Field-effect transistor

Last updated

Cross-sectional view of a field-effect transistor, showing source, gate and drain terminals FET cross section.svg
Cross-sectional view of a field-effect transistor, showing source, gate and drain terminals

The field-effect transistor (FET) is a type of transistor that uses an electric field to control the flow of current in a semiconductor. It comes in two types: junction FET (JFET) and metal-oxide-semiconductor FET (MOSFET). FETs have three terminals: source, gate, and drain. FETs control the flow of current by the application of a voltage to the gate, which in turn alters the conductivity between the drain and source.

Contents

FETs are also known as unipolar transistors since they involve single-carrier-type operation. That is, FETs use either electrons (n-channel) or holes (p-channel) as charge carriers in their operation, but not both. Many different types of field effect transistors exist. Field effect transistors generally display very high input impedance at low frequencies. The most widely used field-effect transistor is the MOSFET (metal–oxide–semiconductor field-effect transistor).

History

Julius Edgar Lilienfeld proposed the concept of a field-effect transistor in 1925. Julius Edgar Lilienfeld (1881-1963).jpg
Julius Edgar Lilienfeld proposed the concept of a field-effect transistor in 1925.

The concept of a field-effect transistor (FET) was first patented by the Austro-Hungarian born physicist Julius Edgar Lilienfeld in 1925 [1] and by Oskar Heil in 1934, but they were unable to build a working practical semiconducting device based on the concept. The transistor effect was later observed and explained by John Bardeen and Walter Houser Brattain while working under William Shockley at Bell Labs in 1947, shortly after the 17-year patent expired. Shockley initially attempted to build a working FET by trying to modulate the conductivity of a semiconductor, but was unsuccessful, mainly due to problems with the surface states, the dangling bond, and the germanium and copper compound materials. In the course of trying to understand the mysterious reasons behind their failure to build a working FET, it led to Bardeen and Brattain instead inventing the point-contact transistor in 1947, which was followed by Shockley's bipolar junction transistor in 1948. [2] [3]

The first FET device to be successfully built was the junction field-effect transistor (JFET). [2] A JFET was first patented by Heinrich Welker in 1945. [4] The static induction transistor (SIT), a type of JFET with a short channel, was invented by Japanese engineers Jun-ichi Nishizawa and Y. Watanabe in 1950. Following Shockley's theoretical treatment on the JFET in 1952, a working practical JFET was built by George C. Dacey and Ian M. Ross in 1953. [5] However, the JFET still had issues affecting junction transistors in general. [6] Junction transistors were relatively bulky devices that were difficult to manufacture on a mass-production basis, which limited them to a number of specialised applications. The insulated-gate field-effect transistor (IGFET) was theorized as a potential alternative to junction transistors, but researchers were unable to build working IGFETs, largely due to the troublesome surface state barrier that prevented the external electric field from penetrating into the material. [6] By the mid-1950s, researchers had largely given up on the FET concept, and instead focused on bipolar junction transistor (BJT) technology. [7]

The foundations of MOSFET technology were laid down by the work of William Shockley, John Bardeen and Walter Brattain. Shockley independently envisioned the FET concept in 1945, but he was unable to build a working device. The next year Bardeen explained his failure in terms of surface states. Bardeen applied the theory of surface states on semiconductors (previous work on surface states was done by Shockley in 1939 and Igor Tamm in 1932) and realized that the external field was blocked at the surface because of extra electrons which are drawn to the semiconductor surface. Electrons become trapped in those localized states forming an inversion layer. Bardeen's hypothesis marked the birth of surface physics. Bardeen then decided to make use of an inversion layer instead of the very thin layer of semiconductor which Shockley had envisioned in his FET designs. Based on his theory, in 1948 Bardeen patented the progenitor of MOSFET, an insulated-gate FET (IGFET) with an inversion layer. The inversion layer confines the flow of minority carriers, increasing modulation and conductivity, although its electron transport depends on the gate's insulator or quality of oxide if used as an insulator, deposited above the inversion layer. Bardeen's patent as well as the concept of an inversion layer forms the basis of CMOS technology today. In 1976 Shockley described Bardeen's surface state hypothesis "as one of the most significant research ideas in the semiconductor program". [8]

After Bardeen's surface state theory the trio tried to overcome the effect of surface states. In late 1947, Robert Gibney and Brattain suggested the use of electrolyte placed between metal and semiconductor to overcome the effects of surface states. Their FET device worked, but amplification was poor. Bardeen went further and suggested to rather focus on the conductivity of the inversion layer. Further experiments led them to replace electrolyte with a solid oxide layer in the hope of getting better results. Their goal was to penetrate the oxide layer and get to the inversion layer. However, Bardeen suggested they switch from silicon to germanium and in the process their oxide got inadvertently washed off. They stumbled upon a completely different transistor, the point-contact transistor. Lillian Hoddeson argues that "had Brattain and Bardeen been working with silicon instead of germanium they would have stumbled across a successful field effect transistor". [8] [9] [10] [11] [12]

By the end of the first half of the 1950s, following theoretical and experimental work of Bardeen, Brattain, Kingston, Morrison and others, it became more clear that there were two types of surface states. Fast surface states were found to be associated with the bulk and a semiconductor/oxide interface. Slow surface states were found to be associated with the oxide layer because of adsorption of atoms, molecules and ions by the oxide from the ambient. The latter were found to be much more numerous and to have much longer relaxation times. At the time Philo Farnsworth and others came up with various methods of producing atomically clean semiconductor surfaces.

In 1955, Carl Frosch and Lincoln Derrick accidentally covered the surface of silicon wafer with a layer of silicon dioxide. They showed that oxide layer prevented certain dopants into the silicon wafer, while allowing for others, thus discovering the passivating effect of oxidation on the semiconductor surface. Their further work demonstrated how to etch small openings in the oxide layer to diffuse dopants into selected areas of the silicon wafer. In 1957, they published a research paper and patented their technique summarizing their work. The technique they developed is known as oxide diffusion masking, which would later be used in the fabrication of MOSFET devices. At Bell Labs, the importance of Frosch's technique was immediately realized. Results of their work circulated around Bell Labs in the form of BTL memos before being published in 1957. At Shockley Semiconductor, Shockley had circulated the preprint of their article in December 1956 to all his senior staff, including Jean Hoerni. [6] [13] [14]

In 1955, Ian Munro Ross filed a patent for a FeFET or MFSFET. Its structure was like that of a modern inversion channel MOSFET, but ferroelectric material was used as a dielectric/insulator instead of oxide. He envisioned it as a form of memory, years before the floating gate MOSFET. In February 1957, John Wallmark filed a patent for FET in which germanium monoxide was used as a gate dielectric, but he didn't pursue the idea. In his other patent filed the same year he described a double gate FET. In March 1957, in his laboratory notebook, Ernesto Labate, a research scientist at Bell Labs, conceived of a device similar to the later proposed MOSFET, although Labate's device didn't explicitly use silicon dioxide as an insulator. [15] [16] [17] [18]

Metal-oxide-semiconductor FET (MOSFET)

Atalla1963.png
Mohamed Atalla (left) and Dawon Kahng (right) invented the MOSFET (MOS field-effect transistor) in 1959.

A breakthrough in FET research came with the work of Egyptian engineer Mohamed Atalla in the late 1950s. [3] In 1958 he presented experimental work which showed that growing thin silicon oxide on clean silicon surface leads to neutralization of surface states. This is known as surface passivation, a method that became critical to the semiconductor industry as it made mass-production of silicon integrated circuits possible. [19] [20]

The metal–oxide–semiconductor field-effect transistor (MOSFET) was then invented by Mohamed Atalla and Dawon Kahng in 1959. [21] [22] The MOSFET largely superseded both the bipolar transistor and the JFET, [2] and had a profound effect on digital electronic development. [23] [22] With its high scalability, [24] and much lower power consumption and higher density than bipolar junction transistors, [25] the MOSFET made it possible to build high-density integrated circuits. [26] The MOSFET is also capable of handling higher power than the JFET. [27] The MOSFET was the first truly compact transistor that could be miniaturised and mass-produced for a wide range of uses. [6] The MOSFET thus became the most common type of transistor in computers, electronics, [20] and communications technology (such as smartphones). [28] The US Patent and Trademark Office calls it a "groundbreaking invention that transformed life and culture around the world". [28]

CMOS (complementary MOS), a semiconductor device fabrication process for MOSFETs, was developed by Chih-Tang Sah and Frank Wanlass at Fairchild Semiconductor in 1963. [29] [30] The first report of a floating-gate MOSFET was made by Dawon Kahng and Simon Sze in 1967. [31] A double-gate MOSFET was first demonstrated in 1984 by Electrotechnical Laboratory researchers Toshihiro Sekigawa and Yutaka Hayashi. [32] [33] FinFET (fin field-effect transistor), a type of 3D non-planar multi-gate MOSFET, originated from the research of Digh Hisamoto and his team at Hitachi Central Research Laboratory in 1989. [34] [35]

Basic information

FETs can be majority-charge-carrier devices, in which the current is carried predominantly by majority carriers, or minority-charge-carrier devices, in which the current is mainly due to a flow of minority carriers. [36] The device consists of an active channel through which charge carriers, electrons or holes, flow from the source to the drain. Source and drain terminal conductors are connected to the semiconductor through ohmic contacts. The conductivity of the channel is a function of the potential applied across the gate and source terminals.

The FET's three terminals are: [37]

  1. source (S), through which the carriers enter the channel. Conventionally, current entering the channel at S is designated by IS.
  2. drain (D), through which the carriers leave the channel. Conventionally, current leaving the channel at D is designated by ID. Drain-to-source voltage is VDS.
  3. gate (G), the terminal that modulates the channel conductivity. By applying voltage to G, one can control ID.

More about terminals

Cross section of an n-type MOSFET Lateral mosfet.svg
Cross section of an n-type MOSFET

All FETs have source, drain, and gate terminals that correspond roughly to the emitter, collector, and base of BJTs. Most FETs have a fourth terminal called the body, base, bulk, or substrate. This fourth terminal serves to bias the transistor into operation; it is rare to make non-trivial use of the body terminal in circuit designs, but its presence is important when setting up the physical layout of an integrated circuit. The size of the gate, length L in the diagram, is the distance between source and drain. The width is the extension of the transistor, in the direction perpendicular to the cross section in the diagram (i.e., into/out of the screen). Typically the width is much larger than the length of the gate. A gate length of 1 µm limits the upper frequency to about 5 GHz, 0.2 µm to about 30 GHz.

The names of the terminals refer to their functions. The gate terminal may be thought of as controlling the opening and closing of a physical gate. This gate permits electrons to flow through or blocks their passage by creating or eliminating a channel between the source and drain. Electron-flow from the source terminal towards the drain terminal is influenced by an applied voltage. The body simply refers to the bulk of the semiconductor in which the gate, source and drain lie. Usually the body terminal is connected to the highest or lowest voltage within the circuit, depending on the type of the FET. The body terminal and the source terminal are sometimes connected together since the source is often connected to the highest or lowest voltage within the circuit, although there are several uses of FETs which do not have such a configuration, such as transmission gates and cascode circuits.

Unlike BJTs, the vast majority of FETs are electrically symmetrical. The source and drain terminals can thus be interchanged in practical circuits with no change in operating characteristics or function. This can be confusing when FET's appear to be connected "backwards" in schematic diagrams and circuits because the physical orientation of the FET was decided for other reasons, such as printed circuit layout considerations.

Effect of gate voltage on current

I-V characteristics and output plot of a JFET n-channel transistor JFET n-channel en.svg
I–V characteristics and output plot of a JFET n-channel transistor
Simulation result for right side: formation of inversion channel (electron density) and left side: current-gate voltage curve (transfer characteristics) in an n-channel nanowire MOSFET. Note that the threshold voltage for this device lies around 0.45 V. Threshold formation nowatermark.gif
Simulation result for right side: formation of inversion channel (electron density) and left side: current-gate voltage curve (transfer characteristics) in an n-channel nanowire MOSFET. Note that the threshold voltage for this device lies around 0.45 V.
FET conventional symbol types FET Symbols.svg
FET conventional symbol types

The FET controls the flow of electrons (or electron holes) from the source to drain by affecting the size and shape of a "conductive channel" created and influenced by voltage (or lack of voltage) applied across the gate and source terminals. (For simplicity, this discussion assumes that the body and source are connected.) This conductive channel is the "stream" through which electrons flow from source to drain.

n-channel FET

In an n-channel "depletion-mode" device, a negative gate-to-source voltage causes a depletion region to expand in width and encroach on the channel from the sides, narrowing the channel. If the active region expands to completely close the channel, the resistance of the channel from source to drain becomes large, and the FET is effectively turned off like a switch (see right figure, when there is very small current). This is called "pinch-off", and the voltage at which it occurs is called the "pinch-off voltage". Conversely, a positive gate-to-source voltage increases the channel size and allows electrons to flow easily (see right figure, when there is a conduction channel and current is large).

In an n-channel "enhancement-mode" device, a conductive channel does not exist naturally within the transistor, and a positive gate-to-source voltage is necessary to create one. The positive voltage attracts free-floating electrons within the body towards the gate, forming a conductive channel. But first, enough electrons must be attracted near the gate to counter the dopant ions added to the body of the FET; this forms a region with no mobile carriers called a depletion region, and the voltage at which this occurs is referred to as the threshold voltage of the FET. Further gate-to-source voltage increase will attract even more electrons towards the gate which are able to create an active channel from source to drain; this process is called inversion.

p-channel FET

In a p-channel "depletion-mode" device, a positive voltage from gate to body widens the depletion layer by forcing electrons to the gate-insulator/semiconductor interface, leaving exposed a carrier-free region of immobile, positively charged acceptor ions.

Conversely, in a p-channel "enhancement-mode" device, a conductive region does not exist and negative voltage must be used to generate a conduction channel.

Effect of drain-to-source voltage on channel

For either enhancement- or depletion-mode devices, at drain-to-source voltages much less than gate-to-source voltages, changing the gate voltage will alter the channel resistance, and drain current will be proportional to drain voltage (referenced to source voltage). In this mode the FET operates like a variable resistor and the FET is said to be operating in a linear mode or ohmic mode. [38] [39]

If drain-to-source voltage is increased, this creates a significant asymmetrical change in the shape of the channel due to a gradient of voltage potential from source to drain. The shape of the inversion region becomes "pinched-off" near the drain end of the channel. If drain-to-source voltage is increased further, the pinch-off point of the channel begins to move away from the drain towards the source. The FET is said to be in saturation mode; [40] although some authors refer to it as active mode, for a better analogy with bipolar transistor operating regions. [41] [42] The saturation mode, or the region between ohmic and saturation, is used when amplification is needed. The in-between region is sometimes considered to be part of the ohmic or linear region, even where drain current is not approximately linear with drain voltage.

Even though the conductive channel formed by gate-to-source voltage no longer connects source to drain during saturation mode, carriers are not blocked from flowing. Considering again an n-channel enhancement-mode device, a depletion region exists in the p-type body, surrounding the conductive channel and drain and source regions. The electrons which comprise the channel are free to move out of the channel through the depletion region if attracted to the drain by drain-to-source voltage. The depletion region is free of carriers and has a resistance similar to silicon. Any increase of the drain-to-source voltage will increase the distance from drain to the pinch-off point, increasing the resistance of the depletion region in proportion to the drain-to-source voltage applied. This proportional change causes the drain-to-source current to remain relatively fixed, independent of changes to the drain-to-source voltage, quite unlike its ohmic behavior in the linear mode of operation. Thus, in saturation mode, the FET behaves as a constant-current source rather than as a resistor, and can effectively be used as a voltage amplifier. In this case, the gate-to-source voltage determines the level of constant current through the channel.

Composition

FETs can be constructed from various semiconductors, out of which silicon is by far the most common. Most FETs are made by using conventional bulk semiconductor processing techniques, using a single crystal semiconductor wafer as the active region, or channel.

Among the more unusual body materials are amorphous silicon, polycrystalline silicon or other amorphous semiconductors in thin-film transistors or organic field-effect transistors (OFETs) that are based on organic semiconductors; often, OFET gate insulators and electrodes are made of organic materials, as well. Such FETs are manufactured using a variety of materials such as silicon carbide (SiC), gallium arsenide (GaAs), gallium nitride (GaN), and indium gallium arsenide (InGaAs).

In June 2011, IBM announced that it had successfully used graphene-based FETs in an integrated circuit. [43] [44] These transistors are capable of about 2.23 GHz cutoff frequency, much higher than standard silicon FETs. [45]

Types

Depletion-type FETs under typical voltages: JFET, poly-silicon MOSFET, double-gate MOSFET, metal-gate MOSFET, MESFET.
.mw-parser-output .legend{page-break-inside:avoid;break-inside:avoid-column}.mw-parser-output .legend-color{display:inline-block;min-width:1.25em;height:1.25em;line-height:1.25;margin:1px 0;text-align:center;border:1px solid black;background-color:transparent;color:black}.mw-parser-output .legend-text{}
Depletion
Electrons
Holes
Metal
Insulator
Top: source, bottom: drain, left: gate, right: bulk. Voltages that lead to channel formation are not shown. FET comparison.png
Depletion-type FETs under typical voltages: JFET, poly-silicon MOSFET, double-gate MOSFET, metal-gate MOSFET, MESFET.
  Depletion
  Electrons
  Holes
  Metal
  Insulator
Top: source, bottom: drain, left: gate, right: bulk. Voltages that lead to channel formation are not shown.

The channel of a FET is doped to produce either an n-type semiconductor or a p-type semiconductor. The drain and source may be doped of opposite type to the channel, in the case of enhancement mode FETs, or doped of similar type to the channel as in depletion mode FETs. Field-effect transistors are also distinguished by the method of insulation between channel and gate. Types of FETs include:

Advantages

Field-effect transistors have high gate-to-drain current resistance, of the order of 100 MΩ or more, providing a high degree of isolation between control and flow. Because base current noise will increase with shaping time[ clarification needed ], [57] a FET typically produces less noise than a bipolar junction transistor (BJT), and is found in noise-sensitive electronics such as tuners and low-noise amplifiers for VHF and satellite receivers. It exhibits no offset voltage at zero drain current and makes an excellent signal chopper. It typically has better thermal stability than a BJT. [37]

Because the FETs are controlled by gate charge, once the gate is closed or open, there is no additional power draw, as there would be with a bipolar junction transistor or with non-latching relays in some states. This allows extremely low-power switching, which in turn allows greater miniaturization of circuits because heat dissipation needs are reduced compared to other types of switches.

Disadvantages

A field-effect transistor has a relatively low gain–bandwidth product compared to a bipolar junction transistor. MOSFETs are very susceptible to overload voltages, thus requiring special handling during installation. [58] The fragile insulating layer of the MOSFET between the gate and the channel makes it vulnerable to electrostatic discharge or changes to threshold voltage during handling. This is not usually a problem after the device has been installed in a properly designed circuit.

FETs often have a very low "on" resistance and have a high "off" resistance. However, the intermediate resistances are significant, and so FETs can dissipate large amounts of power while switching. Thus, efficiency can put a premium on switching quickly, but this can cause transients that can excite stray inductances and generate significant voltages that can couple to the gate and cause unintentional switching. FET circuits can therefore require very careful layout and can involve trades between switching speed and power dissipation. There is also a trade-off between voltage rating and "on" resistance, so high-voltage FETs have a relatively high "on" resistance and hence conduction losses. [59]

Failure modes

Field-effect transistors are relatively robust, especially when operated within the temperature and electrical limitations defined by the manufacturer (proper derating). However, modern FET devices can often incorporate a body diode. If the characteristics of the body diode are not taken into consideration, the FET can experience slow body diode behavior, where a parasitic transistor will turn on and allow high current to be drawn from drain to source when the FET is off. [60]

Uses

The most commonly used FET is the MOSFET. The CMOS (complementary metal oxide semiconductor) process technology is the basis for modern digital integrated circuits. This process technology uses an arrangement where the (usually "enhancement-mode") p-channel MOSFET and n-channel MOSFET are connected in series such that when one is on, the other is off.

In FETs, electrons can flow in either direction through the channel when operated in the linear mode. The naming convention of drain terminal and source terminal is somewhat arbitrary, as the devices are typically (but not always) built symmetrical from source to drain. This makes FETs suitable for switching analog signals between paths (multiplexing). With this concept, one can construct a solid-state mixing board, for example. FET is commonly used as an amplifier. For example, due to its large input resistance and low output resistance, it is effective as a buffer in common-drain (source follower) configuration.

IGBTs are used in switching internal combustion engine ignition coils, where fast switching and voltage blocking capabilities are important.

Source-gated transistor

Source-gated transistors are more robust to manufacturing and environmental issues in large-area electronics such as display screens, but are slower in operation than FETs. [61]

See also

Related Research Articles

<span class="mw-page-title-main">Transistor</span> Solid-state electrically operated switch also used as an amplifier

A transistor is a semiconductor device used to amplify or switch electrical signals and power. It is one of the basic building blocks of modern electronics. It is composed of semiconductor material, usually with at least three terminals for connection to an electronic 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. Some transistors are packaged individually, but many more in miniature form are found embedded in integrated circuits. Because transistors are the key active components in practically all modern electronics, many people consider them one of the 20th century's greatest inventions.

<span class="mw-page-title-main">Semiconductor device</span> Electronic component that exploits the electronic properties of semiconductor materials

A semiconductor device is an electronic component that relies on the electronic properties of a semiconductor material for its function. Its conductivity lies between conductors and insulators. Semiconductor devices have replaced vacuum tubes in most applications. They conduct electric current in the solid state, rather than as free electrons across a vacuum or as free electrons and ions through an ionized gas.

<span class="mw-page-title-main">MOSFET</span> Type of field-effect transistor

The metal–oxide–semiconductor field-effect transistor is a type of field-effect transistor (FET), most commonly fabricated by the controlled oxidation of silicon. It has an insulated gate, the voltage of which determines the conductivity of the device. This ability to change conductivity with the amount of applied voltage can be used for amplifying or switching electronic signals. The term metal–insulator–semiconductor field-effect transistor (MISFET) is almost synonymous with MOSFET. Another near-synonym is insulated-gate field-effect transistor (IGFET).

<span class="mw-page-title-main">JFET</span> Type of field-effect transistor

The junction field-effect transistor (JFET) is one of the simplest types of field-effect transistor. JFETs are three-terminal semiconductor devices that can be used as electronically controlled switches or resistors, or to build amplifiers.

NMOS or nMOS logic uses n-type (-) MOSFETs to implement logic gates and other digital circuits.

A power semiconductor device is a semiconductor device used as a switch or rectifier in power electronics. Such a device is also called a power device or, when used in an integrated circuit, a power IC.

A MESFET is a field-effect transistor semiconductor device similar to a JFET with a Schottky (metal–semiconductor) junction instead of a p–n junction for a gate.

<span class="mw-page-title-main">High-electron-mobility transistor</span> Type of field-effect transistor

A high-electron-mobility transistor, also known as heterostructure FET (HFET) or modulation-doped FET (MODFET), is a field-effect transistor incorporating a junction between two materials with different band gaps as the channel instead of a doped region. A commonly used material combination is GaAs with AlGaAs, though there is wide variation, dependent on the application of the device. Devices incorporating more indium generally show better high-frequency performance, while in recent years, gallium nitride HEMTs have attracted attention due to their high-power performance. Like other FETs, HEMTs are used in integrated circuits as digital on-off switches. FETs can also be used as amplifiers for large amounts of current using a small voltage as a control signal. Both of these uses are made possible by the FET’s unique current–voltage characteristics. HEMT transistors are able to operate at higher frequencies than ordinary transistors, up to millimeter wave frequencies, and are used in high-frequency products such as cell phones, satellite television receivers, voltage converters, and radar equipment. They are widely used in satellite receivers, in low power amplifiers and in the defense industry.

<span class="mw-page-title-main">Threshold voltage</span> Minimum source-to-gate voltage for a field effect transistor to be conducting from source to drain

The threshold voltage, commonly abbreviated as Vth or VGS(th), of a field-effect transistor (FET) is the minimum gate-to-source voltage (VGS) that is needed to create a conducting path between the source and drain terminals. It is an important scaling factor to maintain power efficiency.

<span class="mw-page-title-main">Organic field-effect transistor</span> Type of field-effect transistor

An organic field-effect transistor (OFET) is a field-effect transistor using an organic semiconductor in its channel. OFETs can be prepared either by vacuum evaporation of small molecules, by solution-casting of polymers or small molecules, or by mechanical transfer of a peeled single-crystalline organic layer onto a substrate. These devices have been developed to realize low-cost, large-area electronic products and biodegradable electronics. OFETs have been fabricated with various device geometries. The most commonly used device geometry is bottom gate with top drain and source electrodes, because this geometry is similar to the thin-film silicon transistor (TFT) using thermally grown SiO2 as gate dielectric. Organic polymers, such as poly(methyl-methacrylate) (PMMA), can also be used as dielectric. One of the benefits of OFETs, especially compared with inorganic TFTs, is their unprecedented physical flexibility, which leads to biocompatible applications, for instance in the future health care industry of personalized biomedicines and bioelectronics.

<span class="mw-page-title-main">Power MOSFET</span> MOSFET that can handle significant power levels

A power MOSFET is a specific type of metal–oxide–semiconductor field-effect transistor (MOSFET) designed to handle significant power levels. Compared to the other power semiconductor devices, such as an insulated-gate bipolar transistor (IGBT) or a thyristor, its main advantages are high switching speed and good efficiency at low voltages. It shares with the IGBT an isolated gate that makes it easy to drive. They can be subject to low gain, sometimes to a degree that the gate voltage needs to be higher than the voltage under control.

<span class="mw-page-title-main">VMOS</span>

A VMOS transistor is a type of metal–oxide–semiconductor field-effect transistor (MOSFET). VMOS is also used to describe the V-groove shape vertically cut into the substrate material.

In semiconductor electronics fabrication technology, a self-aligned gate is a transistor manufacturing approach whereby the gate electrode of a MOSFET is used as a mask for the doping of the source and drain regions. This technique ensures that the gate is naturally and precisely aligned to the edges of the source and drain.

<span class="mw-page-title-main">Multigate device</span> MOS field-effect transistor with more than one gate

A multigate device, multi-gate MOSFET or multi-gate field-effect transistor (MuGFET) refers to a metal–oxide–semiconductor field-effect transistor (MOSFET) that has more than one gate on a single transistor. The multiple gates may be controlled by a single gate electrode, wherein the multiple gate surfaces act electrically as a single gate, or by independent gate electrodes. A multigate device employing independent gate electrodes is sometimes called a multiple-independent-gate field-effect transistor (MIGFET). The most widely used multi-gate devices are the FinFET and the GAAFET, which are non-planar transistors, or 3D transistors.

A transistor is a semiconductor device with at least three terminals for connection to an electric circuit. In the common case, the third terminal controls the flow of current between the other two terminals. This can be used for amplification, as in the case of a radio receiver, or for rapid switching, as in the case of digital circuits. The transistor replaced the vacuum-tube triode, also called a (thermionic) valve, which was much larger in size and used significantly more power to operate. The first transistor was successfully demonstrated on December 23, 1947, at Bell Laboratories in Murray Hill, New Jersey. Bell Labs was the research arm of American Telephone and Telegraph (AT&T). The three individuals credited with the invention of the transistor were William Shockley, John Bardeen and Walter Brattain. The introduction of the transistor is often considered one of the most important inventions in history.

The gate oxide is the dielectric layer that separates the gate terminal of a MOSFET from the underlying source and drain terminals as well as the conductive channel that connects source and drain when the transistor is turned on. Gate oxide is formed by thermal oxidation of the silicon of the channel to form a thin insulating layer of silicon dioxide. The insulating silicon dioxide layer is formed through a process of self-limiting oxidation, which is described by the Deal–Grove model. A conductive gate material is subsequently deposited over the gate oxide to form the transistor. The gate oxide serves as the dielectric layer so that the gate can sustain as high as 1 to 5 MV/cm transverse electric field in order to strongly modulate the conductance of the channel.

A carbon nanotube field-effect transistor (CNTFET) is a field-effect transistor that utilizes a single carbon nanotube (CNT) or an array of carbon nanotubes as the channel material, instead of bulk silicon, as in the traditional MOSFET structure. There have been major developments since CNTFETs were first demonstrated in 1998.

<span class="mw-page-title-main">Depletion and enhancement modes</span> Two major types of field effect transistors

In field-effect transistors (FETs), depletion mode and enhancement mode are two major transistor types, corresponding to whether the transistor is in an on state or an off state at zero gate–source voltage.

<span class="mw-page-title-main">FET amplifier</span>

A FET amplifier is an amplifier that uses one or more field-effect transistors (FETs). The most common type of FET amplifier is the MOSFET amplifier, which uses metal–oxide–semiconductor FETs (MOSFETs). The main advantage of a FET used for amplification is that it has very high input impedance and low output impedance.

<span class="mw-page-title-main">QFET</span> Type of MOSFET

A quantum field-effect transistor (QFET) or quantum-well field-effect transistor (QWFET) is a type of MOSFET that takes advantage of quantum tunneling to greatly increase the speed of transistor operation by eliminating the traditional transistor's area of electron conduction which typically causes carriers to slow down by a factor of 3000. The result is an increase in logic speed by a factor of 10 with a simultaneous reduction in component power requirement and size also by a factor of 10. It achieves these things through a manufacturing process known as rapid thermal processing (RTP) that uses ultrafine layers of construction materials.

References

  1. Lilienfeld, J.E. "Method and apparatus for controlling electric current" Archived 2022-04-09 at the Wayback Machine US Patent no. 1,745,175 (filed: 8 October 1926 ; issued: 28 January 1930).
  2. 1 2 3 Lee, Thomas H. (2003). The Design of CMOS Radio-Frequency Integrated Circuits (PDF). Cambridge University Press. ISBN   978-1-139-64377-1. Archived from the original (PDF) on 2019-12-09. Retrieved 2019-07-20.
  3. 1 2 Puers, Robert; Baldi, Livio; Voorde, Marcel Van de; Nooten, Sebastiaan E. van (2017). Nanoelectronics: Materials, Devices, Applications, 2 Volumes. John Wiley & Sons. p. 14. ISBN   978-3-527-34053-8.
  4. Grundmann, Marius (2010). The Physics of Semiconductors. Springer-Verlag. ISBN   978-3-642-13884-3.
  5. Nishizawa, Jun-Ichi (1982). "Junction Field-Effect Devices". In Sittig, Roland; Roggwiller, P. (eds.). Semiconductor Devices for Power Conditioning. Springer. pp. 241–272. doi:10.1007/978-1-4684-7263-9_11. ISBN   978-1-4684-7265-3.
  6. 1 2 3 4 Moskowitz, Sanford L. (2016). Advanced Materials Innovation: Managing Global Technology in the 21st century. John Wiley & Sons. p. 168. ISBN   978-0-470-50892-3.
  7. "The Foundation of Today's Digital World: The Triumph of the MOS Transistor". Computer History Museum. 13 July 2010. Retrieved 21 July 2019.
  8. 1 2 Howard R. Duff (2001). "John Bardeen and transistor physics". AIP Conference Proceedings. Vol. 550. pp. 3–32. doi: 10.1063/1.1354371 .
  9. Hans Camenzind (2005). Designing Analog Chips.
  10. Massoud, Hisham Z. (1997). ULSI Science and Technology/1997. The Electrochemical Society. p. 43. ISBN   978-1-56677-130-6.
  11. Lillian Hoddeson (1994). "Research on crystal rectifiers during World War II and the invention of the transistor". History and Technology. 11 (2): 121–130. doi:10.1080/07341519408581858.
  12. Michael Riordan; Lillian Hoddeson (1997). Crystal Fire: The Birth of the Information Age. W. W. Norton & Company. ISBN   978-0-393-04124-8.
  13. Christophe Lécuyer; David C. Brook; Jay Last (2010). Makers of the Microchip: A Documentary History of Fairchild Semiconductor. MIT Press. pp. 62–63. ISBN   978-0-262-01424-3.
  14. Claeys, Cor L. (2003). ULSI Process Integration III: Proceedings of the International Symposium. The Electrochemical Society. pp. 27–30. ISBN   978-1-56677-376-8.
  15. Lojek, Bo (2007). History of Semiconductor Engineering. Springer Science & Business Media. p. 324. ISBN   978-3-540-34258-8.
  16. Stefan Ferdinand Müller (2016). Development of HfO2-Based Ferroelectric Memories for Future CMOS Technology Nodes. ISBN   978-3-7392-4894-3.
  17. B.G Lowe; R.A. Sareen (2013). Semiconductor X-Ray Detectors. CRC Press. ISBN   978-1-4665-5401-6.
  18. Bassett, Ross Knox (2007). To the Digital Age: Research Labs, Start-up Companies, and the Rise of MOS Technology. Johns Hopkins University Press. p. 22. ISBN   978-0-8018-8639-3.
  19. "Martin Atalla in Inventors Hall of Fame, 2009" . Retrieved 21 June 2013.
  20. 1 2 "Dawon Kahng". National Inventors Hall of Fame . Retrieved 27 June 2019.
  21. "1960 - Metal Oxide Semiconductor (MOS) Transistor Demonstrated". The Silicon Engine. Computer History Museum.
  22. 1 2 Lojek, Bo (2007). History of Semiconductor Engineering . Springer Science & Business Media. pp.  321–3. ISBN   978-3-540-34258-8.
  23. "960 - Metal Oxide Semiconductor (MOS) Transistor Demonstrated". The Silicon Engine. Computer History Museum.
  24. Motoyoshi, M. (2009). "Through-Silicon Via (TSV)". Proceedings of the IEEE. 97 (1): 43–48. doi:10.1109/JPROC.2008.2007462. ISSN   0018-9219. S2CID   29105721.
  25. "Transistors Keep Moore's Law Alive". EETimes . 12 December 2018. Retrieved 18 July 2019.
  26. "Who Invented the Transistor?". Computer History Museum . 4 December 2013. Retrieved 20 July 2019.
  27. Duncan, Ben (1996). High Performance Audio Power Amplifiers. Elsevier. p. 177. ISBN   978-0-08-050804-7.
  28. 1 2 "Remarks by Director Iancu at the 2019 International Intellectual Property Conference". United States Patent and Trademark Office . June 10, 2019. Retrieved 20 July 2019.
  29. "1963: Complementary MOS Circuit Configuration is Invented". Computer History Museum . Retrieved 6 July 2019.
  30. U.S. patent 3,102,230 , filed in 1960, issued in 1963
  31. D. Kahng and S. M. Sze, "A floating gate and its application to memory devices", The Bell System Technical Journal, vol. 46, no. 4, 1967, pp. 1288–1295
  32. Colinge, J.P. (2008). FinFETs and Other Multi-Gate Transistors. Springer Science & Business Media. p. 11. ISBN   978-0-387-71751-7.
  33. Sekigawa, Toshihiro; Hayashi, Yutaka (1 August 1984). "Calculated threshold-voltage characteristics of an XMOS transistor having an additional bottom gate". Solid-State Electronics. 27 (8): 827–828. Bibcode:1984SSEle..27..827S. doi:10.1016/0038-1101(84)90036-4. ISSN   0038-1101.
  34. "IEEE Andrew S. Grove Award Recipients". IEEE Andrew S. Grove Award . Institute of Electrical and Electronics Engineers . Retrieved 4 July 2019.
  35. "The Breakthrough Advantage for FPGAs with Tri-Gate Technology" (PDF). Intel. 2014. Retrieved 4 July 2019.
  36. Jacob Millman (1985). Electronic devices and circuits. Singapore: McGraw-Hill International. p. 397. ISBN   978-0-07-085505-2.
  37. 1 2 Jacob Millman (1985). Electronic devices and circuits. Singapore: McGraw-Hill. pp. 384–385. ISBN   978-0-07-085505-2.
  38. Galup-Montoro, C.; Schneider, M.C. (2007). MOSFET modeling for circuit analysis and design . London/Singapore: World Scientific. p.  83. ISBN   978-981-256-810-6.
  39. Norbert R Malik (1995). Electronic circuits: analysis, simulation, and design. Englewood Cliffs, NJ: Prentice Hall. pp. 315–316. ISBN   978-0-02-374910-0.
  40. Spencer, R.R.; Ghausi, M.S. (2001). Microelectronic circuits. Upper Saddle River NJ: Pearson Education/Prentice-Hall. p. 102. ISBN   978-0-201-36183-4.
  41. Sedra, A. S.; Smith, K.C. (2004). Microelectronic circuits (Fifth ed.). New York: Oxford University Press. p.  552. ISBN   978-0-19-514251-8.
  42. PR Gray; PJ Hurst; SH Lewis; RG Meyer (2001). Analysis and design of analog integrated circuits (Fourth ed.). New York: Wiley. pp. §1.5.2 p. 45. ISBN   978-0-471-32168-2.
  43. Bob Yirka (10 January 2011). "IBM creates first graphene based integrated circuit". Phys.org . Retrieved 14 January 2019.
  44. Lin, Y.-M.; Valdes-Garcia, A.; Han, S.-J.; Farmer, D. B.; Sun, Y.; Wu, Y.; Dimitrakopoulos, C.; Grill, A; Avouris, P; Jenkins, K. A. (2011). "Wafer-Scale Graphene Integrated Circuit". Science. 332 (6035): 1294–1297. Bibcode:2011Sci...332.1294L. doi:10.1126/science.1204428. PMID   21659599. S2CID   3020496.
  45. Belle Dumé (10 December 2012). "Flexible graphene transistor sets new records". Physics World. Retrieved 14 January 2019.
  46. Schöning, Michael J.; Poghossian, Arshak (2002). "Recent advances in biologically sensitive field-effect transistors (BioFETs)" (PDF). Analyst. 127 (9): 1137–1151. Bibcode:2002Ana...127.1137S. doi:10.1039/B204444G. PMID   12375833.
  47. freepatentsonline.com, HIGFET and method - Motorola]
  48. Ionescu, A. M.; Riel, H. (2011). "Tunnel field-effect transistors as energy-efficient electronic switches". Nature . 479 (7373): 329–337. Bibcode:2011Natur.479..329I. doi:10.1038/nature10679. PMID   22094693. S2CID   4322368.
  49. Dumé, Isabelle (12 December 2018). "Topological off-on switch could make new type of transistor". Physics World. IOP Publishing. Retrieved 16 January 2022.
  50. "Organic transistor paves way for new generations of neuro-inspired computers". ScienceDaily . January 29, 2010. Retrieved January 14, 2019.
  51. Sarvari H.; Ghayour, R.; Dastjerdy, E. (2011). "Frequency analysis of graphene nanoribbon FET by Non-Equilibrium Green's Function in mode space". Physica E: Low-dimensional Systems and Nanostructures. 43 (8): 1509–1513. Bibcode:2011PhyE...43.1509S. doi:10.1016/j.physe.2011.04.018.
  52. Jerzy Ruzyllo (2016). Semiconductor Glossary: A Resource for Semiconductor Community. World Scientific. p. 244. ISBN   978-981-4749-56-5.
  53. Appenzeller J, et al. (November 2008). "Toward Nanowire Electronics". IEEE Transactions on Electron Devices. 55 (11): 2827–2845. Bibcode:2008ITED...55.2827A. doi:10.1109/ted.2008.2008011. ISSN   0018-9383. OCLC   755663637. S2CID   703393.
  54. Prakash, Abhijith; Ilatikhameneh, Hesameddin; Wu, Peng; Appenzeller, Joerg (2017). "Understanding contact gating in Schottky barrier transistors from 2D channels". Scientific Reports. 7 (1): 12596. arXiv: 1707.01459 . Bibcode:2017NatSR...712596P. doi:10.1038/s41598-017-12816-3. ISSN   2045-2322. OCLC   1010581463. PMC   5626721 . PMID   28974712.
  55. Miklos, Bolza. "What Are Graphene Field Effect Transistors (GFETs)?". Graphenea. Retrieved 14 January 2019.
  56. IBM Research Unveils 'VTFET': A Revolutionary New Chip Architecture Which is Two Times the Performance finFET Dec 2021
  57. VIII.5. Noise in Transistors
  58. Allen Mottershead (2004). Electronic devices and siraj circuits. New Delhi: Prentice-Hall of India. ISBN   978-81-203-0124-5.
  59. Bhalla, Anup (2021-09-17). "Origins of SiC FETs and Their Evolution Toward the Perfect Switch". Power Electronics News. Retrieved 2022-01-21.
  60. Slow Body Diode Failures of Field Effect Transistors (FETs): A Case Study.
  61. Sporea, R.A.; Trainor, M.J.; Young, N.D.; Silva, S.R.P. (2014). "Source-gated transistors for order-of-magnitude performance improvements in thin-film digital circuits". Scientific Reports. 4: 4295. Bibcode:2014NatSR...4E4295S. doi:10.1038/srep04295. PMC   3944386 . PMID   24599023.