Stochastic Process
Stochastic Process
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
From Wikipedia, the free encyclopedia
Diagram of a copper cathode in a galvanic cell (e.g., a battery). Positively charged cations move
towards the cathode allowing a positive current i to flow out of the cathode.
(Top)
Formula
Justification
Occurrences
Gravitation
Electrostatics
Light and other electromagnetic radiation
Example
Sound in a gas
Inverse-square law
31 languages
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
Color (beta)
Automatic
Light
Dark
Report an issue with dark mode
In science, an inverse-square law is any scientific law stating that the observed
"intensity" of a specified physical quantity is inversely proportional to the square of the
distance from the source of that physical quantity. The fundamental cause for this can
be understood as geometric dilution corresponding to point-source radiation into three-
dimensional space.
Radar energy expands during both the signal transmission and the reflected return, so
the inverse square for both paths means that the radar will receive energy according to
the inverse fourth power of the range.
To prevent dilution of energy while propagating a signal, certain methods can be used
such as a waveguide, which acts like a canal does for water, or how a gun barrel
restricts hot gas expansion to one dimension in order to prevent loss of energy transfer
to a bullet.
Formula[edit]
In mathematical notation the inverse square law can be expressed as an intensity (I)
varying as a function of distance (d) from some centre. The intensity is proportional (see
∝) to the reciprocal of the square of the distance thus:
intensity ∝ 1distance2
intensity1intensity2=distance22distance12
intensity1×distance12=intensity2×distance22
The divergence of a vector field which is the resultant of radial inverse-square law fields
with respect to one or more sources is proportional to the strength of the local sources,
and hence zero outside sources. Newton's law of universal gravitation follows an
inverse-square law, as do the effects of electric, light, sound, and radiation phenomena.
Justification[edit]
The inverse-square law generally applies when some force, energy, or other conserved
quantity is evenly radiated outward from a point source in three-dimensional space.
2
Since the surface area of a sphere (which is 4πr ) is proportional to the square of the
radius, as the emitted radiation gets farther from the source, it is spread out over an
area that is increasing in proportion to the square of the distance from the source.
Hence, the intensity of radiation passing through any unit area (directly facing the point
source) is inversely proportional to the square of the distance from the point source.
Gauss's law for gravity is similarly applicable, and can be used with any physical
quantity that acts in accordance with the inverse-square relationship.
Occurrences[edit]
Gravitation[edit]
Gravitation is the attraction between objects that have mass. Newton's law states:
The gravitational attraction force between two point masses is directly proportional to
the product of their masses and inversely proportional to the square of their separation
[citation
distance. The force is always attractive and acts along the line joining them.
needed]
If the distribution of matter in each body is spherically symmetric, then the objects can
be treated as point masses without approximation, as shown in the shell theorem.
Otherwise, if we want to calculate the attraction between massive bodies, we need to
add all the point-point attraction forces vectorially and the net attraction might not be
exact inverse square. However, if the separation between the massive bodies is much
larger compared to their sizes, then to a good approximation, it is reasonable to treat
the masses as a point mass located at the object's center of mass while calculating the
gravitational force.
As the law of gravitation, this law was suggested in 1645 by Ismaël Bullialdus. But
Bullialdus did not accept Kepler's second and third laws, nor did he appreciate
Christiaan Huygens's solution for circular motion (motion in a straight line pulled aside
by the central force). Indeed, Bullialdus maintained the sun's force was attractive at
aphelion and repulsive at perihelion. Robert Hooke and Giovanni Alfonso Borelli both
[1]
expounded gravitation in 1666 as an attractive force. Hooke's lecture "On gravity" was
[2]
at the Royal Society, in London, on 21 March. Borelli's "Theory of the Planets" was
[3]
published later in 1666. Hooke's 1670 Gresham lecture explained that gravitation
applied to "all celestiall bodys" and added the principles that the gravitating power
decreases with distance and that in the absence of any such power bodies move in
straight lines. By 1679, Hooke thought gravitation had inverse square dependence and
[4]
communicated this in a letter to Isaac Newton: my supposition is that the attraction
[5]
always is in duplicate proportion to the distance from the center reciprocall.
Hooke remained bitter about Newton claiming the invention of this principle, even
though Newton's 1686 Principia acknowledged that Hooke, along with Wren and Halley,
[6]
had separately appreciated the inverse square law in the solar system, as well as
[7]
giving some credit to Bullialdus.
Electrostatics[edit]
F=keq1q2r2
The intensity (or illuminance or irradiance) of light or other linear waves radiating from a
point source (energy per unit of area perpendicular to the source) is inversely
proportional to the square of the distance from the source, so an object (of the same
size) twice as far away receives only one-quarter the energy (in the same time period).
More generally, the irradiance, i.e., the intensity (or power per unit area in the direction
of propagation), of a spherical wavefront varies inversely with the square of the distance
from the source (assuming there are no losses caused by absorption or scattering).
For example, the intensity of radiation from the Sun is 9126 watts per square meter at
the distance of Mercury (0.387 AU); but only 1367 watts per square meter at the
distance of Earth (1 AU)—an approximate threefold increase in distance results in an
approximate ninefold decrease in intensity of radiation.
For non-isotropic radiators such as parabolic antennas, headlights, and lasers, the
effective origin is located far behind the beam aperture. If you are close to the origin,
you don't have to go far to double the radius, so the signal drops quickly. When you are
far from the origin and still have a strong signal, like with a laser, you have to travel very
far to double the radius and reduce the signal. This means you have a stronger signal
or have antenna gain in the direction of the narrow beam relative to a wide beam in all
directions of an isotropic antenna.
In photography and stage lighting, the inverse-square law is used to determine the “fall
off” or the difference in illumination on a subject as it moves closer to or further from the
light source. For quick approximations, it is enough to remember that doubling the
[9]
distance reduces illumination to one quarter; or similarly, to halve the illumination
increase the distance by a factor of 1.4 (the square root of 2), and to double
illumination, reduce the distance to 0.7 (square root of 1/2). When the illuminant is not a
point source, the inverse square rule is often still a useful approximation; when the size
of the light source is less than one-fifth of the distance to the subject, the calculation
[10]
error is less than 1%.
The fractional reduction in electromagnetic fluence (Φ) for indirectly ionizing radiation
with increasing distance from a point source can be calculated using the inverse-square
law. Since emissions from a point source have radial directions, they intercept at a
2
perpendicular incidence. The area of such a shell is 4πr where r is the radial distance
from the center. The law is particularly important in diagnostic radiography and
radiotherapy treatment planning, though this proportionality does not hold in practical
situations unless source dimensions are much smaller than the distance. As stated in
Fourier theory of heat “as the point source is magnification by distances, its radiation is
dilute proportional to the sin of the angle, of the increasing circumference arc from the
point of origin”.
Example[edit]
Let P be the total power radiated from a point source (for example, an omnidirectional
isotropic radiator). At large distances from the source (compared to the size of the
source), this power is distributed over larger and larger spherical surfaces as the
distance from the source increases. Since the surface area of a sphere of radius r is A =
2
4πr , the intensity I (power per unit area) of radiation at distance r is
I=PA=P4πr2.
Sound in a gas[edit]
In acoustics, the sound pressure of a spherical wavefront radiating from a point source
decreases by 50% as the distance r is doubled; measured in dB, the decrease is still
6.02 dB, since dB represents an intensity ratio. The pressure ratio (as opposed to
power ratio) is not inverse-square, but is inverse-proportional (inverse distance law):
p ∝ 1r
p
:
v ∝1r
In the near field is a quadrature component of the particle velocity that is 90° out of
phase with the sound pressure and does not contribute to the time-averaged energy or
the intensity of the sound. The sound intensity is the product of the RMS sound
pressure and the in-phase component of the RMS particle velocity, both of which are
inverse-proportional. Accordingly, the intensity follows an inverse-square behaviour:
I = pv ∝ 1r2.
I∝1rn−1,
[citation needed]
given that the space outside the source is divergence free.
Non-Euclidean implications[edit]
The inverse-square law, fundamental in Euclidean spaces, also applies to non-
Euclidean geometries, including hyperbolic space. The inherent curvature in these
spaces impacts physical laws, underpinning various fields such as cosmology, general
[11]
relativity, and string theory.
John D. Barrow, in his 2020 paper "Non-Euclidean Newtonian Cosmology," elaborates on the
behavior of force (F) and potential (Φ) within hyperbolic 3-space (H3). He illustrates that F and
Φ obey the formulas F ∝ 1 / R^2 sinh^2(r/R) and Φ ∝ coth(r/R), where R and r represent the
[11]
curvature radius and the distance from the focal point, respectively.
Within the realm of non-Euclidean geometries and general relativity, deviations from the
inverse-square law might not stem from the law itself but rather from the assumption
that the force between bodies depends instantaneously on distance, contradicting
special relativity. General relativity instead interprets gravity as a distortion of
spacetime, causing freely falling particles to traverse geodesics in this curved
[13]
spacetime.
History[edit]
John Dumbleton of the 14th-century Oxford Calculators, was one of the first to express
functional relationships in graphical form. He gave a proof of the mean speed theorem
stating that "the latitude of a uniformly difform movement corresponds to the degree of
the midpoint" and used this method to study the quantitative decrease in intensity of
illumination in his Summa logicæ et philosophiæ naturalis (ca. 1349), stating that it was
not linearly proportional to the distance, but was unable to expose the Inverse-square
[14]
law.
German astronomer Johannes Kepler discussed the inverse-square law and how it affects the
intensity of light.
In 1645, in his book Astronomia Philolaica ..., the French astronomer Ismaël Bullialdus
[17]
(1605–1694) refuted Johannes Kepler's suggestion that "gravity" weakens as the
inverse of the distance; instead, Bullialdus argued, "gravity" weakens as the inverse
[18][19]
square of the distance:
In England, the Anglican bishop Seth Ward (1617–1689) publicized the ideas of
Bullialdus in his critique In Ismaelis Bullialdi astronomiae philolaicae fundamenta
inquisitio brevis (1653) and publicized the planetary astronomy of Kepler in his book
Astronomia geometrica (1656).
In 1663–1664, the English scientist Robert Hooke was writing his book Micrographia
(1666) in which he discussed, among other things, the relation between the height of
the atmosphere and the barometric pressure at the surface. Since the atmosphere
surrounds the Earth, which itself is a sphere, the volume of atmosphere bearing on any
unit area of the Earth's surface is a truncated cone (which extends from the Earth's
center to the vacuum of space; obviously only the section of the cone from the Earth's
surface Waveguide
29 languages
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
Color (beta)
Automatic
Light
Dark
Report an issue with dark mode
Electric field Ex component of the TE31 mode inside an x-band hollow metal waveguide.
Without the physical constraint of a waveguide, waves would expand into three-
dimensional space and their intensities would decrease according to the inverse square
law.
There are different types of waveguides for different types of waves. The original and
most common meaning is a hollow conductive metal pipe used to carry high frequency
[1]
radio waves, particularly microwaves. Dielectric waveguides are used at higher radio
frequencies, and transparent dielectric waveguides and optical fibers serve as
waveguides for light. In acoustics, air ducts and horns are used as waveguides for
sound in musical instruments and loudspeakers, and specially-shaped metal rods
conduct ultrasonic waves in ultrasonic machining.
The geometry of a waveguide reflects its function; in addition to more common types
that channel the wave in one dimension, there are two-dimensional slab waveguides
which confine waves to two dimensions. The frequency of the transmitted wave also
dictates the size of a waveguide: each waveguide has a cutoff wavelength determined
by its size and will not conduct waves of greater wavelength; an optical fiber that guides
light will not transmit microwaves which have a much larger wavelength. Some naturally
occurring structures can also act as waveguides. The SOFAR channel layer in the
[2]
ocean can guide the sound of whale song across enormous distances. Any shape of
cross section of waveguide can support EM waves. Irregular shapes are difficult to
analyse. Commonly used waveguides are rectangular and circular in shape.
Uses[edit]
Waveguide supplying power for the Argonne National Laboratory Advanced Photon Source.
The uses of waveguides for transmitting signals were known even before the term was
coined. The phenomenon of sound waves guided through a taut wire have been known
for a long time, as well as sound through a hollow pipe such as a cave or medical
stethoscope. Other uses of waveguides are in transmitting power between the
components of a system such as radio, radar or optical devices. Waveguides are the
fundamental principle of guided wave testing (GWT), one of the many methods of non-
[3]
destructive evaluation.
Specific examples:
● Optical fibers transmit light and signals for long distances with low attenuation
and a wide usable range of wavelengths.
● In a microwave oven a waveguide transfers power from the magnetron,
where waves are formed, to the cooking chamber.
● In a radar, a waveguide transfers radio frequency energy to and from the
antenna, where the impedance needs to be matched for efficient power
transmission (see below).
● Rectangular and circular waveguides are commonly used to connect feeds of
parabolic dishes to their electronics, either low-noise receivers or power
amplifier/transmitters.
● Waveguides are used in scientific instruments to measure optical, acoustic
and elastic properties of materials and objects. The waveguide can be put in
contact with the specimen (as in a medical ultrasonography), in which case
the waveguide ensures that the power of the testing wave is conserved, or
the specimen may be put inside the waveguide (as in a dielectric constant
measurement, so that smaller objects can be tested and the accuracy is
[4]
better.
[5]
● A transmission line is a commonly used specific type of waveguide.
History[edit]
This section
duplicates the
scope of other
articles,
specifically
Waveguide
(electromagnetis
m)#History.
Please discuss
this issue and
help introduce a
summary style
to the section by
replacing the
section with a
link and a
summary or by
splitting the
content into a
new article.
(November
2020)
The first structure for guiding waves was proposed by J. J. Thomson in 1893, and was
first experimentally tested by Oliver Lodge in 1894. The first mathematical analysis of
[6]: 8
electromagnetic waves in a metal cylinder was performed by Lord Rayleigh in 1897.
For sound waves, Lord Rayleigh published a full mathematical analysis of propagation
[7]
modes in his seminal work, "The Theory of Sound". Jagadish Chandra Bose
researched millimeter wavelengths using waveguides, and in 1897 described to the
[8][9]
Royal Institution in London his research carried out in Kolkata.
The study of dielectric waveguides (such as optical fibers, see below) began as early as
the 1920s, by several people, most famous of which are Rayleigh, Sommerfeld and
[10]
Debye. Optical fiber began to receive special attention in the 1960s due to its
importance to the communications industry.
R. Carson and Sallie P. Mead. This work led to the discovery that for the TE01 mode in
circular waveguide losses go down with frequency and at one time this was a serious
[11]: 544–548
contender for the format for long-distance telecommunications.
The importance of radar in World War II gave a great impetus to waveguide research,
at least on the Allied side. The magnetron, developed in 1940 by John Randall and
Harry Boot at the University of Birmingham in the United Kingdom, provided a good
power source and made microwave radar feasible. The most important centre of US
research was at the Radiation Laboratory (Rad Lab) at MIT but many others took part in
the US, and in the UK such as the Telecommunications Research Establishment. The
head of the Fundamental Development Group at Rad Lab was Edward Mills Purcell. His
researchers included Julian Schwinger, Nathan Marcuvitz, Carol Gray Montgomery,
and Robert H. Dicke. Much of the Rad Lab work concentrated on finding lumped
element models of waveguide structures so that components in waveguide could be
analysed with standard circuit theory. Hans Bethe was also briefly at Rad Lab, but while
there he produced his small aperture theory which proved important for waveguide
cavity filters, first developed at Rad Lab. The German side, on the other hand, largely
ignored the potential of waveguides in radar until very late in the war. So much so that
when radar parts from a downed British plane were sent to Siemens & Halske for
analysis, even though they were recognised as microwave components, their purpose
could not be identified.
At that time, microwave techniques were badly neglected in Germany. It was generally
believed that it was of no use for electronic warfare, and those who wanted to do
research work in this field were not allowed to do so.
German academics were even allowed to continue publicly publishing their research in
[12]: 548–554 [13]: 1055, 1057
this field because it was not felt to be important.
Immediately after World War II waveguide was the technology of choice in the
microwave field. However, it has some problems; it is bulky, expensive to produce, and
the cutoff frequency effect makes it difficult to produce wideband devices. Ridged
waveguide can increase bandwidth beyond an octave, but a better solution is to use a
technology working in TEM mode (that is, non-waveguide) such as coaxial conductors
since TEM does not have a cutoff frequency. A shielded rectangular conductor can also
be used and this has certain manufacturing advantages over coax and can be seen as
the forerunner of the planar technologies (stripline and microstrip). However, planar
technologies really started to take off when printed circuits were introduced. These
methods are significantly cheaper than waveguide and have largely taken its place in
most bands. However, waveguide is still favoured in the higher microwave bands from
[12]: 556–557 [14]: 21–27, 21–50
around Ku band upwards.
Properties[edit]
Propagation modes and cutoff frequencies[edit]
A propagation mode in a waveguide is one solution of the wave equations, or, in other
[10]
words, the form of the wave. Due to the constraints of the boundary conditions, there
are only limited frequencies and forms for the wave function which can propagate in the
waveguide. The lowest frequency in which a certain mode can propagate is the cutoff
frequency of that mode. The mode with the lowest cutoff frequency is the fundamental
[15]:
mode of the waveguide, and its cutoff frequency is the waveguide cutoff frequency.
38
Propagation modes are computed by solving the Helmholtz equation alongside a set of
boundary conditions depending on the geometrical shape and materials bounding the
region. The usual assumption for infinitely long uniform waveguides allows us to
assume a propagating form for the wave, i.e. stating that every field component has a
known dependency on the propagation direction (i.e.
). More specifically, the common approach is to first replace all unknown time-varying
fields
u(x,y,z,t)
U(x,y,z)
, (angular frequency
ω=2πf
U(x,y,z)=U^(x,y)e−γz
, where the
term represents the propagation constant (still unknown) along the direction along
which the waveguide extends to infinity. The Helmholtz equation can be rewritten to
accommodate such form and the resulting equality needs to be solved for
and
U^(x,y)
U^(x,y)γ
[16]
for each solution of the former.
The propagation constant
of the guided wave is complex, in general. For a lossless case, the propagation
constant might be found to take on either real or imaginary values, depending on the
chosen solution of the eigenvalue equation and on the angular frequency
. When
is purely real, the mode is said to be "below cutoff", since the amplitude of the field
phasors tends to exponentially decrease with propagation; an imaginary
[17]
.
Impedance matching[edit]
[10]
). A waveguide in circuit theory is described by a transmission line having a length
[18]: 2–3, 6–12 [19]: 14 [20]
and characteristic impedance. In other words, the impedance
indicates the ratio of voltage to current of the circuit component (in this case a
waveguide) during propagation of the wave. This description of the waveguide was
originally intended for alternating current, but is also suitable for electromagnetic and
sound waves, once the wave and material properties (such as pressure, density,
dielectric constant) are properly converted into electrical terms (current and impedance
[21]: 14
for example).
Γ=Z2−Z1Z2+Z1
, where
(Gamma) is the reflection coefficient (0 denotes full transmission, 1 full reflection, and
0.5 is a reflection of half the incoming voltage),
Z1
and
Z2
are the impedance of the first component (from which the wave enters) and the
[22]
second component, respectively.
An impedance mismatch creates a reflected wave, which added to the incoming waves
creates a standing wave. An impedance mismatch can be also quantified with the
standing wave ratio (SWR or VSWR for voltage), which is connected to the impedance
ratio and reflection coefficient by:
VSWR=|V|max|V|min=1+|Γ|1−|Γ|
, where
|V|min/max
are the minimum and maximum values of the voltage absolute value, and
the VSWR is the voltage standing wave ratio, which value of 1 denotes full
transmission, without reflection and thus no standing wave, while very large values
[20]
mean high reflection and standing wave pattern.
Electromagnetic waveguides[edit]
Radio-frequency waveguides[edit]
In this military radar, microwave radiation is transmitted between the source and the reflector by
a waveguide. The figure suggests that microwaves leave the box in a circularly symmetric mode
(allowing the antenna to rotate), then they are converted to a linear mode, and pass through a
flexible stage. Their polarisation is then rotated in a twisted stage and finally they irradiate the
parabolic antenna.
Optical waveguides[edit]
Other types of optical waveguide are also used, including photonic-crystal fiber, which
guides waves by any of several distinct mechanisms. Guides in the form of a hollow
tube with a highly reflective inner surface have also been used as light pipes for
illumination applications. The inner surfaces may be polished metal, or may be covered
with a multilayer film
nts that are used in design and configuration of automated systems in areas such as
electrical and pneumatic domains, and the control of quantities being measured. They
typically work for industries
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
Color (beta)
Automatic
Light
Dark
Report an issue with dark mode
From Wikipedia, the free encyclopedia
SWR of a vertical HB9XBG Antenna for the 40m-band as a function of frequency
[1][2]
Voltage standing wave ratio (VSWR) (pronounced "vizwar" ) is the ratio of
maximum to minimum voltage on a transmission line . For example, a VSWR of 1.2
means a peak voltage 1.2 times the minimum voltage along that line, if the line is at
least one half wavelength long.
A SWR can be also defined as the ratio of the maximum amplitude to minimum
amplitude of the transmission line's currents, electric field strength, or the magnetic field
strength. Neglecting transmission line loss, these ratios are identical.
[3]
The power standing wave ratio (PSWR) is defined as the square of the VSWR,
however, this deprecated term has no direct physical relation to power actually involved
in transmission.
SWR is usually measured using a dedicated instrument called an SWR meter. Since
SWR is a measure of the load impedance relative to the characteristic impedance of the
transmission line in use (which together determine the reflection coefficient as
described below), a given SWR meter can interpret the impedance it sees in terms of
SWR only if it has been designed for the same particular characteristic impedance as
the line. In practice most transmission lines used in these applications are coaxial
cables with an impedance of either 50 or 75 ohms, so most SWR meters correspond to
one of these.
Checking the SWR is a standard procedure in a radio station. Although the same
information could be obtained by measuring the load's impedance with an impedance
analyzer (or "impedance bridge"), the SWR meter is simpler and more robust for this
purpose. By measuring the magnitude of the impedance mismatch at the transmitter
output it reveals problems due to either the antenna or the transmission line.
Impedance matching[edit]
Main article: Impedance matching
Such a mismatch is usually undesired and results in standing waves along the
transmission line which magnifies transmission line losses (significant at higher
frequencies and for longer cables). The SWR is a measure of the depth of those
standing waves and is, therefore, a measure of the matching of the load to the
transmission line. A matched load would result in an SWR of 1:1 implying no reflected
wave. An infinite SWR represents complete reflection by a load unable to absorb
electrical power, with all the incident power reflected back towards the source.
It should be understood that the match of a load to the transmission line is different from
the match of a source to the transmission line or the match of a source to the load seen
through the transmission line. For instance, if there is a perfect match between the load
*
impedance Zload and the source impedance Zsource = Z load, that perfect match will
remain if the source and load are connected through a transmission line with an
electrical length of one half wavelength (or a multiple of one half wavelengths) using a
transmission line of any characteristic impedance Z0. However the SWR will generally
not be 1:1, depending only on Zload and Z0. With a different length of transmission line,
the source will see a different impedance than Zload which may or may not be a good
match to the source. Sometimes this is deliberate, as when a quarter-wave matching
section is used to improve the match between an otherwise mismatched source and
load.
However typical RF sources such as transmitters and signal generators are designed to
look into a purely resistive load impedance such as 50Ω or 75Ω, corresponding to
common transmission lines' characteristic impedances. In those cases, matching the
load to the transmission line, Zload = Z0, always ensures that the source will see the
same load impedance as if the transmission line weren't there. This is identical to a 1:1
SWR. This condition (Zload = Z0) also means that the load seen by the source is
independent of the transmission line's electrical length. Since the electrical length of a
physical segment of transmission line depends on the signal frequency, violation of this
condition means that the impedance seen by the source through the transmission line
becomes a function of frequency (especially if the line is long), even if Zload is
frequency-independent. So in practice, a good SWR (near 1:1) implies a transmitter's
output seeing the exact impedance it expects for optimum and safe operation.
Incident wave (blue) is fully reflected (red wave) out of phase at short-circuited end of transmission line,
creating a net voltage (black) standing wave. Γ = −1, SWR = ∞.
Standing waves on transmission line, net voltage shown in different colors during one period of
oscillation. Incoming wave from left (amplitude = 1) is partially reflected with (top to bottom) Γ = 0.6,
−0.333, and 0.8 ∠60°. Resulting SWR = 4, 2, 9.
Vf
Vr
).
Γ=VrVf.
or
Γ=ZL−ZoZL+Zo
is a complex number that describes both the magnitude and the phase shift of the
reflection. The simplest cases with
● Γ=−1
● : complete negative reflection, when the line is short-circuited,
● Γ=0
● : no reflection, when the line is perfectly matched,
● Γ=+1
● : complete positive reflection, when the line is open-circuited.
At some points along the line the forward and reflected waves interfere constructively,
exactly in phase, with the resulting amplitude
Vmax
|Vmax|=|Vf|+|Vr|=|Vf|+|ΓVf|=(1+|Γ|)|Vf|.
At other points, the waves interfere 180° out of phase with the amplitudes partially
cancelling:
|Vmin|=|Vf|−|Vr|=|Vf|−|ΓVf|=(1−|Γ|)|Vf|.
VSWR=|Vmax||Vmin|=1+|Γ|1−|Γ|.
always falls in the range [0,1], the SWR is always greater than or equal to unity. Note
that the phase of Vf and Vr vary along the transmission line in opposite directions to
each other. Therefore, the complex-valued reflection coefficient
varies as well, but only in phase. With the SWR dependent only on the complex
magnitude of
, it can be seen that the SWR measured at any point along the transmission line
(neglecting transmission line losses) obtains an identical reading.
Since the power of the forward and reflected waves are proportional to the square of
the voltage components due to each wave, SWR can be expressed in terms of forward
and reflected power:
SWR=1+Pr/Pf1−Pr/Pf.
By sampling the complex voltage and current at the point of insertion, an SWR meter is
able to compute the effective forward and reflected voltages on the transmission line for
the characteristic impedance for which the SWR meter has been designed. Since the
forward and reflected power is related to the square of the forward and reflected
voltages, some SWR meters also display the forward and reflected power.
In the special case of a load RL, which is purely resistive but unequal to the
characteristic impedance of the transmission line Z0, the SWR is given simply by their
ratio:
SWR=max{RLZ0,Z0RL}
with the ratio or its reciprocal is chosen to obtain a value greater than unity.
Vactual=Re(ei2πftV) .
Thus taking the real part of the complex quantity inside the parenthesis, the actual
voltage consists of a sine wave at frequency f with a peak amplitude equal to the
complex magnitude of V, and with a phase given by the phase of the complex V. Then
with the position along a transmission line given by x, with the line ending in a load
located at xo, the complex amplitudes of the forward and reverse waves would be
written as:
Vfwd(x)=e−ik(x−xo)AVrev(x)=Γeik(x−xo)A
for some complex amplitude A (corresponding to the forward wave at xo that some
treatments use phasors where the time dependence is according to
e−i2πft
e+ik(x−xo) .
According to the superposition principle the net voltage present at any point x on the
transmission line is equal to the sum of the voltages due to the forward and reflected
waves:
Vnet(x)=Vfwd(x)+Vrev(x)=e−ik(x−xo)(1+Γei2k(x−xo))A
Since we are interested in the variations of the magnitude of Vnet along the line (as a
function of x), we shall solve instead for the squared magnitude of that quantity, which
simplifies the mathematics. To obtain the squared magnitude we multiply the above
quantity by its complex conjugate:
|Vnet(x)|2=Vnet(x)Vnet∗(x)=e−ik(x−xo)(1+Γei2k(x−xo))Ae+ik(x−xo)
(1+Γ∗e−i2k(x−xo))A∗=[ 1+|Γ|2+2 Re(Γei2k(x−xo)) ]|A|2
Depending on the phase of the third term, the maximum and minimum values of Vnet
(the square root of the quantity in the equations) are
(1+|Γ|)|A|
and
(1−|Γ|)|A| ,
SWR=|Vmax||Vmin|=1+|Γ|1−|Γ|
|Vnet(x)|2
|Vmin|2
and
|Vmax|2
with a period of
2π
2k
2π
[7][8]
following NSSL's research. In Canada, Environment Canada constructed the King
[9]
City station, with a 5 cm research Doppler radar, by 1985; McGill University
dopplerized its radar (J. S. Marshall Radar Observatory) in 1993. This led to a complete
[10]
Canadian Doppler network between 1998 and 2004. France and other European
countries had switched to Doppler networks by the early 2000s. Meanwhile, rapid
advances in computer technology led to algorithms to detect signs of severe weather,
and many applications for media outlets and researchers.
After 2000, research on dual polarization technology moved into operational use,
increasing the amount of information available on precipitation type (e.g. rain vs. snow).
"Dual polarization" means that microwave radiation which is polarized both horizontally
and vertically (with respect to the ground) is emitted. Wide-scale deployment was done
by the end of the decade or the beginning of the next in some countries such as the
[11]
United States, France, and Canada. In April 2013, all United States National Weather
[12]
Service NEXRADs were completely dual-polarized.
Since 2003, the U.S. National Oceanic and Atmospheric Administration has been
experimenting with phased-array radar as a replacement for conventional parabolic
antenna to provide more time resolution in atmospheric sounding. This could be
significant with severe thunderstorms, as their evolution can be better evaluated with
more timely data.
Also in 2003, the National Science Foundation established the Engineering Research
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
Color (beta)
Automatic
Light
Dark
Report an issue with dark mode
From Wikipedia, the free encyclopedia
Type Passive
Electronic symbol
The Schottky diode (named after the German physicist Walter H. Schottky), also
known as Schottky barrier diode or hot-carrier diode, is a semiconductor diode
formed by the junction of a semiconductor with a metal. It has a low forward voltage
drop and a very fast switching action. The cat's-whisker detectors used in the early days
of wireless and metal rectifiers used in early power applications can be considered
primitive Schottky diodes.
When sufficient forward voltage is applied, a current flows in the forward direction. A
silicon p–n diode has a typical forward voltage of 600–700 mV, while the Schottky's
forward voltage is 150–450 mV. This lower forward voltage requirement allows higher
switching speeds and better system efficiency.
Construction[edit]
T
h
i
s
s
e
c
t
i
o
n
n
e
e
d
s
a
d
d
i
t
i
o
n
a
l
c
i
t
a
t
i
o
n
s
f
o
r
v
e
r
i
f
i
c
a
t
i
o
n
.
P
l
e
a
s
e
h
e
l
p
i
m
p
r
o
v
e
t
h
i
s
a
r
t
i
c
l
e
b
y
a
d
d
i
n
g
c
i
t
a
t
i
o
n
s
t
o
r
e
l
i
a
b
l
e
s
o
u
r
c
e
s
i
n
t
h
i
s
s
e
c
t
i
o
n
.
U
n
s
o
u
r
c
e
d
m
a
t
e
r
i
a
l
m
a
y
b
e
c
h
a
l
l
e
n
g
e
d
a
n
d
r
e
m
o
v
e
d
.
(
J
u
l
y
2
0
1
5
)
(
L
e
a
r
n
h
o
w
a
n
d
w
h
e
n
t
o
r
e
m
o
v
e
t
h
i
s
m
e
s
s
a
g
e
)
1N5822 Schottky diode with cut-open packaging. The semiconductor in the center makes a
Schottky barrier against one metal electrode (providing rectifying action) and an ohmic contact
with the other electrode.
HP 5082-2800 Schottky Barrier Diodes for General Purpose Applications
The choice of the combination of the metal and semiconductor determines the forward
voltage of the diode. Both n- and p-type semiconductors can develop Schottky barriers.
However, the p-type typically has a much lower forward voltage. As the reverse leakage
current increases dramatically with lowering the forward voltage, it cannot be too low, so
the usually employed range is about 0.15–0.45 V, and p-type semiconductors are
employed only rarely. Titanium silicide and other refractory silicides, which are able to
withstand the temperatures needed for source/drain annealing in CMOS processes,
usually have too low a forward voltage to be useful, so processes using these silicides
[clarification needed]
therefore usually do not offer Schottky diodes.
With increased doping of the semiconductor, the width of the depletion region drops.
Below a certain width, the charge carriers can tunnel through the depletion region. At
very high doping levels, the junction does not behave as a rectifier any more and
becomes an ohmic contact. This can be used for the simultaneous formation of ohmic
contacts and diodes, as a diode will form between the silicide and lightly doped n-type
region, and an ohmic contact will form between the silicide and the heavily doped n- or
p-type region. Lightly doped p-type regions pose a problem, as the resulting contact has
too high a resistance for a good ohmic contact, but too low a forward voltage and too
high a reverse leakage to make a good diode.
As the edges of the Schottky contact are fairly sharp, a high electric field gradient
occurs around them, which limits how large the reverse breakdown voltage threshold
can be. Various strategies are used, from guard rings to overlaps of metallization to
spread out the field gradient. The guard rings consume valuable die area and are used
primarily for larger higher-voltage diodes, while overlapping metallization is employed
primarily with smaller low-voltage diodes.
For power Schottky diodes, the parasitic resistances of the buried n+ layer and the
epitaxial n-type layer become important. The resistance of the epitaxial layer is more
important than it is for a transistor, as the current must cross its entire thickness.
However, it serves as a distributed ballasting resistor over the entire area of the junction
and, under usual conditions, prevents localized thermal runaway.
In comparison with the power p–n diodes, the Schottky diodes are less rugged. The
junction is in direct contact with the thermally sensitive metallization; a Schottky diode
can therefore dissipate less power than an equivalent-size p–n counterpart with a deep-
buried junction before failing (especially during reverse breakdown). The relative
advantage of the lower forward voltage of Schottky diodes is diminished at higher
[2]
forward currents, where the voltage drop is dominated by the series resistance.
This "instantaneous" switching is not always the case. In higher voltage Schottky
devices, in particular, the guard ring structure needed to control breakdown field
geometry creates a parasitic p–n diode with the usual recovery time attributes. As long
as this guard ring diode is not forward biased, it adds only capacitance. If the Schottky
junction is driven hard enough however, the forward voltage eventually will bias both
diodes forward and actual trr will be greatly impacted.
It is often said that the Schottky diode is a "majority carrier" semiconductor device. This
means that if the semiconductor body is a doped n-type, only the n-type carriers (mobile
electrons) play a significant role in the normal operation of the device. The majority
carriers are quickly injected into the conduction band of the metal contact on the other
side of the diode to become free moving electrons. Therefore, no slow random
recombination of n and p-type carriers is involved, so that this diode can cease
conduction faster than an ordinary p–n rectifier diode. This property, in turn, allows a
smaller device area, which also makes for a faster transition. This is another reason
why Schottky diodes are useful in switch-mode power converters: the high speed of the
diode means that the circuit can operate at frequencies in the range 200 kHz to 2 MHz,
allowing the use of small inductors and capacitors with greater efficiency than would be
possible with other diode types. Small-area Schottky diodes are the heart of RF
detectors and mixers, which often operate at frequencies up to 50 GHz.
Limitations[edit]
The most evident limitations of Schottky diodes are their relatively low reverse voltage
ratings, and their relatively high reverse leakage current. For silicon-metal Schottky
diodes, the reverse voltage is typically 50 V or less. Some higher-voltage designs are
available (200 V is considered a high reverse voltage). Reverse leakage current, since it
increases with temperature, leads to a thermal instability issue. This often limits the
useful reverse voltage to well below the actual rating.
While higher reverse voltages are achievable, they would present a higher forward
voltage, comparable to other types of standard diodes. Such Schottky diodes would
[4]
have no advantage unless great switching speed is required.
Silicon carbide has a high thermal conductivity, and temperature has little influence on
its switching and thermal characteristics. With special packaging, silicon carbide
Schottky diodes can operate at junction temperatures of over 500 K (about 200 °C),
[5]
which allows passive radiative cooling in aerospace applications.
Applications[edit]
Voltage clamping[edit]
While standard silicon diodes have a forward voltage drop of about 0.7 V and
germanium diodes 0.3 V, Schottky diodes' voltage drop at forward biases of around 1
[6] [7]
mA is in the range of 0.15 V to 0.46 V (see the 1N5817 and 1N5711 ), which makes
them useful in voltage clamping applications and prevention of transistor saturation.
This is due to the higher current density in the Schottky diode.
The Schottky diode's low forward voltage drop is good for energy-efficient applications,
because little energy is wasted to heat. This makes them useful as blocking diodes in
stand-alone ("off-grid") photovoltaic (PV) systems which prevent batteries from
discharging through the solar panels at night. They are also used in grid-connected
systems with multiple strings connected in parallel, in order to prevent reverse current
flowing from adjacent strings through shaded strings if the bypass diodes have failed.
Schottky diodes are also used as rectifiers in switched-mode power supplies. The low
forward voltage and fast recovery time leads to increased efficiency.
They can also be used in power supply "OR"ing circuits in products that have both an
internal battery and a mains adapter input, or similar. However, the high reverse
leakage current presents a problem in this case, as any high-impedance voltage
sensing circuit (e.g., monitoring the battery voltage or detecting whether a mains
adapter is present) will see the voltage from the other power source through the diode
leakage.
Sample-and-hold circuits[edit]
Schottky diodes can be used in diode-bridge based sample and hold circuits. When
compared to regular p–n junction based diode bridges, Schottky diodes can offer
advantages. A forward-biased Schottky diode does not have any minority carrier charge
storage. This allows them to switch more quickly than regular diodes, resulting in lower
transition time from the sample to the hold step. The absence of minority carrier charge
storage also results in a lower hold step or sampling error, resulting in a more accurate
[8]
sample at the output.
Charge control[edit]
Due to its efficient electric field control, Schottky diodes can be used to accurately load
or unload single electrons in semiconductor nanostructures such as quantum wells or
[9]
quantum dots.
Designation[edit]
[10]
(surface mount version of 1N5819)
Commonly encountered Schottky diodes include the 1N58xx series rectifiers, such as
[6][11]
the 1N581x (1 A) and 1N582x (3 A) through-hole parts, and the SS1x (1 A) and
[10][12]
SS3x (3 A) surface-mount parts. Schottky rectifiers are available in numerous
[13][14]
surface-mount package styles.
Schottky metal–semiconductor junctions are featured in the successors to the 7400 TTL
family of logic devices, the 74S, 74LS and 74ALS series, where they are employed as
Baker clamps in parallel with the collector-base junctions of the bipolar transistors to
prevent their saturation, thereby greatly reducing their turn-off delays.
Alternatives[edit]
When less power dissipation is desired, a MOSFET and a control circuit can be used
instead, in an operation mode known as active rectification.
Electrowetting[edit]
Electrowetting can be observed when a Schottky diode is formed using a droplet of
liquid metal, e.g. mercury, in contact with a semiconductor, e.g. silicon. Depending on
the doping type and density in the semiconductor, the droplet spreading depends on the
[20]
magnitude and sign of the voltage applied to the mercury droplet. This effect has
[21]
been termed ‘Schottky electrowetting’.
See also[edit]
● Heterostructure barrier varactor – Semiconductor device with variable
capacitance
● List of 1N58xx Schottky diodes
● Schottky effect – a phenomenon in condensed matter physics
References[edit]
● ^ ‘’Laughton, M. A. (2003). "17. Power Semiconductor Devices". Electrical
engineer's reference book. Newnes. pp. 25–27. ISBN 978-0-7506-4637-6.
Retrieved 2011-05-16.
● ^ Hastings, Alan (2005). The Art of Analog Layout (2nd ed.). Prentice Hall. ISBN
0-13-146410-8.
● ^ Pierret, Robert F. (1996). Semiconductor Device Fundamentals. Addison-
Wesley. ISBN 978-0-131-78459-8.
● ^ "Introduction to Schottky Rectifiers" (PDF). MicroNotes. 401. Schottky rectifiers
seldom exceed 100 volts in their working peak reverse voltage since devices
moderately above this rating level will result in forward voltages equal to or
greater than equivalent pn junction rectifiers.
● ^
● Jump up to:
ab
● "Schottky Diodes: the Old Ones Are Good, the New Ones Are Better".
Electronic Design. March 1, 2011.
● ^
● Jump up to:
ab
● "1N5817 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^
● Jump up to:
ab
● "1N5711 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^ Johns, David A. and Martin, Ken. Analog Integrated Circuit Design (1997),
Wiley. Page 351. ISBN 0-471-14448-7
● ^ Couto, O. D. D.; Puebla, J.; Chekhovich, E. A.; Luxmoore, I. J.; Elliott, C. J.;
Babazadeh, N.; Skolnick, M. S.; Tartakovskii, A. I.; Krysa, A. B. (2011-09-01).
"Charge control in InP/(Ga,In)P single quantum dots embedded in Schottky
diodes". Physical Review B. 84 (12). American Physical Society (APS): 125301.
arXiv:1107.2522. Bibcode:2011PhRvB..84l5301C.
doi:10.1103/physrevb.84.125301. ISSN 1098-0121. S2CID 119215237.
● ^
● Jump up to:
ab
● "SS14 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-11-23.
● ^ "1N5820 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-11-23.
● ^ "SS34 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-11-23.
● ^ Bourns Schottky Rectifiers.
● ^ Vishay Schottky Rectifiers.
● ^ "1N6263 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^ "1SS106 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^ "1SS108 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^ "BAT4 Datasheets (PDF)". Datasheetcatalog.com. Retrieved 2013-01-14.
● ^ Vishay Small-Signal Schottky Diodes.
● ^ Arscott, Steve; Gaudet, Matthieu (2013-08-12). "Electrowetting at a liquid
metal-semiconductor junction" (PDF). Applied Physics Letters. 103 (7). AIP
Publishing: 074104. Bibcode:2013ApPhL.103g4104A. doi:10.1063/1.4818715.
ISSN 0003-6951.
● ^ Arscott, Steve (2014-07-04). "Electrowetting and semiconductors". RSC
Advances. 4 (55). Royal Society of Chemistry (RSC): 29223.
Bibcode:2014RSCAd...429223A. doi:10.1039/c4ra04187a. ISSN 2046-2069.
External links[edit]
show
● V
● T
● E
Electronic components
● Germany
Authority control ● Israel
databases: National ● United States
Category:
DiodesContents hide
(Top)
The critical parameter: Schottky barrier height
Schottky–Mott rule and Fermi level pinning
History
See also
References
Further reading
Metal–semiconductor
junction
7 languages
Article
Talk
Read
Edit
View history
Tools
Appearance hide
Text
Small
Standard
Large
Width
Standard
Wide
Color (beta)
Automatic
Light
Dark
Report an issue with dark mode
From Wikipedia, the free encyclopedia
In solid-state physics, a metal–semiconductor (M–S) junction is a type of electrical
junction in which a metal comes in close contact with a semiconductor material. It is the
oldest practical semiconductor device. M–S junctions can either be rectifying or non-
rectifying. The rectifying metal–semiconductor junction forms a Schottky barrier, making
a device known as a Schottky diode, while the non-rectifying junction is called an ohmic
[1]
contact. (In contrast, a rectifying semiconductor–semiconductor junction, the most
common semiconductor device today, is known as a p–n junction.)
Band diagram for metal-semiconductor junction at zero bias (equilibrium). Shown is the graphical
definition of the Schottky barrier height, ΦB, for an n-type semiconductor as the difference
between the interfacial conduction band edge EC and Fermi level EF.
The Schottky barrier height is defined differently for n-type and p-type semiconductors
(being measured from the conduction band edge and valence band edge, respectively).
The alignment of the semiconductor's bands near the junction is typically independent of
the semiconductor's doping level, so the n-type and p-type Schottky barrier heights are
ideally related to each other by:
ΦB(n)+ΦB(p)=Eg
In practice, the Schottky barrier height is not precisely constant across the interface,
[2]
and varies over the interfacial surface.
[3]
Band diagrams for models of formation of junction between silver and n-doped silicon. In
The Schottky–Mott rule of Schottky barrier formation, named for Walter H. Schottky
and Nevill Mott, predicts the Schottky barrier height based on the vacuum work function
of the metal relative to the vacuum electron affinity (or vacuum ionization energy) of the
semiconductor:
ΦB(n)≈Φmetal−χsemi
This model is derived based on the thought experiment of bringing together the two
materials in vacuum, and is closely related in logic to Anderson's rule for
semiconductor-semiconductor junctions. Different semiconductors respect the
[5]
Schottky–Mott rule to varying degrees.
Although the Schottky–Mott model correctly predicted the existence of band bending in
the semiconductor, it was found experimentally that it would give grossly incorrect
predictions for the height of the Schottky barrier. A phenomenon referred to as "Fermi
level pinning" caused some point of the band gap, at which finite DOS exists, to be
locked (pinned) to the Fermi level. This made the Schottky barrier height almost
[5]
completely insensitive to the metal's work function:
ΦB≈12Ebandgap
In fact, empirically, it is found that neither of the above extremes is quite correct. The
choice of metal does have some effect, and there appears to be a weak correlation
between the metal work function and the barrier height, however the influence of the
[6]: 143
work function is only a fraction of that predicted by the Schottky-Mott rule.
It was noted in 1947 by John Bardeen that the Fermi level pinning phenomenon would
naturally arise if there were chargeable states in the semiconductor right at the
interface, with energies inside the semiconductor's gap. These would either be induced
during the direct chemical bonding of the metal and semiconductor (metal-induced gap
states) or be already present in the semiconductor–vacuum surface (surface states).
These highly dense surface states would be able to absorb a large quantity of charge
donated from the metal, effectively shielding the semiconductor from the details of the
metal. As a result, the semiconductor's bands would necessarily align to a location
relative to the surface states which are in turn pinned to the Fermi level (due to their
[3]
high density), all without influence from the metal.
The Fermi level pinning effect is strong in many commercially important semiconductors
[5]
(Si, Ge, GaAs), and thus can be problematic for the design of semiconductor devices.
For example, nearly all metals form a significant Schottky barrier to n-type germanium
and an ohmic contact to p-type germanium, since the valence band edge is strongly
[7]
pinned to the metal's Fermi level. The solution to this inflexibility requires additional
processing steps such as adding an intermediate insulating layer to unpin the bands. (In
[8]
the case of germanium, germanium nitride has been used )
History[edit]
The rectification property of metal–semiconductor contacts was discovered by
Ferdinand Braun in 1874 using mercury metal contacted with copper sulfide and iron
[9]
sulfide semiconductors. Sir Jagadish Chandra Bose applied for a US patent for a
metal-semiconductor diode in 1901. This patent was awarded in 1904.
JC Bose patent
The first theory that predicted the correct direction of rectification of the metal–
semiconductor junction was given by Nevill Mott in 1939. He found the solution for both
the diffusion and drift currents of the majority carriers through the semiconductor
surface space charge layer which has been known since about 1948 as the Mott barrier.
Walter H. Schottky and Spenke extended Mott's theory by including a donor ion whose
density is spatially constant through the semiconductor surface layer. This changed the
constant electric field assumed by Mott to a linearly decaying electric field. This
semiconductor space-charge layer under the metal is known as the Schottky barrier. A
similar theory was also proposed by Davydov in 1939. Although it gives the correct
direction of rectification, it has also been proven that the Mott theory and its Schottky-
Davydov extension gives the wrong current limiting mechanism and wrong current-
voltage formulae in silicon metal/semiconductor diode rectifiers. The correct theory was
developed by Hans Bethe and reported by him in a M.I.T. Radiation Laboratory Report
dated November 23, 1942. In Bethe's theory, the current is limited by thermionic
emission of electrons over the metal–semiconductor potential barrier. Thus, the
appropriate name for the metal–semiconductor diode should be the Bethe diode,
instead of the Schottky diode, since the Schottky theory does not predict the modern
[13]
metal–semiconductor diode characteristics correctly.
See also[edit]
● Schottky barrier
References[edit]
● ^ Semiconductor Devices: Modelling and Technology, Nandita Dasgupta,
Amitava Dasgupta.(2004) ISBN 81-203-2398-X.
● ^ "Inhomogeneous Schottky Barrier".
● ^
● Jump up to:
ab
● Bardeen, J. (1947). "Surface States and Rectification at a Metal Semi-
Conductor Contact". Physical Review. 71 (10): 717–727.
Bibcode:1947PhRv...71..717B. doi:10.1103/PhysRev.71.717.
● ^ Tung, R. (2001). "Formation of an electric dipole at metal-semiconductor
interfaces". Physical Review B. 64 (20): 205310. Bibcode:2001PhRvB..64t5310T.
doi:10.1103/PhysRevB.64.205310.
● ^
● Jump up to:
abc
● "Barrier Height Correlations and Systematics".
● ^ Sze, S. M. Ng, Kwok K. (2007). Physics of semiconductor devices. John Wiley
& Sons. ISBN 978-0-471-14323-9. OCLC 488586029.
● ^ Nishimura, T.; Kita, K.; Toriumi, A. (2007). "Evidence for strong Fermi-level
pinning due to metal-induced gap states at metal/germanium interface". Applied
Physics Letters. 91 (12): 123123. Bibcode:2007ApPhL..91l3123N.
doi:10.1063/1.2789701.
● ^ Lieten, R. R.; Degroote, S.; Kuijk, M.; Borghs, G. (2008). "Ohmic contact
formation on n-type Ge". Applied Physics Letters. 92 (2): 022106.
Bibcode:2008ApPhL..92b2106L. doi:10.1063/1.2831918.
● ^ Braun, F. (1874), "Ueber die Stromleitung durch Schwefelmetalle" [On current
conduction through metal sulfides], Annalen der Physik und Chemie (in German),
153 (4): 556–563, Bibcode:1875AnP...229..556B,
doi:10.1002/andp.18752291207
● ^ Pierce, G. W. (1907). "Crystal Rectifiers for Electric Currents and Electric
Oscillations. Part I. Carborundum". Physical Review. Series I. 25 (1): 31–60.
Bibcode:1907PhRvI..25...31P. doi:10.1103/PhysRevSeriesI.25.31.
● ^ US 1745175 "Method and apparatus for controlling electric current" first filed in
Canada on 22.10.1925.
● ^ US 755840, Bose, Jagadis Chunder, "Detector for electrical disturbances",
published September 30, 1901, issued March 29, 1904
● ^ Sah, Chih-Tang (1991). Fundamentals of Solid-State Electronics. World
Scientific. ISBN 9810206372.
● ^ S. Arscott and M. Gaudet "Electrowetting at a liquid metal-semiconductor
junction" Appl. Phys. Lett. 103, 074104 (2013). doi:10.1063/1.4818715
● ^ S. Arscott "Electrowetting and semiconductors" RSC Advances 4, 29223
(2014). doi:10.1039/C4RA04187A
● ^ Bassett, Ross Knox (2007). To the Digital Age: Research Labs, Start-up
Companies, and the Rise of MOS Technology. Johns Hopkins University Press.
p. 328. ISBN 9780801886393.
● ^ The Industrial Reorganization Act: The communications industry. U.S.
Government Printing Office. 1973. p. 1475.
● ^ Atalla, M.; Kahng, D. (November 1962). "A new "Hot electron" triode structure
with semiconductor-metal emitter". IRE Transactions on Electron Devices. 9 (6):
507–508. Bibcode:1962ITED....9..507A. doi:10.1109/T-ED.1962.15048. ISSN
0096-2430. S2CID 51637380.
● ^ Kasper, E. (2018). Silicon-Molecular Beam Epitaxy. CRC Press. ISBN
9781351093514.
● ^
● Jump up to:
ab
● Siegel, Peter H.; Kerr, Anthony R.; Hwang, Wei (March 1984). NASA
Technical Paper 2287: Topics in the Optimization of Millimeter-Wave Mixers
(PDF). NASA. pp. 12–13.
● ^
● Jump up to:
ab
● Button, Kenneth J. (1982). Infrared and Millimeter Waves V6: Systems and
Components. Elsevier. p. 214. ISBN 9780323150590.
● ^ Anand, Y. (2013). "Microwave Schottky Barrier Diodes". Metal-Semiconductor
Schottky Barrier Junctions and Their Applications. Springer Science & Business
Media. p. 220. ISBN 9781468446555.
● ^ Archer, R. J.; Atalla, M. M. (January 1963). "Metals Contacts on Cleaved
Silicon Surfaces". Annals of the New York Academy of Sciences. 101 (3): 697–
708. Bibcode:1963NYASA.101..697A. doi:10.1111/j.1749-6632.1963.tb54926.x.
ISSN 1749-6632. S2CID 84306885.
Further reading[edit]
● Streetman, Ben G.; Banerjee, Sanjay Kumar (2016). Solid state electronic
devices. Boston: Pearson. p. 251-257. ISBN 978-1-292-06055-2. OCLC
908999844.
hide
Int
● FAST
Na
● France
● BnF data
● Germany
● Israel
● United States
Category:
Semiconductor structures
This page was last edited on 12 February 2024, at 23:21 (UTC).
Text is available under the Creative Commons Attribution-ShareAlike License 4.0; additional terms may apply.
By using this site, you agree to the Terms of U