through the diode. The characteristics at this time are called the reverse bias characteristics and VR and IR are examples of these characteristics.

Types of transistor
There are two types of standard transistors, NPN and PNP, with different circuit symbols. The letters refer to the layers of semiconductor material used to make the transistor. Most transistors used today are NPN because this is the easiest type to make from silicon. This page is mostly about NPN transistors and if you are new to electronics it is best to start by learning how to use these first. The leads are labelled base (B), collector (C) and emitter (E).
These terms refer to the internal operation of a transistor but they are not much help in understanding how a transistor is used, so just treat them as labels!

Transistor circuit symbols

A Darlington pair is two transistors connected together to give a very high current gain. In addition to standard (bipolar junction) transistors, there are field-effect transistors which are usually referred to as FETs. They have different circuit symbols and properties and they are not (yet) covered by this page.

Transistor currents
The diagram shows the two current paths through a transistor. You can build this circuit with two standard 5mm red LEDs and any general purpose low power NPN transistor (BC108, BC182 or BC548 for example). The small base current controls the larger collector current. When the switch is closed a small current flows into the base (B) of the transistor. It is just enough to make LED B glow dimly. The transistor amplifies this small current to allow a larger current to flow through from its collector (C) to its emitter (E). This collector current is large enough to make LED C light brightly. When the switch is open no base current flows, so the transistor switches off the collector current. Both LEDs are off. A transistor amplifies current and can be used as a switch.
This arrangement where the emitter (E) is in the controlling circuit (base current) and in the controlled circuit (collector current) is called common emitter mode. It is the most widely used arrangement for transistors so it is the one to learn first.

Functional model of an NPN transistor


The operation of a transistor is difficult to explain and understand in terms of its internal structure. It is more helpful to use this functional model:

The base-emitter junction behaves like a diode. A base current IB flows only when the voltage VBE across the base-emitter junction is 0.7V or more. The small base current IB controls the large collector current Ic. Ic = hFE IB (unless the transistor is full on and saturated) hFE is the current gain (strictly the DC current gain), a typical value for hFE is 100 (it has no units because it is a ratio) The collector-emitter resistance RCE is controlled by the base current IB: o IB = 0 RCE = infinity transistor off o IB small RCE reduced transistor partly on o IB increased RCE = 0 transistor full on ('saturated')

Additional notes:

A resistor is often needed in series with the base connection to limit the base current IB and prevent the transistor being damaged. Transistors have a maximum collector current Ic rating. The current gain hFE can vary widely, even for transistors of the same type!

A transistor that is full on (with RCE = 0) is said to be 'saturated'. When a transistor is saturated the collectoremitter voltage VCE is reduced to almost 0V. When a transistor is saturated the collector current Ic is determined by the supply voltage and the external resistance in the collector circuit, not by the transistor's current gain. As a result the ratio Ic/IB for a saturated transistor is less than the current gain hFE. The emitter current IE = Ic + IB, but Ic is much larger than IB, so roughly IE = Ic.

There is a table showing technical data for some popular transistors on the transistors page.

Touch switch circuit

Darlington pair
This is two transistors connected together so that the current amplified by the first is amplified further by the second transistor. The overall current gain is equal to the two individual gains multiplied together: Darlington pair current gain, hFE = hFE1 hFE2 (hFE1 and hFE2 are the gains of the individual transistors) This gives the Darlington pair a very high current gain, such as 10000, so that only a tiny base current is required to make the pair switch on. A Darlington pair behaves like a single transistor with a very high current gain. It has three leads (B, C and E) which are equivalent to the leads of a standard individual transistor. To turn on there must be 0.7V across both the base-emitter junctions which are connected in series inside the Darlington pair, therefore it requires 1.4V to turn on. Darlington pairs are available as complete packages but you can make up your own from two transistors; TR1 can be a low power type, but normally TR2 will need to be high power. The maximum collector current Ic(max) for the pair is the same as Ic(max) for TR2. A Darlington pair is sufficiently sensitive to respond to the small current passed by your skin and it can be used to make a touch-switch as shown in the diagram. For this circuit

which just lights an LED the two transistors can be any general purpose low power transistors. The 100k resistor protects the transistors if the contacts are linked with a piece of wire.

Using a transistor as a switch


When a transistor is used as a switch it must be either OFF or fully ON. In the fully ON state the voltage VCE across the transistor is almost zero and the transistor is said to be saturated because it cannot pass any more collector current Ic. The output device switched by the transistor is usually called the 'load'. The power developed in a switching transistor is very small:

In the OFF state: power = Ic VCE, but Ic = 0, so the power is zero. In the full ON state: power = Ic VCE, but VCE = 0 (almost), so the power is very small.

This means that the transistor should not become hot in use and you do not need to consider its maximum power rating. The important ratings in switching circuits are the maximum collector current Ic(max) and the minimum current gain hFE(min). The transistor's voltage ratings may be ignored unless you are using a supply voltage of more than about 15V. There is a table showing technical data for some popular transistors on the transistors page. For information about the operation of a transistor please see the functional model above.

Protection diode
If the load is a motor, relay or solenoid (or any other device with a coil) a diode must be connected across the load to protect the transistor from the brief high voltage produced when the load is switched off. The diagram shows how a protection diode is connected 'backwards' across the load, in this case a relay coil.
Current flowing through a coil creates a magnetic field which collapses suddenly when the current is switched off. The sudden collapse of the magnetic field induces a brief high voltage across the coil which is very likely to damage transistors and ICs. The protection diode allows the induced voltage to drive a brief current through the coil (and diode) so the magnetic field dies away quickly rather than instantly. This prevents the induced voltage becoming high enough to cause damage to transistors and ICs.

When to use a relay


Transistors cannot switch AC or high voltages (such as mains electricity) and they are not usually a good choice for switching large currents (> 5A). In these cases a relay will be needed, but note that a low power transistor may still be needed to switch the current for the relay's coil! Advantages of relays:

Relays Relays can switch AC and DC, transistors can only switch DC. Photographs Rapid Electronics Relays can switch high voltages, transistors cannot. Relays are a better choice for switching large currents (> 5A). Relays can switch many contacts at once.

Disadvantages of relays:
Relays are bulkier than transistors for switching small currents. Relays cannot switch rapidly, transistors can switch many times per second. Relays use more power due to the current flowing through their coil. Relays require more current than many ICs can provide, so a low power transistor may be needed to switch the current for the relay's coil.

Connecting a transistor to the output from an IC


Most ICs cannot supply large output currents so it may be necessary to use a transistor to switch the larger current required for output devices such as lamps, motors and relays. The 555 timer IC is unusual because it can supply a relatively large current of up to 200mA which is sufficient for some output devices such as low current lamps, buzzers and many relay coils without needing to use a transistor. A transistor can also be used to enable an IC connected to a low voltage supply (such as 5V) to switch the current for an output device with a separate higher voltage supply (such as 12V). The two power supplies must be linked, normally this is done by linking their 0V connections. In this case you should use an NPN transistor. A resistor RB is required to limit the current flowing into the base of the transistor and prevent it being damaged. However, RB must be sufficiently low to ensure that the transistor is thoroughly saturated to prevent it overheating, this is particularly important if the transistor is switching a large current (> 100mA). A safe rule is to make the base current IB about five times larger than the value which should just saturate the transistor.

Choosing a suitable NPN transistor


The circuit diagram shows how to connect an NPN transistor, this will switch on the load when the IC output is high. If you need the opposite action, with the load switched on when the IC output is low (0V) please see the circuit for a PNP transistor below. The procedure below explains how to choose a suitable switching transistor. 1. The transistor's maximum collector current Ic(max) must be greater than the load current Ic.
load current Ic = supply voltage Vs load resistance RL

NPN transistor switch


(load is on when IC output is high)

Using units in calculations


Remember to use V, A and or V, mA and k . For more details please see the Ohm's Law page.

2. The transistor's minimum current gain hFE(min) must be at least five times the load current Ic divided by the maximum output current from the IC.
hFE(min) > 5 load current Ic max. IC current

3. Choose a transistor which meets these requirements and make a note of its properties: Ic(max) and hFE(min).
There is a table showing technical data for some popular transistors on the transistors page.

4. Calculate an approximate value for the base resistor:


RB =

5.

Vc hFE where Vc = IC supply voltage 5 Ic (in a simple circuit with one supply this is Vs) For a simple circuit where the IC and the load share the same power supply (Vc = Vs) you may prefer to use: RB = 0.2 RL hFE

6. Then choose the nearest standard value for the base resistor. 7. Finally, remember that if the load is a motor or relay coil a protection diode is required.

Example
The output from a 4000 series CMOS IC is required to operate a relay with a 100 coil. The supply voltage is 6V for both the IC and load. The IC can supply a maximum current of 5mA.

1. Load current = Vs/RL = 6/100 = 0.06A = 60mA, so transistor must have Ic(max) > 60mA. 2. The maximum current from the IC is 5mA, so transistor must have hFE(min) > 60
(5 60mA/5mA).

3. Choose general purpose low power transistor BC182 with Ic(max) = 100mA and hFE(min) = 100. 4. RB = 0.2 RL hFE = 0.2 100 100 = 2000 . so choose RB = 1k8 or 2k2. 5. The relay coil requires a protection diode.

Choosing a suitable PNP transistor


The circuit diagram shows how to connect a PNP

PNP transistor switch


(load is on when IC output is low)

transistor, this will switch on the load when the IC output is low (0V). If you need the opposite action, with the load switched on when the IC output is high please see the circuit for an NPN transistor above.
LED lights when the LDR is dark

The procedure for choosing a suitable PNP transistor is exactly the same as that for an NPN transistor described above.

LED lights when the LDR is bright

Using a transistor switch with sensors


The top circuit diagram shows an LDR (light sensor) connected so that the LED lights when the LDR is in darkness. The variable resistor adjusts the brightness at which the transistor switches on and off. Any general purpose low power transistor can be used in this circuit. The 10k fixed resistor protects the transistor from excessive base current (which will destroy it) when the variable resistor is reduced to zero. To make this circuit switch at a suitable brightness you may need to experiment with different values for the fixed resistor, but it must not be less than 1k . If the transistor is switching a load with a coil, such as a motor or relay, remember to add a protection diode across the load.

The switching action can be inverted, so the LED lights when the LDR is brightly lit, by swapping the LDR and variable resistor. In this case the fixed resistor can be omitted because the LDR resistance cannot be reduced to zero. Note that the switching action of this circuit is not particularly good because there will be an intermediate brightness when the transistor will be partly on (not saturated). In this state the transistor is in danger of overheating unless it is switching a small current. There is no problem with the small LED current, but the larger current for a lamp, motor or relay is likely to cause overheating. Other sensors, such as a thermistor, can be used with this circuit, but they may require a different variable resistor. You can calculate an approximate value for the variable resistor (Rv) by using a multimeter to find the minimum and maximum values of the sensor's resistance (Rmin and Rmax): Variable resistor, Rv = square root of (Rmin Rmax)
For example an LDR: Rmin = 100 , Rmax = 1M , so Rv = square root of (100 1M) = 10k .

You can make a much better switching circuit with sensors connected to a suitable IC (chip). The switching action will be much sharper with no partly on state.

A transistor inverter (NOT gate)


Inverters (NOT gates) are available on logic ICs but if you only require one inverter it is usually better to use this circuit. The output signal (voltage) is the inverse of the input signal:

When the input is high (+Vs) the output is low (0V). When the input is low (0V) the output is high (+Vs).

Any general purpose low power NPN transistor can be used. For general use RB = 10k and RC = 1k , then the inverter output can be connected to a device with an input impedance (resistance) of at least 10k such as a logic IC or a 555 timer (trigger and reset inputs). If you are connecting the inverter to a CMOS logic IC input (very high impedance) you can increase RB to 100k and RC to 10k , this will reduce the current used by the inverter. Next Page: Analogue and Digital Systems | Studying Electronics

Electronics Club Home Page

Site Map Example Projects Construction of Projects Soldering Guide Study Electronics Electronic Components 555 Timer Circuit Symbols Frequently Asked Questions Links to other Electronics sites

John Hewes 2010, The Electronics Club, www.kpsec.freeuk.com

You might also like