Sei sulla pagina 1di 39

.

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 or switch electronic signals and electrical power. It
is composed of semiconductor material usually with at least three terminals for connection to an external
circuit. A voltage or current applied to one pair of the transistor's terminals controls the current through
another pair of terminals. Because the controlled (output) power can be higher than the controlling (input)
power, a transistor can amplify 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. Julius Edgar Lilienfeld patented a field-effect transistor in 1926 but it was not possible
[1]

to actually construct a working device at that time. The first practically implemented device was a point-
contact transistor invented in 1947 by American physicists John Bardeen, Walter Brattain, and William
Shockley. The transistor revolutionized the field of electronics, and paved the way for smaller and
cheaper radios, calculators, and computers, among other things. The transistor is on the list of IEEE
milestones in electronics, and Bardeen, Brattain, and Shockley shared the 1956 Nobel Prize in Physics for
[2]

their achievement. [3]

Most transistors are made from very pure silicon or germanium, but certain other semiconductor materials
can also be used. A transistor may have only one kind of charge carrier, in a field effect transistor, or may
have two kinds of charge carriers in bipolar junction transistor devices. Compared with the vacuum tube,
transistors are generally smaller, and require less power to operate. Certain vacuum tubes have advantages
over transistors at very high operating frequencies or high operating voltages. Many types of transistors are
made to standardized specifications by multiple manufacturers.

History
A replica of the first working transistor.

The thermionic triode, a vacuum tube invented in 1907, enabled amplified radio technology and long-
distance telephony. The triode, however, was a fragile device that consumed a substantial amount of
power. In 1909 physicist William Eccles discovered the crystal diode oscillator. German physicist Julius
[4]

Edgar Lilienfeld filed a patent for a field-effect transistor(FET) in Canada in 1925, which was intended to
be a solid-state replacement for the triode. Lilienfeld also filed identical patents in the United States in
[5][6]

1926 and 1928. However, Lilienfeld did not publish any research articles about his devices nor did his
[7] [8][9]

patents cite any specific examples of a working prototype. Because the production of high-quality
semiconductor materials was still decades away, Lilienfeld's solid-state amplifier ideas would not have
found practical use in the 1920s and 1930s, even if such a device had been built. In 1934, German
[10]

inventor Oskar Heil patented a similar device in Europe. [11]

John Bardeen, William Shockleyand Walter Brattain at Bell Labs, 1948.

From November 17, 1947, to December 23, 1947, John Bardeen and Walter Brattain at AT&T's Bell
Labs in Murray Hill, New Jersey of the United States performed experiments and observed that when two
gold point contacts were applied to a crystal of germanium, a signal was produced with the output power
greater than the input. Solid State Physics Group leader William Shockley saw the potential in this, and
[12]

over the next few months worked to greatly expand the knowledge of semiconductors. The
term transistor was coined by John R. Pierce as a contraction of the term transresistance. According to
[13][14][15]

Lillian Hoddeson and Vicki Daitch, authors of a biography of John Bardeen, Shockley had proposed that
Bell Labs' first patent for a transistor should be based on the field-effect and that he be named as the
inventor. Having unearthed Lilienfeld’s patents that went into obscurity years earlier, lawyers at Bell Labs
advised against Shockley's proposal because the idea of a field-effect transistor that used an electric field
as a "grid" was not new. Instead, what Bardeen, Brattain, and Shockley invented in 1947 was the
first point-contact transistor. In acknowledgement of this accomplishment, Shockley, Bardeen, and
[10]

Brattain were jointly awarded the 1956 Nobel Prize in Physics "for their researches on semiconductors and
their discovery of the transistor effect". [16][17]
Herbert F. Mataré (1950)

In 1948, the point-contact transistor was independently invented by German physicists Herbert
Mataré and Heinrich Welker while working at the Compagnie des Freins et Signaux,
a Westinghouse subsidiary located in Paris. Mataré had previous experience in developingcrystal
rectifiers from silicon and germanium in the German radar effort during World War II. Using this
knowledge, he began researching the phenomenon of "interference" in 1947. By June 1948, witnessing
currents flowing through point-contacts, Mataré produced consistent results using samples of germanium
produced by Welker, similar to what Bardeen and Brattain had accomplished earlier in December 1947.
Realizing that Bell Labs' scientists had already invented the transistor before them, the company rushed to
get its "transistron" into production for amplified use in France's telephone network.
[18]

The first bipolar junction transistors were invented by Bell Labs' William Shockley, which applied for
patent (2,569,347) on June 26, 1948. On April 12, 1950, Bell Labs chemists Gordon Teal and Morgan
Sparks had successfully produced a working bipolar NPN junction amplifying germanium transistor. Bell
Labs had made this new "sandwich" transistor discovery announcement, in a press release on July 4,
1951. [19][20]
Philco surface-barrier transistor developed and produced in 1953

The first high-frequency transistor was the surface-barrier germanium transistor developed by Philco in
1953, capable of operating up to60 MHz. These were made by etching depressions into an N-type
[21]

germanium base from both sides with jets of Indium(III) sulfate until it was a few ten-thousandths of an
inch thick. Indium electroplated into the depressions formed the collector and emitter.[22][23]

The first "prototype" pocket transistor radio was shown by INTERMETALL (a company founded
by Herbert Mataré in 1952) at the Internationale Funkausstellung Düsseldorf between August 29, 1953
and September 9, 1953. [24]

The first "production" pocket transistor radio was the Regency TR-1, released in October 1954 . Produced [17]

as a joint venture between the Regency Division of Industrial Development Engineering Associates,
I.D.E.A. and Texas Instruments of Dallas Texas, the TR-1 was manufactured in Indianapolis, Indiana. It
was a near pocket-sized radio featuring 4 transistors and one germanium diode. The industrial design was
outsourced to the Chicago firm of Painter, Teague and Petertil. It was initially released in one of four
different colours: black, bone white, red, and gray. Other colours were to shortly follow. [25][26][27]

The first "production" all-transistor car radio was developed by Chrysler and Philco corporations and it
was announced in the April 28th 1955 edition of the Wall Street Journal. Chrysler had made the all-
transistor car radio, Mopar model 914HR, available as an option starting in fall 1955 for its new line of
1956 Chrysler and Imperial cars, which first hit the dealership showroom floors on October 21, 1955. [28][29][30]

The first working silicon transistor was developed at Bell Labs on January 26, 1954 by Morris Tanenbaum.
The first commercial silicon transistor was produced by Texas Instruments in 1954. This was the work
of Gordon Teal, an expert in growing crystals of high purity, who had previously worked at Bell
Labs. The first MOSFET actually built was by Kahng and Atalla at Bell Labs in 1960.
[31][32][33] [34]

Importance[edit]
A Darlington transistor opened up so the actual transistor chip (the small square) can be seen inside. A Darlington transistor is effectively two

transistors on the same chip. One transistor is much larger than the other, but both are large in comparison to transistors in large-scale

integration because this particular example is intended for power applications.

The transistor is the key active component in practically all modern electronics. Many consider it to be one
of the greatest inventions of the 20th century. Its importance in today's society rests on its ability to
[35]

be mass-produced using a highly automated process (semiconductor device fabrication) that achieves
astonishingly low per-transistor costs. The invention of the first transistor at Bell Labs was named an IEEE
Milestone in 2009. [36]

Although several companies each produce over a billion individually packaged (known as discrete)
transistors every year, the vast majority of transistors are now produced in integrated circuits (often
[37]

shortened to IC, microchips or simply chips), along withdiodes, 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 3 billion transistors
(MOSFETs). "About 60 million transistors were built in 2002… for [each] man, woman, and child on
[38]

Earth." [39]

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 standardmicrocontroller and write a computer program to
carry out a control function than to design an equivalent mechanical system to control that same function.

Simplified operation[edit]
This section does not cite any sources. Please help improve this section by adding citations to
reliable sources. Unsourced material may be challenged and removed. (November 2010) (Learn how and
when to remove this template message)
A simple circuit diagram to show the labels of a n–p–n 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. It can produce a stronger output signal, a voltage or current, which is proportional to a weaker
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.
There are two types of transistors, which 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 between the base and 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 represents a typical bipolar transistor in a circuit. Charge will flow between emitter and
collector terminals depending on the current in the base. Because 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 V . BE

Transistor as a switch[edit]

BJT used as an electronic switch, in grounded-emitter configuration.


Transistors are commonly used in digital circuits as electronic switches which can be either in an "on" or
"off" state, both for high-power applications such as switched-mode power supplies and for low-power
applications such as logic gates. Important parameters for this application include the current switched, the
voltage handled, and the switching speed, characterised by the rise and fall times.
In a grounded-emitter transistor circuit, such as the light-switch circuit shown, as the base voltage rises, the
emitter and collector currents rise exponentially. The collector voltage drops because of reduced resistance
from collector to emitter. If the voltage difference between the collector and emitter were zero (or near
zero), the collector current would be limited only by the load resistance (light bulb) and the supply voltage.
This is called saturation because current is flowing from collector to emitter freely. When saturated, the
switch is said to be on. [40]

Providing sufficient base drive current is a key problem in the use of bipolar transistors as switches. The
transistor provides current gain, allowing a relatively large current in the collector to be switched by a
much smaller current into the base terminal. The ratio of these currents varies depending on the type of
transistor, and even for a particular type, varies depending on the collector current. In the example light-
switch circuit shown, the resistor is chosen to provide enough base current to ensure the transistor will be
saturated.
In a switching circuit, the idea is to simulate, as near as possible, the ideal switch having the properties of
open circuit when off, short circuit when on, and an instantaneous transition between the two states.
Parameters are chosen such that the "off" output is limited to leakage currents too small to affect connected
circuitry; the resistance of the transistor in the "on" state is too small to affect circuitry; and the transition
between the two states is fast enough not to have a detrimental effect.
Transistor as an amplifier[edit]

Amplifier circuit, common-emitter configuration with a voltage-divider bias circuit.

The common-emitter amplifier is designed so that a small change in voltage (V ) changes the small current
in

through the base of the transistor; the transistor's current amplification combined with the properties of the
circuit means that small swings inV produce large changes in V .
in out

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.

Comparison with vacuum tubes[edit]


Before transistors were developed, vacuum (electron) tubes (or in the UK "thermionic valves" or just
"valves") were the main active components in electronic equipment.
Advantages[edit]
The key advantages that have allowed transistors to replace vacuum tubes in most applications are

 no cathode heater (which produces the characteristic orange glow of tubes), reducing power
consumption, eliminating delay as tube heaters warm up, and immune from cathode poisoning and
depletion;
 very small size and weight, reducing equipment size;
 large numbers of extremely small transistors can be manufactured as a single integrated circuit;
 low operating voltages compatible with batteries of only a few cells;
 circuits with greater energy efficiency are usually possible. For low-power applications (e.g., voltage
amplification) in particular, energy consumption can be very much less than for tubes;
 complementary devices available, providing design flexibility including complementary-
symmetry circuits, not possible with vacuum tubes;
 very low sensitivity to mechanical shock and vibration, providing physical ruggedness and virtually
eliminating shock-induced spurious signals (e.g., microphonics in audio applications);
 not susceptible to breakage of a glass envelope, leakage, outgassing, and other physical damage.
Limitations[edit]
Transistors have the following limitations:

 silicon transistors can age and fail;


[41]

 high-power, high-frequency operation, such as that used in over-the-air television broadcasting, is


better achieved in vacuum tubes due to improved electron mobility in a vacuum;
 solid-state devices are susceptible to damage from very brief electrical and thermal events,
including electrostatic discharge in handling; vacuum tubes are electrically much more rugged;
 sensitivity to radiation and cosmic rays (special radiation-hardened chips are used for spacecraft
devices);
 vacuum tubes in audio applications create significant lower-harmonic distortion, the so-called tube
sound, which some people prefer. [42]

Types[edit]

PNP P-channel

NPN N-channel
BJT JFET

BJT and JFET symbols

P-channel

N-channel

JFET MOSFET enh MOSFET dep

JFET and MOSFET symbols

Transistors are categorized by

 semiconductor material: the metalloids germanium (first used in 1947) and silicon (first used in
1954)—in amorphous,polycrystalline and monocrystalline form—, the compounds gallium
arsenide (1966) and silicon carbide (1997), thealloy silicon-germanium (1989), the allotrope of
carbon graphene (research ongoing since 2004), etc. (seeSemiconductor material);
 structure: BJT, JFET, IGFET (MOSFET), insulated-gate bipolar transistor, "other types";
 electrical polarity (positive and negative): n–p–n, p–n–p (BJTs), n-channel, p-channel (FETs);
 maximum power rating: low, medium, high;
 maximum operating frequency: low, medium, high, radio (RF), microwave frequency (the maximum
effective frequency of a transistor in a common-emitter or common-source circuit is denoted by the
term f , an abbreviation fortransition frequency—the frequency of transition is the frequency at which
T

the transistor yields unity voltage 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 (see Packaging);
 amplification factor h , β (transistor beta) or g (transconductance).
FE F
[43]
m

Hence, a particular transistor may be described as silicon, surface-mount, BJT, n–p–n, low-power, high-
frequency switch.
A popular way to remember which symbol represents which type of transistor is to look at the arrow and
how it is arranged. Within an NPN transistor symbol, the arrow will Not Point iN. Conversely, within the
PNP symbol you see that the arrow Points iN Proudly.
Bipolar junction transistor (BJT)[edit]
Main article: Bipolar junction transistor
Bipolar transistors are so named because they conduct by using both majority and minoritycarriers. The
bipolar junction transistor, the first type of transistor to be mass-produced, is a combination of two junction
diodes, and is formed of either a thin layer of p-type semiconductor sandwiched between two n-type
semiconductors (an n–p–n transistor), or a thin layer of n-type semiconductor sandwiched between two p-
type semiconductors (a p–n–p transistor). This construction produces two p–n junctions: a base–emitter
junction and a base–collector junction, separated by a thin region of semiconductor known as the base
region (two junction diodes wired together without sharing an intervening semiconducting region will not
make a transistor).
BJTs have three terminals, corresponding to the three layers of semiconductor—an emitter, a base, and
a collector. They are useful in amplifiers because the currents at the emitter and collector are controllable
by a relatively small base current. In an n–p–n transistor operating in the active region, the emitter–base
[44]

junction is forward biased (electrons and holes recombine at the junction), and the base-collector junction
is reverse biased (electrons and holes are formed at, and move away from the junction), and electrons are
injected into the base region. Because the base is narrow, most of these electrons will diffuse into the
reverse-biased base–collector junction and be swept into the collector; perhaps one-hundredth of the
electrons will recombine in the base, which is the dominant mechanism in the base current. By controlling
the number of electrons that can leave the base, the number of electrons entering the collector can be
controlled. Collector current is approximately β (common-emitter current gain) times the base current. It
[44]

is typically greater than 100 for small-signal transistors but can be smaller in transistors designed for high-
power applications.
Unlike the field-effect transistor (see below), the BJT is a low-input-impedance device. Also, as the base–
emitter voltage (V ) is increased the base–emitter current and hence the collector–emitter current (I )
BE CE

increase exponentially according to the Shockley diode model and the Ebers-Moll model. Because of this
exponential relationship, the BJT has a higher transconductance than the FET.
Bipolar transistors can be made to conduct by exposure to light, because absorption of photons in the base
region generates a photocurrent that acts as a base current; the collector current is approximately β times
the photocurrent. Devices designed for this purpose have a transparent window in the package and are
called phototransistors.
Field-effect transistor (FET)[edit]
Main articles: Field-effect transistor, MOSFET, and JFET

Operation of a FET and its Id-Vg curve. At first, when no gate voltage is applied. There is no inversion electron in the channel, the device is OFF. As

gate voltage increase, inversion electron density in the channel increase, current increase, the device turns on.

The field-effect transistor, sometimes called a unipolar transistor, uses either electrons (in n-channel FET)
or holes (in p-channel FET) for conduction. The four terminals of the FET are named source, gate, drain,
andbody (substrate). On most FETs, the body is connected to the source inside the package, and this will
be assumed for the following description.
In a FET, the drain-to-source current flows via a conducting channel that connects the source region to
thedrain region. The conductivity is varied by the electric field that is produced when a voltage is applied
between the gate and source terminals; hence the current flowing between the drain and source is
controlled by the voltage applied between the gate and source. As the gate–source voltage (V ) is GS

increased, the drain–source current (I ) increases exponentially for V below threshold, and then at a
DS GS

roughly quadratic rate (I ∝ (V − V ) ) (where V is the threshold voltage at which drain current
DS GS T
2
T

begins) in the "space-charge-limited" region above threshold. A quadratic behavior is not observed in
[45]

modern devices, for example, at the 65 nm technology node. [46]

For low noise at narrow bandwidth the higher input resistance of the FET is advantageous.
FETs are divided into two families: junction FET (JFET) and insulated gate FET (IGFET). The IGFET is
more commonly known as a metal–oxide–semiconductor FET(MOSFET), reflecting its original
construction from layers of metal (the gate), oxide (the insulation), and semiconductor. Unlike IGFETs, the
JFET gate forms a p–n diode with the channel which lies between the source and drain. Functionally, this
makes the n-channel JFET the solid-state equivalent of the vacuum tube triode which, similarly, forms a
diode between its grid and cathode. Also, both devices operate in the depletion mode, they both have a
high input impedance, and they both conduct current under the control of an input voltage.
Metal–semiconductor FETs (MESFETs) are JFETs in which the reverse biased p–n junction is replaced by
a metal–semiconductor junction. These, and the HEMTs (high-electron-mobility transistors, or HFETs), in
which a two-dimensional electron gas with very high carrier mobility is used for charge transport, are
especially suitable for use at very high frequencies (microwave frequencies; several GHz).
FETs are further divided into depletion-mode and enhancement-mode types, depending on whether the
channel is turned on or off with zero gate-to-source voltage. For enhancement mode, the channel is off at
zero bias, and a gate potential can "enhance" the conduction. For the depletion mode, the channel is on at
zero bias, and a gate potential (of the opposite polarity) can "deplete" the channel, reducing conduction.
For either mode, a more positive gate voltage corresponds to a higher current for n-channel devices and a
lower current for p-channel devices. Nearly all JFETs are depletion-mode because the diode junctions
would forward bias and conduct if they were enhancement-mode devices; most IGFETs are enhancement-
mode types.
Usage of bipolar and field-effect transistors[edit]
The bipolar junction transistor (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 amplifiers because of their greater linearity and ease of manufacture. In integrated circuits,
the desirable properties of MOSFETs allowed them to capture nearly all market share for digital circuits.
Discrete MOSFETs can be applied in transistor applications, including analog circuits, voltage regulators,
amplifiers, power transmitters and motor drivers.
Other transistor types[edit]

Transistor symbol created on Portuguese pavement in the University of Aveiro.

For early bipolar transistors, see Bipolar junction transistor § Bipolar transistors.

 Bipolar junction transistor (BJT):


 heterojunction bipolar transistor, up to several hundred GHz, common in modern ultrafast and RF
circuits;
 Schottky transistor;
 avalanche transistor:
 Darlington transistors are two BJTs connected together to provide a high current gain equal to the
product of the current gains of the two transistors;
 insulated-gate bipolar transistors (IGBTs) use a medium-power IGFET, similarly connected to a
power BJT, to give a high input impedance. Power diodes are often connected between certain
terminals depending on specific use. IGBTs are particularly suitable for heavy-duty industrial
applications. The ASEA Brown Boveri (ABB) 5SNA2400E170100 , , intended for three-phase
[47]

power supplies, houses three n–p–n IGBTs in a case measuring 38 by 140 by 190 mm and
weighing 1.5 kg. Each IGBT is rated at 1,700 volts and can handle 2,400 amperes;
 Phototransistor.
 multiple-emitter transistor, used in transistor–transistor logic and integrated current mirrors;
 multiple-base transistor, used to amplify very-low-level signals in noisy environments such as the
pickup of a record player or radio front ends. Effectively, it is a very large number of transistors
in parallel where, at the output, the signal is added constructively, but random noise is added
only stochastically.
[48]

 Field-effect transistor (FET):


 carbon nanotube field-effect transistor (CNFET), where the channel material is replaced by a
carbon nanotube;
 junction gate field-effect transistor (JFET), where the gate is insulated by a reverse-biased p–n
junction;
 metal–semiconductor field-effect transistor (MESFET), similar to JFET with a Schottky junction
instead of a p–n junction;
 high-electron-mobility transistor (HEMT);
 metal–oxide–semiconductor field-effect transistor (MOSFET), where the gate is insulated by a
shallow layer of insulator;
 inverted-T field-effect transistor (ITFET);
 fin field-effect transistor (FinFET), source/drain region shapes fins on the silicon surface;
 fast-reverse epitaxial diode field-effect transistor (FREDFET);
 thin-film transistor, in LCDs;
 organic field-effect transistor (OFET), in which the semiconductor is an organic compound;
 ballistic transistor (disambiguation);
 floating-gate transistor, for non-volatile storage;
 FETs used to sense environment;
 ion-sensitive field-effect transistor (IFSET), to measure ion concentrations in solution,
 electrolyte–oxide–semiconductor field-effect transistor (EOSFET), neurochip,
 deoxyribonucleic acid field-effect transistor (DNAFET).
 Tunnel field-effect transistor, where it switches by modulating quantum tunnelling through a barrier.
 Diffusion transistor, formed by diffusing dopants into semiconductor substrate; can be both BJT and
FET.
 Unijunction transistor, can be used as simple pulse generators. It comprise a main body of either P-
type or N-type semiconductor with ohmic contacts at each end (terminalsBase1 and Base2). A
junction with the opposite semiconductor type is formed at a point along the length of the body for the
third terminal (Emitter).
 Single-electron transistors (SET), consist of a gate island between two tunneling junctions. The
tunneling current is controlled by a voltage applied to the gate through a capacitor.
[49]

 Nanofluidic transistor, controls the movement of ions through sub-microscopic, water-filled


channels.[50]

 Multigate devices:
 tetrode transistor;
 pentode transistor;
 trigate transistor (prototype by Intel);
 dual-gate field-effect transistors have a single channel with two gates in cascode; a configuration
optimized for high-frequency amplifiers, mixers, and oscillators.
 Junctionless nanowire transistor (JNT), uses a simple nanowire of silicon surrounded by an
electrically isolated "wedding ring" that acts to gate the flow of electrons through the wire.
 Vacuum-channel transistor, when in 2012, NASA and the National Nanofab Center in South Korea
were reported to have built a prototype vacuum-channel transistor in only 150 nanometers in size, can
be manufactured cheaply using standard silicon semiconductor processing, can operate at high speeds
even in hostile environments, and could consume just as much power as a standard transistor. [51]

 Organic electrochemical transistor.


 A Solaristor (from solar cell transistor) is a two-terminal gate-less self-powered phototransistor.
Part numbering standards/specifications[edit]
The types of some transistors can be parsed from the part number. There are three major semiconductor
naming standards; in each the alphanumeric prefix provides clues to type of the device.
Japanese Industrial Standard (JIS)[edit]
JIS transistor prefix table

Prefix Type of transistor

2SA high-frequency p–n–p BJT

2SB audio-frequency p–n–p BJT

2SC high-frequency n–p–n BJT

2SD audio-frequency n–p–n BJT

2SJ P-channel FET (both JFET and MOSFET)

2SK N-channel FET (both JFET and MOSFET)

The JIS-C-7012 specification for transistor part numbers starts with "2S", e.g. 2SD965, but sometimes
[52]

the "2S" prefix is not marked on the package – a 2SD965 might only be marked "D965"; a 2SC1815 might
be listed by a supplier as simply "C1815". This series sometimes has suffixes (such as "R", "O", "BL",
standing for "red", "orange", "blue", etc.) to denote variants, such as tighter h (gain) groupings.
FE

European Electronic Component Manufacturers Association (EECA)[edit]


The Pro Electron standard, the European Electronic Component Manufacturers Association part
numbering scheme, begins with two letters: the first gives the semiconductor type (A for germanium, B for
silicon, and C for materials like GaAs); the second letter denotes the intended use (A for diode, C for
general-purpose transistor, etc.). A 3-digit sequence number (or one letter then two digits, for industrial
types) follows. With early devices this indicated the case type. Suffixes may be used, with a letter (e.g. "C"
often means high h , such as in: BC549C ) or other codes may follow to show gain (e.g. BC327-25) or
FE
[53]

voltage rating (e.g. BUK854-800A ). The more common prefixes are:


[54]

Pro Electron / EECA transistor prefix table

Prefix class Type and usage Example Equivalent Reference

AC Germanium small-signal AF transistor AC126 NTE102A Datasheet

AD Germanium AF power transistor AD133 NTE179 Datasheet

AF Germanium small-signal RF transistor AF117 NTE160 Datasheet

AL Germanium RF power transistor ALZ10 NTE100 Datasheet


AS Germanium switching transistor ASY28 NTE101 Datasheet

AU Germanium power switching transistor AU103 NTE127 Datasheet

BC Silicon, small-signal transistor ("general purpose") BC548 2N3904 Datasheet

BD Silicon, power transistor BD139 NTE375 Datasheet

BF Silicon, RF (high frequency) BJT or FET BF245 NTE133 Datasheet

BS Silicon, switching transistor (BJT or MOSFET) BS170 2N7000 Datasheet

BL Silicon, high frequency, high power (for transmitters) BLW60 NTE325 Datasheet

BU Silicon, high voltage (for CRT horizontal deflection circuits) BU2520A NTE2354 Datasheet

CF Gallium arsenide small-signal microwave transistor (MESFET) CF739 — Datasheet

CL Gallium arsenide microwave power transistor (FET) CLY10 — Datasheet

Joint Electron Device Engineering Council (JEDEC)[edit]


The JEDEC EIA370 transistor device numbers usually start with "2N", indicating a three-terminal device
(dual-gate field-effect transistors are four-terminal devices, so begin with 3N), then a 2, 3 or 4-digit
sequential number with no significance as to device properties (although early devices with low numbers
tend to be germanium). For example, 2N3055is a silicon n–p–n power transistor, 2N1301 is a p–n–p
germanium switching transistor. A letter suffix (such as "A") is sometimes used to indicate a newer
variant, but rarely gain groupings.
Proprietary[edit]
Manufacturers of devices may have their own proprietary numbering system, for example CK722. Since
devices are second-sourced, a manufacturer's prefix (like "MPF" in MPF102, which originally would
denote a Motorola FET) now is an unreliable indicator of who made the device. Some proprietary naming
schemes adopt parts of other naming schemes, for example a PN2222A is a (possibly Fairchild
Semiconductor) 2N2222A in a plastic case (but a PN108 is a plastic version of a BC108, not a 2N108,
while the PN100 is unrelated to other xx100 devices).
Military part numbers sometimes are assigned their own codes, such as the British Military CV Naming
System.
Manufacturers buying large numbers of similar parts may have them supplied with "house numbers",
identifying a particular purchasing specification and not necessarily a device with a standardized registered
number. For example, an HP part 1854,0053 is a (JEDEC) 2N2218 transistor which is also assigned the
[55][56]

CV number: CV7763 [57]

Naming problems[edit]
With so many independent naming schemes, and the abbreviation of part numbers when printed on the
devices, ambiguity sometimes occurs. For example, two different devices may be marked "J176" (one the
J176 low-power JFET, the other the higher-powered MOSFET 2SJ176).
As older "through-hole" transistors are given surface-mount packaged counterparts, they tend to be
assigned many different part numbers because manufacturers have their own systems to cope with the
variety in pinout arrangements and options for dual or matched n–p–n + p–n–p devices in one pack. So
even when the original device (such as a 2N3904) may have been assigned by a standards authority, and
well known by engineers over the years, the new versions are far from standardized in their naming.

Construction[edit]
Semiconductor material[edit]
Semiconductor material characteristics

Junction forward Max.


Semiconductor Electron mobility Hole mobility
voltage junction temp.
material m2/(V·s) @ 25 °C m2/(V·s) @ 25 °C
V @ 25 °C °C

Ge 0.27 0.39 0.19 70 to 100

Si 0.71 0.14 0.05 150 to 200

GaAs 1.03 0.85 0.05 150 to 200

Al-Si junction 0.3 — — 150 to 200

The first BJTs were made from germanium (Ge). Silicon (Si) types currently predominate but certain
advanced microwave and high-performance versions now employ the compound
semiconductor material gallium arsenide (GaAs) and the semiconductor alloy silicon germanium (SiGe).
Single element semiconductor material (Ge and Si) is described aselemental.
Rough parameters for the most common semiconductor materials used to make transistors are given in the
adjacent table; these parameters will vary with increase in temperature, electric field, impurity level, strain,
and sundry other factors.
The junction forward voltage is the voltage applied to the emitter–base junction of a BJT in order to make
the base conduct a specified current. The current increases exponentially as the junction forward voltage is
increased. The values given in the table are typical for a current of 1 mA (the same values apply to
semiconductor diodes). The lower the junction forward voltage the better, as this means that less power is
required to "drive" the transistor. The junction forward voltage for a given current decreases with increase
in temperature. For a typical silicon junction the change is −2.1 mV/°C. In some circuits special
[58]

compensating elements (sensistors) must be used to compensate for such changes.


The density of mobile carriers in the channel of a MOSFET is a function of the electric field forming the
channel and of various other phenomena such as the impurity level in the channel. Some impurities, called
dopants, are introduced deliberately in making a MOSFET, to control the MOSFET electrical behavior.
The electron mobility and hole mobility columns show the average speed that electrons and holes diffuse
through the semiconductor material with an electric field of 1 volt per meter applied across the material. In
general, the higher the electron mobility the faster the transistor can operate. The table indicates that Ge is
a better material than Si in this respect. However, Ge has four major shortcomings compared to silicon and
gallium arsenide:

 Its maximum temperature is limited;


 it has relatively high leakage current;
 it cannot withstand high voltages;
 it is less suitable for fabricating integrated circuits.
Because the electron mobility is higher than the hole mobility for all semiconductor materials, a given
bipolar n–p–n transistor tends to be swifter than an equivalent p–n–p transistor. GaAs has the highest
electron mobility of the three semiconductors. It is for this reason that GaAs is used in high-frequency
applications. A relatively recent FET development, the high-electron-mobility transistor (HEMT), has
[when?]

a heterostructure (junction between different semiconductor materials) of aluminium gallium arsenide


(AlGaAs)-gallium arsenide (GaAs) which has twice the electron mobility of a GaAs-metal barrier junction.
Because of their high speed and low noise, HEMTs are used in satellite receivers working at frequencies
around 12 GHz. HEMTs based on gallium nitride and aluminium gallium nitride (AlGaN/GaN HEMTs)
provide a still higher electron mobility and are being developed for various applications.
Max. junction temperature values represent a cross section taken from various manufacturers' data
sheets. This temperature should not be exceeded or the transistor may be damaged.
Al–Si junction refers to the high-speed (aluminum–silicon) metal–semiconductor barrier diode, commonly
known as a Schottky diode. This is included in the table because some silicon power IGFETs have
a parasitic reverse Schottky diode formed between the source and drain as part of the fabrication process.
This diode can be a nuisance, but sometimes it is used in the circuit.
Packaging[edit]
See also: Semiconductor package and Chip carrier

Assorted discrete transistors

Soviet KT315b transistors

Discrete transistors can be individually packaged transistors or unpackaged transisor chips (dice).
Transistors come in many different semiconductor packages (see image). The two main categories
are through-hole (or leaded), andsurface-mount, also known as surface-mount device (SMD). The ball grid
array (BGA) is the latest surface-mount package (currently only for large integrated circuits). It has solder
"balls" on the underside in place of leads. Because they are smaller and have shorter interconnections,
SMDs have better high-frequency characteristics but lower power rating.
Transistor packages are made of glass, metal, ceramic, or plastic. The package often dictates the power
rating and frequency characteristics. Power transistors have larger packages that can be clamped to heat
sinks for enhanced cooling. Additionally, most power transistors have the collector or drain physically
connected to the metal enclosure. At the other extreme, some surface-mount microwavetransistors are as
small as grains of sand.
Often a given transistor type is available in several packages. Transistor packages are mainly standardized,
but the assignment of a transistor's functions to the terminals is not: other transistor types can assign other
functions to the package's terminals. Even for the same transistor type the terminal assignment can vary
(normally indicated by a suffix letter to the part number, q.e. BC212L and BC212K).
Nowadays most transistors come in a wide range of SMT packages, in comparison the list of available
through-hole packages is relatively small, here is a short list of the most common through-hole transistors
packages in alphabetical order: ATV, E-line, MRT, HRT, SC-43, SC-72, TO-3, TO-18, TO-39, TO-92,
TO-126, TO220, TO247, TO251, TO262, ZTX851.
Unpackaged transistor chips (die) may be assembled into hybrid devices. The IBM SLT module of the
[59]

1960s is one example of such a hybrid circuit module using glass passivated transistor (and diode) die.
Other packaging techniques for discrete transistors as chips include Direct Chip Attach (DCA) and Chip
On Board (COB). [59]

14.9 JUNCTION TRANSISTOR The credit of inventing the transistor in the year 1947 goes

to J. Bardeen and W.H. Brattain of Bell Telephone Laboratories, U.S.A. That transistor was a

point-contact transistor. The first junction transistor consisting of two back-to-back p-n

junctions was invented by William Schockley in 1951. As long as only the junction transistor

was known, it was known simply as transistor. But over the years new types of transistors

were invented and to differentiate it from the new ones it is now called the Bipolar Junction

Transistor (BJT). Even now, often the word transistor 491 Semiconductor Electronics:

Materials, Devices and Simple Circuits is used to mean BJT when there is no confusion.

Since our study is limited to only BJT, we shall use the word transistor for BJT without any

ambiguity. 14.9.1 Transistor: structure and action A transistor has three doped regions

forming two p-n junctions between them. Obviously, there are two types of transistors, as

shown in Fig. 14.27. (i) n-p-n transistor: Here two segments of n-type semiconductor (emitter

and collector) are separated by a segment of p-type semiconductor (base). (ii) p-n-p

transistor: Here two segments of p-type semiconductor (termed as emitter and collector) are

separated by a segment of n-type semiconductor (termed as base). The schematic

representations of an n-p-n and a p-n-p configuration are shown in Fig. 14.27(a). All the three

segments of a transistor have different thickness and their doping levels are also different. In

the schematic symbols used for representing p-n-p and n-p-n transistors [Fig. 14.27(b)] the

arrowhead shows the direction of conventional current in the transistor. A brief description of

the three segments of a transistor is given below: • Emitter: This is the segment on one side of

the transistor shown in Fig. 14.27(a). It is of moderate size and heavily doped. It supplies a

large number of majority carriers for the current flow through the transistor. • Base: This is

the central segment. It is very thin and lightly doped. • Collector: This segment collects a
major portion of the majority carriers supplied by the emitter. The collector side is

moderately doped and larger in size as compared to the emitter. We have seen earlier in the

case of a p-n junction, that there is a formation of depletion region acorss the junction. In case

of a transistor depletion regions are formed at the emitter base-junction and the basecollector

junction. For understanding the action of a transistor, we have to consider the nature of

depletion regions formed at these junctions. The charge carriers move across different regions

of the transistor when proper voltages are applied across its terminals. The biasing of the

transistor is done differently for different uses. The transistor can be used in two distinct

ways. Basically, it was invented to function as an amplifier, a device which produces a

enlarged copy of a signal. But later its use as a switch acquired equal importance. We shall

study both these functions and the ways the transistor is biased to achieve these mutually

exclusive functions. First we shall see what gives the transistor its amplifying capabilities.

The transistor works as an amplifier, with its emitter-base junction forward biased and the

base-collector junction reverse biased. This situation is shown in Fig. 14.28, where VCC and

VEE are used for creating the respective biasing. When the transistor is biased in this way it

is said to be in active state.We represent the voltage between emitter and base as VEB and

that between the collector and the base as VCB. In FIGURE 14.27 (a) Schematic

representations of a n-p-n transistor and p-n-p transistor, and (b) Symbols for n-p-n and p-n-p

transistors. Physics 492 Fig. 14.28, base is a common terminal for the two power supplies

whose other terminals are connected to emitter and collector, respectively. So the two power

supplies are represented as VEE, and VCC, respectively. In circuits, where emitter is the

common terminal, the power supply between the base and the emitter is represented as VBB

and that between collector and emitter as VCC. Let us see now the paths of current carriers in

the transistor with emitter-base junction forward biased and base-collector junction reverse

biased. The heavily doped emitter has a high concentration of majority carriers, which will be

holes in a p-n-p transistor and electrons in an n-p-n transistor. These majority carriers enter

the base region in large numbers. The base is thin and lightly doped. So the majority carriers
there would be few. In a p-n-p transistor the majority carriers in the base are electrons since

base is of n-type semiconductor. The large number of holes entering the base from the emitter

swamps the small number of electrons there. As the base collector-junction is reversebiased,

these holes, which appear as minority carriers at the junction, can easily cross the junction

and enter the collector. The holes in the base could move either towards the base terminal to

combine with the electrons entering from outside or cross the junction to enter into the

collector and reach the collector terminal. The base is made thin so that most of the holes find

themselves near the reverse-biased base-collector junction and so cross the junction instead of

moving to the base terminal. It is interesting to note that due to forward bias a large current

enters the emitter-base junction, but most of it is diverted to adjacent reverse-biased base-

collector junction and the current coming out of the base becomes a very small fraction of the

current that entered the junction. If we represent the hole current and the electron current

crossing the forward biased junction by Ih and I e respectively then the total current in a

forward biased diode is the sum I h + I e. We see that the emitter current I E = I h + I e but

the base current I B << I h + I e, because a major part of I E goes to collector instead of

coming out of the base terminal. The base current is thus a small fraction of the emitter

current. The current entering into the emitter from outside is equal to the emitter current I E.

Similarly the current emerging from the base terminal is I B and that from collector terminal

is I C. It is obvious from the above description and also from a straight forward application of

Kirchhoff’s law to Fig. 14.28(a) that the emitter current is the sum of collector current and

base current: FIGURE 14.28 Bias Voltage applied on: (a) p-n-p transistor and (b) n-p-n

transistor. 493 Semiconductor Electronics: Materials, Devices and Simple Circuits I E = I C +

I B (14.7) We also see that I C ≈ I E. Our description of the direction of motion of the holes is

identical with the direction of the conventional current. But the direction of motion of

electrons is just opposite to that of the current. Thus in a p-n-p transistor the current enters

from emitter into base whereas in a n-p-n transistor it enters from the base into the emitter.

The arrowhead in the emitter shows the direction of the conventional current. The description
about the paths followed by the majority and minority carriers in a n-p-n is exactly the same

as that for the p-n-p transistor. But the current paths are exactly opposite, as shown in Fig.

14.28. In Fig. 14.28(b) the electrons are the majority carriers supplied by the n-type emitter

region. They cross the thin p-base region and are able to reach the collector to give the

collector current, I C . From the above description we can conclude that in the active state of

the transistor the emitter-base junction acts as a low resistance while the base collector acts as

a high resistance. 14.9.2 Basic transistor circuit configurations and transistor characteristics

In a transistor, only three terminals are available, viz., Emitter (E), Base (B) and Collector

(C). Therefore, in a circuit the input/output connections have to be such that one of these (E,

B or C) is common to both the input and the output. Accordingly, the transistor can be

connected in either of the following three configurations: Common Emitter (CE), Common

Base (CB), Common Collector (CC) The transistor is most widely used in the CE

configuration and we shall restrict our discussion to only this configuration. Since more

commonly used transistors are n-p-n Si transistors, we shall confine our discussion to such

transistors only. With p-n-p transistors the polarities of the external power supplies are to be

inverted. Common emitter transistor characteristics When a transistor is used in CE

configuration, the input is between the base and the emitter and the output is between the

collector and the emitter. The variation of the base current I B with the base-emitter voltage

VBE is called the input characteristic. Similarly, the variation of the collector current IC with

the collector-emitter voltage VCE is called the output characteristic. You will see that the

output characteristics are controlled by the input characteristics. This implies that the

collector current changes with the base current. The input and the output characteristics of an

n-p-n transistors can be studied by using the circuit shown in Fig. 14.29. To study the input

characteristics of the transistor in CE configuration, a curve is plotted between the base

current I B against the base-emitter voltage VBE. The FIGURE 14.29 Circuit arrangement

for studying the input and output characteristics of n-p-n transistor in CE configuration.

Physics 494 collector-emitter voltage VCE is kept fixed while studying the dependence of I B
on VBE. We are interested to obtain the input characteristic when the transistor is in active

state. So the collector-emitter voltage VCE is kept large enough to make the base collector

junction reverse biased. Since VCE = VCB + VBE and for Si transistor VBE is 0.6 to 0.7 V,

VCE must be sufficiently larger than 0.7 V. Since the transistor is operated as an amplifier

over large range of VCE, the reverse bias across the basecollector junction is high most of the

time. Therefore, the input characteristics may be obtained for VCE somewhere in the range of

3 V to 20 V. Since the increase in VCE appears as increase in VCB, its effect on IB is

negligible. As a consequence, input characteristics for various values of VCE will give

almost identical curves. Hence, it is enough to determine only one input characteristics. The

input characteristics of a transistor is as shown in Fig. 14.30(a). The output characteristic is

obtained by observing the variation of I C as VCE is varied keeping I B constant. It is

obvious that if VBE is increased by a small amount, both hole current from the emitter region

and the electron current from the base region will increase. As a consequence both IB and IC

will increase proportionately. This shows that when IB increases I C also increases. The plot

of I C versus VCE for different fixed values of I B gives one output characteristic. So there

will be different output characteristics corresponding to different values of I B as shown in

Fig. 14.30(b). The linear segments of both the input and output characteristics can be used to

calculate some important ac parameters of transistors as shown below. (i) Input resistance (ri

): This is defined as the ratio of change in baseemitter voltage (ΔVBE) to the resulting change

in base current (ΔI B) at constant collector-emitter voltage (VCE). This is dynamic (ac

resistance) and as can be seen from the input characteristic, its value varies with the operating

current in the transistor: CE BE i B V V r I ⎛ ⎞ Δ = ⎜ ⎟ ⎝ ⎠ Δ (14.8) The value of ri can be

anything from a few hundreds to a few thousand ohms. FIGURE 14.30 (a) Typical input

characteristics, and (b) Typical output characteristics. 495 Semiconductor Electronics:

Materials, Devices and Simple Circuits EXAMPLE 14.8 (ii) Output resistance (ro): This is

defined as the ratio of change in collector-emitter voltage (ΔVCE) to the change in collector

current (ΔI C) at a constant base current I B. B CE o C I V r I ⎛ ⎞ Δ = ⎜ ⎟ ⎝ ⎠ Δ (14.9)


The output characteristics show that initially for very small values of VCE, IC increases

almost linearly. This happens because the base-collector junction is not reverse biased and the

transistor is not in active state. In fact, the transistor is in the saturation state and the current is

controlled by the supply voltage VCC (=VCE) in this part of the characteristic. When VCE is

more than that required to reverse bias the base-collector junction, I C increases very little

with VCE. The reciprocal of the slope of the linear part of the output characteristic gives the

values of ro. The output resistance of the transistor is mainly controlled by the bias of the

basecollector junction. The high magnitude of the output resistance (of the order of 100 kΩ)

is due to the reverse-biased state of this diode. This also explains why the resistance at the

initial part of the characteristic, when the transistor is in saturation state, is very low. (iii)

Current amplification factor (β): This is defined as the ratio of the change in collector current

to the change in base current at a constant collector-emitter voltage (VCE) when the transistor

is in active state. CE C ac B V I I β ⎛ ⎞ Δ = ⎜ ⎟ ⎝ ⎠ Δ (14.10) This is also known as small

signal current gain and its value is very large. If we simply find the ratio of IC and I B we get

what is called dc β of the transistor. Hence, C dc B I I β = (14.11) Since I C increases with I

B almost linearly and I C = 0 when I B = 0, the values of both βdc and βac are nearly equal.

So, for most calculations βdc can be used. Both βac and βdc vary with VCE and I B (or I C)

slightly. Example 14.8 From the output characteristics shown in Fig. 14.30(b), calculate the

values of βac and βdc of the transistor when VCE is 10 V and I C = 4.0 mA. Solution CE C

ac B V I I β ⎛ ⎞ Δ = ⎜ ⎟ ⎝ ⎠ Δ , C dc B I I β = For determining βac and βdc at the stated

values of VCE and I C one can proceed as follows. Consider any two characteristics for two

values of IB which lie above and below the given value of IC . Here IC = 4.0 mA. (Choose

characteristics for IB= 30 and 20 μA.) At VCE = 10 V we read the two values of IC from the

graph. Then Physics 496 EXAMPLE 14.8 ΔI B = (30 – 20) μA = 10 μA, ΔI C = (4.5 – 3.0)

mA = 1.5 mA Therefore, βac = 1.5 mA/ 10 μA = 150 For determining βdc, either estimate the

value of I B corresponding to I C = 4.0 mA at VCE = 10 V or calculate the two values of βdc

for the two characteristics chosen and find their mean. Therefore, for IC = 4.5 mA and I B =
30 μA, βdc = 4.5 mA/ 30 μA = 150 and for I C = 3.0 mA and IB = 20 μA βdc =3.0 mA / 20

μA = 150 Hence, βdc =(150 + 150) /2 = 150 14.9.3 Transistor as a device The transistor can

be used as a device application depending on the configuration used (namely CB, CC and

CE), the biasing of the E-B and B-C junction and the operation region namely cutoff, active

region and saturation. As mentioned earlier we have confined only to the CE configuration

and will be concentrating on the biasing and the operation region to understand the working

of a device. When the transistor is used in the cutoff or saturation state it acts as a switch. On

the other hand for using the transistor as an amplifier, it has to operate in the active region. (i)

Transistor as a switch We shall try to understand the operation of the transistor as a switch by

analysing the behaviour of the base-biased transistor in CE configuration as shown in Fig.

14.31(a). Applying Kirchhoff’s voltage rule to the input and output sides of this circuit, we

get VBB = I BRB + VBE (14.12) and VCE = VCC – I CRC. (14.13) We shall treat VBB as

the dc input voltage Vi and VCE as the dc output voltage VO. So, we have Vi = I BRB +

VBE and Vo = VCC – I CRC. Let us see how Vo changes as Vi increases from zero onwards.

In the case of Si transistor, as long as input Vi is less than 0.6 V,the transistor will be in cut

off state and current I C will be zero. Hence Vo = VCC When Vi becomes greater than 0.6 V

the transistor is in active state with some current I C in the output path and the output Vo

decrease as the FIGURE 14.31 (a) Base-biased transistor in CE configuration, (b) Transfer

characteristic. 497 Semiconductor Electronics: Materials, Devices and Simple Circuits term I

CRC increases. With increase of Vi , I C increases almost linearly and so Vo decreases

linearly till its value becomes less than about 1.0 V. Beyond this, the change becomes non

linear and transistor goes into saturation state. With further increase in Vi the output voltage

is found to decrease further towards zero though it may never become zero. If we plot the Vo

vs Vi curve, [also called the transfer characteristics of the base-biased transistor (Fig.

14.31(b)], we see that between cut off state and active state and also between active state and

saturation state there are regions of non-linearity showing that the transition from cutoff state

to active state and from active state to saturation state are not sharply defined. Let us see now
how the transistor is operated as a switch. As long as Vi is low and unable to forward-bias the

transistor, Vo is high (at VCC ). If Vi is high enough to drive the transistor into saturation,

then Vo is low, very near to zero. When the transistor is not conducting it is said to be

switched off and when it is driven into saturation it is said to be switched on. This shows that

if we define low and high states as below and above certain voltage levels corresponding to

cutoff and saturation of the transistor, then we can say that a low input switches the transistor

off and a high input switches it on. Alternatively, we can say that a low input to the transistor

gives a high output and a high input gives a low output. The switching circuits are designed

in such a way that the transistor does not remain in active state. (ii) Transistor as an amplifier

For using the transistor as an amplifier we will use the active region of the Vo versus Vi

curve. The slope of the linear part of the curve represents the rate of change of the output

with the input. It is negative because the output is VCC – ICRC and not I CRC. That is why

as input voltage of the CE amplifier increases its output voltage decreases and the output is

said to be out of phase with the input. If we consider ΔVo and ΔVi as small changes in the

output and input voltages then ΔVo/ΔVi is called the small signal voltage gain AV of the

amplifier. If the VBB voltage has a fixed value corresponding to the mid point of the active

region, the circuit will behave as a CE amplifier with voltage gain ΔVo/ ΔVi . We can

express the voltage gain AV in terms of the resistors in the circuit and the current gain of the

transistor as follows. We have, Vo = VCC – I CRC Therefore, ΔVo = 0 – RC Δ IC Similarly,

from Vi = IBRB + VBE ΔVi = RB ΔI B + ΔVBE But ΔVBE is negligibly small in

comparison to ΔI BRB in this circuit. So, the voltage gain of this CE amplifier (Fig. 14.32) is

given by AV = – RC Δ IC / RB ΔI B = –βac(RC /RB ) (14.14) where βac is equal to Δ

IC/ΔIB from Eq. (14.10). Thus the linear portion of the active region of the transistor can be

exploited for the use in amplifiers. Transistor as an amplifier (CE configuration) is discussed

in detail in the next section. Physics 498 14.9.4 Transistor as an Amplifier (CE-

Configuration) To operate the transistor as an amplifier it is necessary to fix its operating

point somewhere in the middle of its active region. If we fix the value of VBB corresponding
to a point in the middle of the linear part of the transfer curve then the dc base current IB

would be constant and corresponding collector current I C will also be constant. The dc

voltage VCE = VCC - I CRC would also remain constant. The operating values of VCE and I

B determine the operating point, of the amplifier. If a small sinusoidal voltage with amplitude

vs is superposed on the dc base bias by connecting the source of that signal in series with the

VBB supply, then the base current will have sinusoidal variations superimposed on the value

of I B. As a consequence the collector current also will have sinusoidal variations

superimposed on the value of I C, producing in turn corresponding change in the value of

VO. We can measure the ac variations across the input and output terminals by blocking the

dc voltages by large capacitors. In the discription of the amplifier given above we have not

considered any ac signal. In general, amplifiers are used to amplify alternating signals. Now

let us superimpose an ac input signal vi (to be amplified) on the bias VBB (dc) as shown in

Fig. 14.32. The output is taken between the collector and the ground. The working of an

amplifier can be easily understood, if we first assume that vi = 0. Then applying Kirchhoff’s

law to the output loop, we get Vcc = VCE + Ic RL (14.15) Likewise, the input loop gives

VBB = VBE + IB RB (14.16) When vi is not zero, we get VBE + vi = VBE + IB RB + ΔI B

(RB + ri ) The change in VBE can be related to the input resistance ri [see Eq. (14.8)] and the

change in IB. Hence vi = ΔI B (RB + ri ) = r ΔI B The change in I B causes a change in I c .

We define a parameter βac, which is similar to the βdc defined in Eq. (14.11), as c c ac B b I i

I i β Δ = = Δ (14.17) which is also known as the ac current gain Ai . Usually βac is close to

βdc in the linear region of the output characteristics. The change in I c due to a change in I B

causes a change in VCE and the voltage drop across the resistor RL because VCC is fixed.

FIGURE 14.32 A simple circuit of a CE-transistor amplifier. 499 Semiconductor Electronics:

Materials, Devices and Simple Circuits EXAMPLE 14.9 These changes can be given by Eq.

(14.15) as ΔVCC = ΔVCE + RL ΔI C = 0 or ΔVCE = –RL ΔI C The change in VCE is the

output voltage v0. From Eq. (14.10), we get v0 = ΔVCE = –βac RL ΔIB The voltage gain of

the amplifier is 0 CE v i B v V A v rI Δ = = Δ – ac L R r β = (14.18) The negative sign


represents that output voltage is opposite with phase with the input voltage. From the

discussion of the transistor characteristics you have seen that there is a current gain βac in the

CE configuration. Here we have also seen the voltage gain Av. Therefore the power gain A p

can be expressed as the product of the current gain and voltage gain. Mathematically A p =

βac × Av (14.19) Since βac and Av are greater than 1, we get ac power gain. However it

should be realised that transistor is not a power generating device. The energy for the higher

ac power at the output is supplied by the battery. Example 14.9 In Fig. 14.31(a), the VBB

supply can be varied from 0V to 5.0 V. The Si transistor has βdc = 250 and RB = 100 kΩ, RC

= 1 KΩ, VCC = 5.0V. Assume that when the transistor is saturated, VCE = 0V and VBE =

0.8V. Calculate (a) the minimum base current, for which the transistor will reach saturation.

Hence, (b) determine V1 when the transistor is ‘switched on’. (c) find the ranges of V1 for

which the transistor is ‘switched off’ and ‘switched on’. Solution Given at saturation VCE =

0V, VBE = 0.8V VCE = VCC – I CRC IC = VCC/RC = 5.0V/1.0kΩ = 5.0 mA Therefore I B

= I C/β = 5.0 mA/250 = 20μA The input voltage at which the transistor will go into saturation

is given by VIH = VBB = IBRB +VBE = 20μA × 100 kΩ + 0.8V = 2.8V The value of input

voltage below which the transistor remains cutoff is given by VIL = 0.6V, VIH = 2.8V

Between 0.0V and 0.6V, the transistor will be in the ‘switched off’ state. Between 2.8V and

5.0V, it will be in ‘switched on’ state. Note that the transistor is in active state when IB varies

from 0.0mA to 20mA. In this range, IC = βIB is valid. In the saturation range, I C ≤ βI B.

Physics 500 EXAMPLE 14.10 Example 14.10 For a CE transistor amplifier, the audio signal

voltage across the collector resistance of 2.0 kΩ is 2.0 V. Suppose the current amplification

factor of the transistor is 100, What should be the value of RB in series with VBB supply of

2.0 V if the dc base current has to be 10 times the signal current. Also calculate the dc drop

across the collector resistance. (Refer to Fig. 14.33). Solution The output ac voltage is 2.0 V.

So, the ac collector current i C = 2.0/2000 = 1.0 mA. The signal current through the base is,

therefore given by i B = i C /β = 1.0 mA/100 = 0.010 mA. The dc base current has to be 10×

0.010 = 0.10 mA. From Eq.14.16, RB = (VBB - VBE ) /IB. Assuming VBE = 0.6 V, RB =
(2.0 – 0.6 )/0.10 = 14 kΩ. The dc collector current I C = 100×0.10 = 10 mA. 14.9.5 Feedback

amplifier and transistor oscillator In an amplifier, we have seen that a sinusoidal input is

given which appears as an amplified signal in the output. This means that an external input is

necessary to sustain ac signal in the output for an amplifier. In an oscillator, we get ac output

without any external input signal. In other words, the output in an oscillator is self-sustained.

To attain this, an amplifier is taken. A portion of the output power is returned back (feedback)

to the input in phase with the starting power (this process is termed positive feedback) as

shown in Fig. 14.33(a). The feedback can be achieved by inductive coupling (through mutual

inductance) or LC or RC networks. Different types of oscillators essentially use different

methods of coupling the output to the input (feedback network), apart from the resonant

circuit for obtaining oscillation at a particular frequency. For understanding the oscillator

action, we consider the circuit shown in Fig. 14.33(b) in which the feedback is accomplished

by inductive coupling from one coil winding (T1) to another coil winding (T2). Note that the

coils T2 and T1 are wound on the same core and hence are inductively coupled through their

mutual inductance. As in an amplifier, the base-emitter junction is forward biased while the

base-collector junction is reverse biased. Detailed biasing circuits actually used have been

omitted for simplicity. Let us try to understand how oscillations are built. Suppose switch S1

is put on to FIGURE 14.33 (a) Principle of a transistor amplifier with positive feedback

working as an oscillator and (b) Tuned collector oscillator, (c) Rise and fall (or built up) of

current I c and I e due to the inductive coupling. 501 Semiconductor Electronics: Materials,

Devices and Simple Circuits apply proper bias for the first time. Obviously, a surge of

collector current flows in the transistor. This current flows through the coil T2 where

terminals are numbered 3 and 4 [Fig. 14.33(b)]. This current does not reach full amplitude

instantaneously but increases from X to Y, as shown in Fig. [14.33(c)(i)]. The inductive

coupling between coil T2 and coil T1 now causes a current to flow in the emitter circuit (note

that this actually is the ‘feedback’ from input to output). As a result of this positive feedback,

this current (in T1; emitter current) also increases from X´ to Y´ [Fig. In September
1976, in the midst of the Cold War, Victor Ivanovich
Belenko, a disgruntled Soviet pilot, veered off course from a
training flight over Siberia in his MiG-25 Foxbat, flew low
and fast across the Sea of Japan, and landed the plane at a
civilian airport in Hokkaido with just 30 seconds of fuel
remaining. His dramatic defection was a boon for U.S.
military analysts, who for the first time had an opportunity
to examine up close this high-speed Soviet fighter, which
they had thought to be one of the world’s most capable
aircraft. What they discovered astonished them.

For one thing, the airframe was more crudely built than
those of contemporary U.S. fighters, being made mostly of
steel rather than titanium. What’s more, they found the
plane’s avionics bays to be filled with equipment based on
vacuum tubes rather than transistors. The obvious
conclusion, previous fears aside, was that even the Soviet
Union’s most cutting-edge technology lagged laughably
behind the West’s.

After all, in the United States vacuum tubes had given way
to smaller and less power-hungry solid-state devices two
decades earlier, not long after William Shockley, John
Bardeen, and Walter Brattain cobbled together the first
transistor at Bell Laboratories in 1947. By the mid-1970s,
the only vacuum tubes you could find in Western electronics
were hidden away in certain kinds of specialized
equipment—not counting the ubiquitous picture tubes of
television sets. Today even those are gone, and outside of a
few niches, vacuum tubes are an extinct technology. So it
might come as a surprise to learn that some very modest
changes to the fabrication techniques now used to build
integrated circuits could yet breathe vacuum electronics
back to life.

At the NASA Ames Research Center, we’ve been working


for the past few years to develop vacuum-channel
transistors. Our research is still at an early stage, but the
prototypes we’ve constructed show that this novel device
holds extraordinary promise. Vacuum-channel transistors
could work 10 times as fast as ordinary silicon transistors
and may eventually be able to operate at terahertz
frequencies, which have long been beyond the reach of any
solid-state device. And they are considerably more tolerant
of heat and radiation. To understand why, it helps to know a
bit about the construction and functioning of good old-
fashioned vacuum tubes.

Photo: Gregory MaxwellLightbulb Descendant: Vacuum tubes were an outgrowth of


ordinary lightbulbs, a development spurred on by Thomas Edison’s investigations into the
ability of heated filaments to emit electrons. This 1906 example, an early Audion tube, shows
the close resemblance to a lightbulb, although the filament in this particular tube is not
visible, having long ago burned out. That filament once acted as the cathode from which
electrons flew toward the anode or plate, which is located in the center of the glass tube.
Current flow from cathode to anode could be controlled by varying the voltage applied to the
grid, the zigzag wire seen below the plate.

The thumb-size vacuum tubes that amplified signals in


countless radio and television sets during the first half of the
20th century might seem nothing like the metal-oxide
semiconductor field-effect transistors (MOSFETs) that
regularly dazzle us with their capabilities in today’s digital
electronics. But in many ways, they are quite similar. For
one, they both are three-terminal devices. The voltage
applied to one terminal—the grid for a simple triode vacuum
tube and the gate for a MOSFET—controls the amount of
current flowing between the other two: from cathode to
anode in a vacuum tube and from source to drain in a
MOSFET. This ability is what allows each of these devices
to function as an amplifier or, if driven hard enough, as a
switch.

How electric current flows in a vacuum tube is very


different from how it flows in a transistor, though. Vacuum
tubes rely on a process called thermionic emission: Heating
the cathode causes it to shed electrons into the surrounding
vacuum. The current in transistors, on the other hand, comes
from the drift and diffusion of electrons (or of “holes,” spots
where electrons are missing) between the source and the
drain through the solid semiconducting material that
separates them.

Why did vacuum tubes give way to solid-state electronics so


many decades ago? The advantages of semiconductors
include lower costs, much smaller size, superior lifetimes,
efficiency, ruggedness, reliability, and consistency.
Notwithstanding these advantages, when considered purely
as a medium for transporting charge, vacuum wins over
semiconductors. Electrons propagate freely through the
nothingness of a vacuum, whereas they suffer from
collisions with the atoms in a solid (a process called crystal-
lattice scattering). What’s more, a vacuum isn’t prone to the
kind of radiation damage that plagues semiconductors, and
it produces less noise and distortion than solid-state
materials.

The drawbacks of tubes weren’t so vexing when you just


needed a handful of them to run your radio or television set.
But they proved really troublesome with more complicated
circuits. For example, the 1946 ENIAC computer, which
used 17,468 vacuum tubes, consumed 150 kilowatts of
power, weighed more than 27 metric tons, and took up
almost 200 square meters of floor space. And it kept
breaking down all the time, with a tube failing every day or
two.

Illustration: James ProvostChip in a Bottle: The simplest vacuum tube capable of


amplification is the triode, so named because it contains three electrodes: a cathode, an
anode, and a grid. Typically, the structure is cylindrically symmetrical, with the cathode
surrounded by the grid and the grid surrounded by the anode. Operation is similar to that of a
field-effect transistor, here with the voltage applied to the grid controlling the current flow
between the other two electrodes. (Triode tubes often have five pins to accommodate two
additional electrical connections for the heated filament.)

The transistor revolution put an end to such frustrations. But


the ensuing sea change in electronics came about not so
much because of the intrinsic advantages of semiconductors
but because engineers gained the ability to mass-produce
and combine transistors in integrated circuits by chemically
engraving, or etching, a silicon wafer with the appropriate
pattern. As the technology of integrated-circuit fabrication
progressed, more and more transistors could be squeezed
onto microchips, allowing the circuitry to become more
elaborate from one generation to the next. The electronics
also became faster without costing any more.

That speed benefit stemmed from the fact that as the


transistors became smaller, electrons moving through them
had to travel increasingly shorter distances between the
source and the drain, allowing each transistor to be turned
on and off more quickly. Vacuum tubes, on the other hand,
were big and bulky and had to be fabricated individually by
mechanical machining. While they were improved over the
years, tubes never benefited from anything remotely
resembling Moore’s Law.

But after four decades of shrinking transistor dimensions,


the oxide layer that insulates the gate electrode of a typical
MOSFET is now only a few nanometers thick, and just a
few tens of nanometers separate its source and drain.
Conventional transistors really can’t get much smaller. Still,
the quest for faster and more energy-efficient chips
continues. What will the next transistor technology
be? Nanowires, carbon nanotubes, and graphene are all
being developed intensively. Perhaps one of these
approaches will revamp the electronics industry. Or maybe
they’ll all fizzle.
We’ve been working to develop yet another candidate to
replace the MOSFET, one that researchers have been
dabbling with off and on for many years: the vacuum-
channel transistor. It’s the result of a marriage between
traditional vacuum-tube technology and modern
semiconductor-fabrication techniques. This curious hybrid
combines the best aspects of vacuum tubes and transistors
and can be made as small and as cheap as any solid-state
device. Indeed, making them small is what eliminates the
well-known drawbacks of vacuum tubes.

Illustration: James Provost; Inset Image: NASA Ames Research CenterTransistorizing


the Vacuum Tube: A vacuum-channel transistor closely resembles an ordinary metal-oxide
semiconductor field-effect transistor or MOSFET [left]. In a MOSFET, voltage applied to the
gate sets up an electric field in the semiconductor material below. This field in turn draws
charge carriers into the channel between the source and drain regions, allowing current to
flow. No current flows into the gate, which is insulated from the substrate below it by a thin
oxide layer. The vacuum-channel transistor the authors developed [right] similarly uses a thin
layer of oxide to insulate the gate from the cathode and anode, which are sharply pointed to
intensify the electric field at the tips.

In a vacuum tube, an electric filament, similar to the


filament in an incandescent lightbulb, is used to heat the
cathode sufficiently for it to emit electrons. This is why
vacuum tubes need time to warm up and why they consume
so much power. It’s also why they frequently burn out (often
as a result of a minuscule leak in the tube’s glass envelope).
But vacuum-channel transistors don’t need a filament or hot
cathode. If the device is made small enough, the electric
field across it is sufficient to draw electrons from the source
by a process known as field emission. Eliminating the
power-sapping heating element reduces the area each device
takes up on a chip and makes this new kind of transistor
energy efficient.

Another weak point of tubes is that they must maintain a


high vacuum, typically a thousandth or so of atmospheric
pressure, to avoid collisions between electrons and gas
molecules. Under such low pressure, the electric field causes
positive ions generated from the residual gas in a tube to
accelerate and bombard the cathode, creating sharp,
nanometer-scale protrusions, which degrade and, ultimately,
destroy it.

These long-standing problems of vacuum electronics aren’t


insurmountable. What if the distance between cathode and
anode were less than the average distance an electron travels
before hitting a gas molecule, a distance known as the mean
free path? Then you wouldn’t have to worry about collisions
between electrons and gas molecules. For example, the
mean free path of electrons in air under normal atmospheric
pressure is about 200 nanometers, which on the scale of
today’s transistors is pretty large. Use helium instead of air
and the mean free path goes up to about 1 micrometer. That
means an electron traveling across, say, a 100-nm gap
bathed in helium would have only about a 10 percent
probability of colliding with the gas. Make the gap smaller
still and the chance of collision diminishes further.

But even with a low probability of hitting, many electrons


are still going to collide with gas molecules. If the impact
knocks a bound electron from the gas molecule, it will
become a positively charged ion, which means that the
electric field will send it flying toward the cathode. Under
the bombardment of all those positive ions, cathodes
degrade. So you really want to avoid this as much as
possible.

Fortunately, if you keep the voltage low, the electrons will


never acquire enough energy to ionize helium. So if the
dimensions of the vacuum transistor are substantially
smaller than the mean free path of electrons (which is not
hard to arrange), and the working voltage is low enough (not
difficult either), the device can operate just fine at
atmospheric pressure. That is, you don’t, in fact, need to
maintain any sort of vacuum at all for what is nominally a
miniaturized piece of “vacuum” electronics!

But how do you turn this new kind of transistor on and off?
With a triode vacuum tube, you control the current flowing
through it by varying the voltage applied to the grid—
a meshlike electrode situated between the cathode and the
anode. Positioning the grid close to the cathode enhances
the grid’s electrostatic control, although that close
positioning tends to increase the amount of current flowing
into the grid. Ideally, no current would ever flow into the
grid, because it wastes energy and can even cause the tube to
malfunction. But in practice there’s always a little grid
current.
To avoid such problems, we control current flow in our
vacuum-channel transistor just as it’s done in ordinary
MOSFETs, using a gate electrode that has an insulating
dielectric material (silicon dioxide) separating it from the
current channel. The dielectric insulator transfers the electric
field where it’s needed while preventing the flow of current
into the gate.

So you see, the vacuum-channel transistor isn’t at all


complicated. Indeed, it operates much more simply than any
of the transistor varieties that came before it.

Although we are still at an early stage with our research,


we believe the recent improvements we’ve made to the
vacuum-channel transistor could one day have a huge
influence on the electronics industry, particularly for
applications where speed is paramount. Our very first effort
to fashion a prototype produced a device that could operate
at 460 gigahertz—roughly 10 times as fast as the best silicon
transistor can manage. This makes the vacuum-channel
transistor very promising for operating in what is sometimes
known as the terahertz gap, the portion of the
electromagnetic spectrum above microwaves and below
infrared.

Filling the Gap: Vacuum-channel transistors hold the promise of being able to operate
at frequencies above microwaves and below infrared—a region of the spectrum sometimes
known as the terahertz gap because of the difficulty that most semiconductor devices have
operating at those frequencies. Promising applications for terahertz equipment include
directional high-speed communications and hazardous-materials sensing.
Such frequencies, which run from about 0.1 to 10 terahertz,
are useful for sensing hazardous materials and for secure
high-speed telecommunications, to give just a couple of
possible applications. But terahertz waves are difficult to
take advantage of because conventional semiconductors
aren’t capable of generating or detecting this radiation.
Vacuum transistors could—pardon the expression—fill that
void. These transistors might also find their way into future
microprocessors, their method of manufacture being
completely compatible with conventional CMOS
fabrication. But several problems will need to be solved
before that can happen.

Our prototype vacuum transistor operates at 10 volts, an


order of magnitude higher than modern CMOS chips use.
But researchers at the University of Pittsburgh have been
able to build vacuum transistors that operate at just 1 or 2 V,
albeit with significant compromises in design flexibility.
We’re confident we can reduce the voltage requirements of
our device to similar levels by shrinking the distance
between its anode and cathode. Also, the sharpness of these
electrodes determines how much they concentrate the
electric field, and the makeup of the cathode material
governs how large a field is needed to extract electrons from
it. So we might also be able to reduce the voltage needed by
designing electrodes with sharper points or a more
advantageous chemical composition that lowers the barrier
for the electron escaping from the cathode. This will no
doubt be something of a balancing act, because changes
made to reduce operating voltage could compromise the
long-term stability of the electrodes and the resultant
lifetime of the transistor.

The next big step for us is to build a large number of


vacuum-channel transistors into an integrated circuit. For
that, we should be able to use many of the existing
computer-aided design tools and simulation software
developed for constructing CMOS ICs. Before we attempt
this, however, we’ll need to refine our computer models for
this new transistor and to work out suitable design rules for
wiring lots of them together. And we’ll have to devise
proper packaging methods for these 1-atmosphere, helium-
filled devices. Most likely, the techniques currently used to
package various microelectromechanical sensors, such as
accelerometers and gyroscopes, can be applied to vacuum-
channel transistors without too much fuss.

Admittedly, a great deal of work remains to be done before


we can begin to envision commercial products emerging.
But when they eventually do, this new generation of vacuum
electronics will surely boast some surprising capabilities.
Expect that. Otherwise you might end up feeling a bit like
those military analysts who examined that Soviet MiG-25 in
Japan back in 1976: Later they realized that its vacuum-
based avionics could withstand the electromagnetic pulse
from a nuclear blast better than anything the West had in its
planes. Only then did they begin to appreciate the value of a
little nothingness.
14.33(c)(ii)]. The current in T2 (collector current) connected in the collector circuit acquires the value
Y when the transistor becomes saturated. This means that maximum collector current is flowing and
can increase no further. Since there is no further change in collector current, the magnetic field around
T2 ceases to grow. As soon as the field becomes static, there will be no further feedback from T2 to
T1. Without continued feedback, the emitter current begins to fall. Consequently, collector current
decreases from Y towards Z [Fig. 14.33(c)(i)]. However, a decrease of collector current causes the
magnetic field to decay around the coil T2. Thus, T1 is now seeing a decaying field in T2 (opposite
from what it saw when the field was growing at the initial start operation). This causes a further
decrease in the emitter current till it reaches Z′when the transistor is cut-off. This means that both I E
and I C cease to flow. Therefore, the transistor has reverted back to its original state (when the power
was first switched on). The whole process now repeats itself. That is, the transistor is driven to
saturation, then to cut-off, and then back to saturation. The time for change from saturation to cut-off
and back is determined by the constants of the tank circuit or tuned circuit (inductance L of coil T2
and C connected in parallel to it). The resonance frequency (ν) of this tuned circuit determines the
frequency at which the oscillator will oscillate. π 1 2 LC ν ⎛ ⎞ = ⎜ ⎟ ⎝ ⎠ (14.20) In the circuit of
Fig. 14.33(b), the tank or tuned circuit is connected in the collector side. Hence, it is known as tuned
collector oscillator. If the tuned circuit is on the base side, it will be known as tuned base oscillator.
There are many other types of tank circuits (say RC) or feedback circuits giving different types of
oscillators like Colpitt’s oscillator, Hartley oscillator, RC-oscillator.

Potrebbero piacerti anche