You are on page 1of 16

Distributed-element circuit

Distributed-element circuits are electrical circuits


composed of lengths of transmission lines or other
distributed components. These circuits perform the
same functions as conventional circuits composed of
passive components, such as capacitors, inductors,
and transformers. They are used mostly at
microwave frequencies, where conventional
components are difficult (or impossible) to
implement.

Conventional circuits consist of individual


components manufactured separately then connected A low-noise block converter with distributed elements.
together with a conducting medium. Distributed- The circuitry on the right is lumped elements. The
element circuits are built by forming the medium distributed-element circuitry is centre and left of centre,
itself into specific patterns. A major advantage of and is constructed in microstrip.
distributed-element circuits is that they can be
produced cheaply as a printed circuit board for
consumer products, such as satellite television. They are also made in coaxial and waveguide formats for
applications such as radar, satellite communication, and microwave links.

A phenomenon commonly used in distributed-element circuits is that a length of transmission line can be made
to behave as a resonator. Distributed-element components which do this include stubs, coupled lines, and
cascaded lines. Circuits built from these components include filters, power dividers, directional couplers, and
circulators.

Distributed-element circuits were studied during the 1920s and 1930s but did not become important until
World War II, when they were used in radar. After the war their use was limited to military, space, and
broadcasting infrastructure, but improvements in materials science in the field soon led to broader applications.
They can now be found in domestic products such as satellite dishes and mobile phones.

Contents
Circuit modelling
Construction with
transmission lines
Advantages and
disadvantages
Media A low-pass filter as conventional discrete components connected on a
Paired conductors printed circuit board (left), and as a distributed-element design printed on
the board itself (right)
Coaxial
Planar
Waveguide
Mechanical
Circuit components
Stub
Coupled lines
Cascaded lines
Cavity resonator
Dielectric resonator
Helical resonator
Fractals
Taper
Distributed resistance
Circuit blocks
Filters and impedance
matching
Power dividers,
combiners and
directional couplers
Hybrids
Circulators
Active components
History
References
Bibliography

Circuit modelling
Distributed-element circuits are designed with the distributed-element model, an alternative to the lumped-
element model in which the passive electrical elements of electrical resistance, capacitance and inductance are
assumed to be "lumped" at one point in space in a resistor, capacitor or inductor, respectively. The distributed-
element model is used when this assumption no longer holds, and these properties are considered to be
distributed in space. The assumption breaks down when there is significant time for electromagnetic waves to
travel from one terminal of a component to the other; "significant", in this context, implies enough time for a
noticeable phase change. The amount of phase change is dependent on the wave's frequency (and inversely
dependent on wavelength). A common rule of thumb amongst engineers is to change from the lumped to the
distributed model when distances involved are more than one-tenth of a wavelength (a 36° phase change). The
lumped model completely fails at one-quarter wavelength (a 90° phase change), with not only the value, but
the nature of the component not being as predicted. Due to this dependence on wavelength, the distributed-
element model is used mostly at higher frequencies; at low frequencies, distributed-element components are
too bulky. Distributed designs are feasible above 300 MHz, and are the technology of choice at microwave
frequencies above 1 GHz.[1]

There is no clear-cut demarcation in the frequency at which these models should be used. Although the
changeover is usually somewhere in the 100-to-500 MHz range, the technological scale is also significant;
miniaturised circuits can use the lumped model at a higher frequency. Printed circuit boards (PCBs) using
through-hole technology are larger than equivalent designs using surface-mount technology. Hybrid integrated
circuits are smaller than PCB technologies, and monolithic integrated circuits are smaller than both. Integrated
circuits can use lumped designs at higher frequencies than printed circuits, and this is done in some radio
frequency integrated circuits. This choice is particularly significant for hand-held devices, because lumped-
element designs generally result in a smaller product.[2]
Construction with transmission lines

The overwhelming majority of distributed-


element circuits are composed of lengths of
transmission line, a particularly simple form to
model. The cross-sectional dimensions of the
line are unvarying along its length, and are
small compared to the signal wavelength;
thus, only distribution along the length of the
line need be considered. Such an element of a
distributed circuit is entirely characterised by
its length and characteristic impedance. A Frequency response of a fifth-order Chebyshev filter
further simplification occurs in commensurate constructed from lumped (top) and distributed components
line circuits, where all the elements are the (bottom)
same length. With commensurate circuits, a
lumped circuit design prototype consisting of
capacitors and inductors can be directly converted into a distributed circuit with a one-to-one correspondence
between the elements of each circuit.[3]

Commensurate line circuits are important because a design theory for producing them exists; no general theory
exists for circuits consisting of arbitrary lengths of transmission line (or any arbitrary shapes). Although an
arbitrary shape can be analysed with Maxwell's equations to determine its behaviour, finding useful structures
is a matter of trial and error or guesswork.[4]

An important difference between distributed-element circuits and lumped-element circuits is that the frequency
response of a distributed circuit periodically repeats as shown in the Chebyshev filter example; the equivalent
lumped circuit does not. This is a result of the transfer function of lumped forms being a rational function of
complex frequency; distributed forms are an irrational function. Another difference is that cascade-connected
lengths of line introduce a fixed delay at all frequencies (assuming an ideal line). There is no equivalent in
lumped circuits for a fixed delay, although an approximation could be constructed for a limited frequency
range.[5]

Advantages and disadvantages


Distributed-element circuits are cheap and easy to manufacture in some formats, but take up more space than
lumped-element circuits. This is problematic in mobile devices (especially hand-held ones), where space is at a
premium. If the operating frequencies are not too high, the designer may miniaturise components rather than
switching to distributed elements. However, parasitic elements and resistive losses in lumped components are
greater with increasing frequency as a proportion of the nominal value of the lumped-element impedance. In
some cases, designers may choose a distributed-element design (even if lumped components are available at
that frequency) to benefit from improved quality. Distributed-element designs tend to have greater power-
handling capability; with a lumped component, all the energy passed by a circuit is concentrated in a small
volume.[6]

Media

Paired conductors
Several types of transmission line exist, and any of them can be used to construct distributed-element circuits.
The oldest (and still most widely used) is a pair of conductors; its most common form is twisted pair, used for
telephone lines and Internet connections. It is not often used for distributed-element circuits because the
frequencies used are lower than the point where distributed-element designs become advantageous. However,
designers frequently begin with a lumped-element design and convert it to an open-wire distributed-element
design. Open wire is a pair of parallel uninsulated conductors used, for instance, for telephone lines on
telegraph poles. The designer does not usually intend to implement the circuit in this form; it is an intermediate
step in the design process. Distributed-element designs with conductor pairs are limited to a few specialised
uses, such as Lecher lines and the twin-lead used for antenna feed lines.[7]

Coaxial

Coaxial line, a centre conductor surrounded by an insulated shielding


conductor, is widely used for interconnecting units of microwave
equipment and for longer-distance transmissions. Although coaxial
distributed-element devices were commonly manufactured during the
second half of the 20th century, they have been replaced in many
applications by planar forms due to cost and size considerations. Air-
dielectric coaxial line is used for low-loss and high-power
applications. Distributed-element circuits in other media still A collection of coaxial directional
commonly transition to coaxial connectors at the circuit ports for couplers. One has the cover
interconnection purposes.[8] removed, showing its internal
structure.

Planar

The majority of modern distributed-element circuits use planar transmission lines, especially those in mass-
produced consumer items. There are several forms of planar line, but the kind known as microstrip is the most
common. It can be manufactured by the same process as printed circuit boards and hence is cheap to make. It
also lends itself to integration with lumped circuits on the same board. Other forms of printed planar lines
include stripline, finline and many variations. Planar lines can also be used in monolithic microwave integrated
circuits, where they are integral to the device chip.[9]

Waveguide

Many distributed-element designs can be directly implemented in


waveguide. However, there is an additional complication with
waveguides in that multiple modes are possible. These sometimes
exist simultaneously, and this situation has no analogy in conducting
lines. Waveguides have the advantages of lower loss and higher
quality resonators over conducting lines, but their relative expense and
bulk means that microstrip is often preferred. Waveguide mostly finds
uses in high-end products, such as high-power military radars and the
upper microwave bands (where planar formats are too lossy).
A waveguide filter
Waveguide becomes bulkier with lower frequency, which militates
against its use on the lower bands.[10]

Mechanical
In a few specialist applications, such as the mechanical filters in high-end radio transmitters (marine, military,
amateur radio), electronic circuits can be implemented as mechanical components; this is done largely because
of the high quality of the mechanical resonators. They are used in the radio frequency band (below microwave
frequencies), where waveguides might otherwise be used. Mechanical circuits can also be implemented, in
whole or in part, as distributed-element circuits. The frequency at which the transition to distributed-element
design becomes feasible (or necessary) is much lower with mechanical circuits. This is because the speed at
which signals travel through mechanical media is much lower than the speed of electrical signals.[11]

Circuit components
There are several structures that are repeatedly used in distributed-element circuits. Some of the common ones
are described below.

Stub

A stub is a short length of line that branches to the side of a main line. The end of the stub is often left open- or
short-circuited, but may also be terminated with a lumped component. A stub can be used on its own (for
instance, for impedance matching), or several of them can be used together in a more complex circuit such as a
filter. A stub can be designed as the equivalent of a lumped capacitor, inductor, or resonator.[12]

Departures from constructing with uniform transmission lines in


distributed-element circuits are rare. One such departure that is widely
used is the radial stub, which is shaped like a sector of a circle. They
are often used in pairs, one on either side of the main transmission
line. Such pairs are called butterfly or bowtie stubs.[13]

Butterfly stub filter


Coupled lines

Coupled lines are two transmission lines between which there is some electromagnetic coupling. The coupling
can be direct or indirect. In indirect coupling, the two lines are run closely together for a distance with no
screening between them. The strength of the coupling depends on the distance between the lines and the cross-
section presented to the other line. In direct coupling, branch lines directly connect the two main lines together
at intervals.[14]

Coupled lines are a common method of constructing power dividers and directional couplers. Another
property of coupled lines is that they act as a pair of coupled resonators. This property is used in many
distributed-element filters.[15]

Cascaded lines

Cascaded lines are lengths of transmission line where the output of one line is connected to the input of the
next. Multiple cascaded lines of different characteristic impedances can be used to construct a filter or a wide-
band impedance matching network. This is called a stepped impedance structure.[16] A single, cascaded line
one-quarter wavelength long forms a quarter-wave impedance transformer. This has the useful property of
transforming any impedance network into its dual; in this role, it is called an impedance inverter. This structure
can be used in filters to implement a lumped-element prototype in ladder topology as a distributed-element
circuit. The quarter-wave transformers are alternated with a distributed-element resonator to achieve this.
However, this is now a dated design; more compact inverters, such as
the impedance step, are used instead. An impedance step is the
discontinuity formed at the junction of two cascaded transmission
lines with different characteristic impedances.[17]

Cavity resonator

A cavity resonator is an empty (or sometimes dielectric-filled) space


surrounded by conducting walls. Apertures in the walls couple the
An orthomode transducer (a variety
resonator to the rest of the circuit. Resonance occurs due to
of duplexer) with stepped impedance
electromagnetic waves reflected back and forth from the cavity walls
matching
setting up standing waves. Cavity resonators can be used in many
media, but are most naturally formed in waveguide from the already
existing metal walls of the guide.[18]

Dielectric resonator

A dielectric resonator is a piece of dielectric material exposed to electromagnetic waves. It is most often in the
form of a cylinder or thick disc. Although cavity resonators can be filled with dielectric, the essential difference
is that in cavity resonators the electromagnetic field is entirely contained within the cavity walls. A dielectric
resonator has some field in the surrounding space. This can lead to undesirable coupling with other
components. The major advantage of dielectric resonators is that they are considerably smaller than the
equivalent air-filled cavity.[19]

Helical resonator

A helical resonator is a helix of wire in a cavity; one end is unconnected, and the other is bonded to the cavity
wall. Although they are superficially similar to lumped inductors, helical resonators are distributed-element
components and are used in the VHF and lower UHF bands.[20]

Fractals

The use of fractal-like curves as a circuit component is an emerging field in


distributed-element circuits.[22] Fractals have been used to make resonators
for filters and antennae. One of the benefits of using fractals is their space-
filling property, making them smaller than other designs.[23] Other advantages
include the ability to produce wide-band and multi-band designs, good in-
band performance, and good out-of-band rejection.[24] In practice, a true
fractal cannot be made because at each fractal iteration the manufacturing
tolerances become tighter and are eventually greater than the construction
method can achieve. However, after a small number of iterations, the
performance is close to that of a true fractal. These may be called pre-fractals Three-iteration Hilbert fractal
or finite-order fractals where it is necessary to distinguish from a true resonator in microstrip[21]
fractal.[25]

Fractals that have been used as a circuit component include the Koch snowflake, Minkowski island, Sierpiński
curve, Hilbert curve, and Peano curve.[26] The first three are closed curves, suitable for patch antennae. The
latter two are open curves with terminations on opposite sides of the fractal. This makes them suitable for use
where a connection in cascade is required.[27]
Taper

A taper is a transmission line with a gradual change in cross-section. It can be considered the limiting case of
the stepped impedance structure with an infinite number of steps.[28] Tapers are a simple way of joining two
transmission lines of different characteristic impedances. Using tapers greatly reduces the mismatch effects that
a direct join would cause. If the change in cross-section is not too great, no other matching circuitry may be
needed.[29] Tapers can provide transitions between lines in different media, especially different forms of planar
media.[30] Tapers commonly change shape linearly, but a variety of other profiles may be used. The profile
that achieves a specified match in the shortest length is known as a Klopfenstein taper and is based on the
Chebychev filter design.[31]

Tapers can be used to match a transmission line to an antenna. In some designs, such as the horn antenna and
Vivaldi antenna, the taper is itself the antenna. Horn antennae, like other tapers, are often linear, but the best
match is obtained with an exponential curve. The Vivaldi antenna is a flat (slot) version of the exponential
taper.[32]

Distributed resistance

Resistive elements are generally not useful in a distributed-element circuit. However, distributed resistors may
be used in attenuators and line terminations. In planar media they can be implemented as a meandering line of
high-resistance material, or as a deposited patch of thin-film or thick-film material.[33] In waveguide, a card of
microwave absorbent material can be inserted into the waveguide.[34]

Circuit blocks

Filters and impedance matching

Filters are a large percentage of circuits constructed with


distributed elements. A wide range of structures are used
for constructing them, including stubs, coupled lines and
cascaded lines. Variations include interdigital filters,
combline filters and hairpin filters. More-recent
developments include fractal filters.[35] Many filters are
Microstrip band-pass hairpin filter (left), followed
constructed in conjunction with dielectric resonators.[36]
by a low-pass stub filter
As with lumped-element filters, the more elements used,
the closer the filter comes to an ideal response; the
structure can become quite complex.[37] For simple, narrow-band requirements, a single resonator may suffice
(such as a stub or spurline filter).[38]

Impedance matching for narrow-band applications is frequently achieved with a single matching stub.
However, for wide-band applications the impedance-matching network assumes a filter-like design. The
designer prescribes a required frequency response, and designs a filter with that response. The only difference
from a standard filter design is that the filter's source and load impedances differ.[39]

Power dividers, combiners and directional couplers


A directional coupler is a four-port device which couples power flowing in
one direction from one path to another. Two of the ports are the input and
output ports of the main line. A portion of the power entering the input port is
coupled to a third port, known as the coupled port. None of the power
entering the input port is coupled to the fourth port, usually known as the
isolated port. For power flowing in the reverse direction and entering the
output port, a reciprocal situation occurs; some power is coupled to the
isolated port, but none is coupled to the coupled port.[41]
Microstrip sawtooth
directional coupler, a variant
A power divider is often constructed as a directional coupler, with the isolated
of the coupled-lines
port permanently terminated in a matched load (making it effectively a three-
directional coupler[40]
port device). There is no essential difference between the two devices. The
term directional coupler is usually used when the coupling factor (the
proportion of power reaching the coupled port) is low, and power divider
when the coupling factor is high. A power combiner is simply a power splitter used in reverse. In distributed-
element implementations using coupled lines, indirectly coupled lines are more suitable for low-coupling
directional couplers; directly-coupled branch line couplers are more suitable for high-coupling power
dividers.[42]

Distributed-element designs rely on an element length of one-quarter wavelength (or some other length); this
will hold true at only one frequency. Simple designs, therefore, have a limited bandwidth over which they will
work successfully. Like impedance matching networks, a wide-band design requires multiple sections and the
design begins to resemble a filter.[43]

Hybrids

A directional coupler which splits power equally between the output and
coupled ports (a 3 dB coupler) is called a hybrid.[44] Although "hybrid"
originally referred to a hybrid transformer (a lumped device used in
telephones), it now has a broader meaning. A widely used distributed-element
hybrid which does not use coupled lines is the hybrid ring or rat-race coupler.
Each of its four ports is connected to a ring of transmission line at a different
point. Waves travel in opposite directions around the ring, setting up standing
waves. At some points on the ring, destructive interference results in a null; no
Hybrid ring, used to produce
power will leave a port set at that point. At other points, constructive
sum and difference signals
interference maximises the power transferred.[45]

Another use for a hybrid coupler is to produce the sum and difference of two
signals. In the illustration, two input signals are fed into the ports marked 1 and 2. The sum of the two signals
appears at the port marked Σ, and the difference at the port marked Δ.[46] In addition to their uses as couplers
and power dividers, directional couplers can be used in balanced mixers, frequency discriminators, attenuators,
phase shifters, and antenna array feed networks.[47]

Circulators

A circulator is usually a three- or four-port device in which power entering one port is transferred to the next
port in rotation, as if round a circle. Power can flow in only one direction around the circle (clockwise or
anticlockwise), and no power is transferred to any of the other ports. Most distributed-element circulators are
based on ferrite materials.[48] Uses of circulators include as an isolator to protect a transmitter (or other
equipment) from damage due to reflections from the antenna, and as a duplexer connecting the antenna,
transmitter and receiver of a radio system.[49]
An unusual application of a circulator is in a reflection amplifier, where the
negative resistance of a Gunn diode is used to reflect back more power than it
received. The circulator is used to direct the input and output power flows to
separate ports.[50]

Passive circuits, both lumped and distributed, are nearly always reciprocal;
however, circulators are an exception. There are several equivalent ways to
define or represent reciprocity. A convenient one for circuits at microwave
frequencies (where distributed-element circuits are used) is in terms of their S-
parameters. A reciprocal circuit will have an S-parameter matrix, [S], which is
A coaxial ferrite circulator
symmetric. From the definition of a circulator, it is clear that this will not be
operating at 1 GHz
the case,

for an ideal three-port circulator, showing that circulators are non-reciprocal by definition. It follows that it is
impossible to build a circulator from standard passive components (lumped or distributed). The presence of a
ferrite, or some other non-reciprocal material or system, is essential for the device to work.[51]

Active components
Distributed elements are usually passive, but most applications will
require active components in some role. A microwave hybrid
integrated circuit uses distributed elements for many passive
components, but active components (such as diodes, transistors, and
some passive components) are discrete. The active components may
be packaged, or they may be placed on the substrate in chip form Microstrip circuit with discrete
without individual packaging to reduce size and eliminate packaging- transistors in miniature surface-
induced parasitics.[52] mount packages, capacitors and
resistors in chip form, and biasing
Distributed amplifiers consist of a number of amplifying devices filters as distributed elements
(usually FETs), with all their inputs connected via one transmission
line and all their outputs via another transmission line. The lengths of
the two lines must be equal between each transistor for the circuit to work correctly, and each transistor adds to
the output of the amplifier. This is different from a conventional multistage amplifier, where the gain is
multiplied by the gain of each stage. Although a distributed amplifier has lower gain than a conventional
amplifier with the same number of transistors, it has significantly greater bandwidth. In a conventional
amplifier, the bandwidth is reduced by each additional stage; in a distributed amplifier, the overall bandwidth is
the same as the bandwidth of a single stage. Distributed amplifiers are used when a single large transistor (or a
complex, multi-transistor amplifier) would be too large to treat as a lumped component; the linking
transmission lines separate the individual transistors.[53]

History
Distributed-element modelling was first used in electrical network analysis by Oliver Heaviside[54] in 1881.
Heaviside used it to find a correct description of the behaviour of signals on the transatlantic telegraph cable.
Transmission of early transatlantic telegraph had been difficult and slow due to dispersion, an effect which was
not well understood at the time. Heaviside's analysis, now known as the
telegrapher's equations, identified the problem and suggested[55] methods for
overcoming it. It remains the standard analysis of transmission lines.[56]

Warren P. Mason was the first to investigate the possibility of distributed-


element circuits, and filed a patent[57] in 1927 for a coaxial filter designed by
this method. Mason and Sykes published the definitive paper on the method
in 1937. Mason was also the first to suggest a distributed-element acoustic
filter in his 1927 doctoral thesis, and a distributed-element mechanical filter in
a patent[58] filed in 1941. Mason's work was concerned with the coaxial form
and other conducting wires, although much of it could also be adapted for
waveguide. The acoustic work had come first, and Mason's colleagues in the Oliver Heaviside
Bell Labs radio department asked him to assist with coaxial and waveguide
filters.[59]

Before World War II, there was little demand for distributed-element circuits; the frequencies used for radio
transmissions were lower than the point at which distributed elements became advantageous. Lower
frequencies had a greater range, a primary consideration for broadcast purposes. These frequencies require
long antennae for efficient operation, and this led to work on higher-frequency systems. A key breakthrough
was the 1940 introduction of the cavity magnetron which operated in the microwave band and resulted in
radar equipment small enough to install in aircraft.[60] A surge in distributed-element filter development
followed, filters being an essential component of radars. The signal loss in coaxial components led to the first
widespread use of waveguide, extending the filter technology from the coaxial domain into the waveguide
domain.[61]

The wartime work was mostly unpublished until after the war for security reasons, which made it difficult to
ascertain who was responsible for each development. An important centre for this research was the MIT
Radiation Laboratory (Rad Lab), but work was also done elsewhere in the US and Britain. The Rad Lab work
was published[62] by Fano and Lawson.[63] Another wartime development was the hybrid ring. This work
was carried out at Bell Labs, and was published[64] after the war by W. A. Tyrrell. Tyrrell describes hybrid
rings implemented in waveguide, and analyses them in terms of the well-known waveguide magic tee. Other
researchers[65] soon published coaxial versions of this device.[66]

George Matthaei led a research group at Stanford Research Institute which included Leo Young and was
responsible for many filter designs. Matthaei first described the interdigital filter[67] and the combline filter.[68]
The group's work was published[69] in a landmark 1964 book covering the state of distributed-element circuit
design at that time, which remained a major reference work for many years.[70]

Planar formats began to be used with the invention of stripline by Robert M. Barrett. Although stripline was
another wartime invention, its details were not published[71] until 1951. Microstrip, invented in 1952,[72]
became a commercial rival of stripline; however, planar formats did not start to become widely used in
microwave applications until better dielectric materials became available for the substrates in the 1960s.[73]
Another structure which had to wait for better materials was the dielectric resonator. Its advantages (compact
size and high quality) were first pointed out[74] by R. D. Richtmeyer in 1939, but materials with good
temperature stability were not developed until the 1970s. Dielectric resonator filters are now common in
waveguide and transmission line filters.[75]

Important theoretical developments included Paul I. Richards' commensurate line theory, which was
published[76] in 1948, and Kuroda's identities, a set of transforms which overcame some practical limitations
of Richards theory, published[77] by Kuroda in 1955.[78] According to Nathan Cohen, the log-periodic
antenna, invented by Raymond DuHamel and Dwight Isbell in 1957, should be considered the first fractal
antenna. However, its self-similar nature, and hence its relation to fractals was missed at the time. It is still not
usually classed as a fractal antenna. Cohen was the first to explicitly identify the class of fractal antennae after
being inspired by a lecture of Benoit Mandelbrot in 1987, but he could not get a paper published until
1995.[79]

References
1. Vendelin et al., pp. 35–37
2. Nguyen, p. 28
Vendelin et al., pp. 35–36
3. Hunter, pp. 137–138
4. Hunter, p. 137
5. Hunter, pp. 139–140
6. Doumanis et al., pp. 45–46
Nguyen, pp. 27–28
7. Hura & Singhal, pp. 178–179
Magnusson et al., p. 240
Gupta, p. 5.5
Craig, pp. 291–292
Henderson & Camargo, pp. 24–25
Chen et al., p. 73
8. Natarajan, pp. 11–12
9. Ghione & Pirola, pp. 18–19
10. Ghione & Pirola, p. 18
11. Taylor & Huang pp. 353–358
Johnson (1983), p. 102
Mason (1961)
Johnson et al. (1971), pp. 155, 169
12. Edwards & Steer, pp. 78, 345–347
Banerjee, p. 74
13. Edwards & Steer, pp. 347–348
14. Magnusson et al., p. 199
Garg et al., p. 433
Chang & Hsieh, pp. 227–229
Bhat & Koul, pp. 602–609
15. Bhat & Koul, pp. 10, 602, 622
16. Lee, p. 787
17. Helszajn, p. 189
18. Hunter, pp. 209–210
19. Penn & Alford, pp. 524–530
20. Whitaker, p. 227
Doumanis et al., pp. 12–14
21. Janković et al., p. 197
22. Ramadan et al., p. 237
23. Janković et al., p. 191
24. Janković et al., pp. 191–192
25. Janković et al., p. 196
26. Janković et al., p. 196
27. Janković et al., p. 196
28. Zhurbenko, p. 310
29. Garg et al., pp. 180–181
30. Garg et al., pp. 404–406, 540
Edwards & Steer, p. 493
31. Zhurbenko, p. 311
Misra, p. 276
Lee, p. 100
32. Bakshi & Bakshi
pp. 3-68–3-70
Milligan, p. 513
33. Maloratsky (2012), p. 69
Hilty, p. 425
Bahl (2014), p. 214
34. Hilty, pp. 426–427
35. Cohen, p. 220
36. Hong & Lancaster, pp. 109, 235
Makimoto & Yamashita, p. 2
37. Harrell, p. 150
38. Awang, p. 296
39. Bahl (2009), p. 149
40. Maloratsky (2004), p. 160
41. Sisodia & Raghuvansh, p. 70
42. Ishii, p. 226
43. Bhat & Khoul, pp. 622–627
44. Maloratsky (2004), p. 117
45. Chang & Hsieh, pp. 197–198
46. Ghione & Pirola, pp. 172–173
47. Chang & Hsieh, p. 227
Maloratsky (2004), p. 117
48. Sharma, pp. 175–176
Linkhart, p. 29
49. Meikle, p. 91
Lacomme et al., pp. 6–7
50. Roer, pp. 255–256
51. Maloratsky (2004), pp. 285–286
52. Bhat & Khoul, pp. 9–10, 15
53. Kumar & Grebennikov, pp. 153–154
54. Heaviside (1925)
55. Heaviside (1887), p. 81
56. Brittain, p. 39
57. Mason (1930)
58. Mason (1961)
59. Johnson et al. (1971), p. 155
Fagen & Millman, p. 108
Levy & Cohn, p. 1055
Polkinghorn (1973)
60. Borden, p. 3
61. Levy & Cohn, p. 1055
62. Fano & Lawson (1948)
63. Levy & Cohn, p. 1055
64. Tyrrell (1947)
65. Sheingold & Morita (1953)
Albanese & Peyser (1958)
66. Ahn, p. 3
67. Matthaei (1962)
68. Matthaei (1963)
69. Matthaei et al. (1964)
70. Levy and Cohn, pp. 1057–1059
71. Barrett & Barnes (1951)
72. Grieg and Englemann (1952)
73. Bhat & Koul, p. 3
74. Richtmeyer (1939)
75. Makimoto & Yamashita, pp. 1–2
76. Richards (1948)
77. First English publication:
Ozaki & Ishii (1958)
78. Levy & Cohn, pp. 1056–1057
79. Cohen, pp. 210–211

Bibliography
Ahn, Hee-Ran, Asymmetric Passive Components in Microwave Integrated Circuits, John Wiley
& Sons, 2006 ISBN 0470036958.
Albanese, V J; Peyser, W P, "An analysis of a broad-band coaxial hybrid ring" (https://ieeexplor
e.ieee.org/document/1125207/), IRE Transactions on Microwave Theory and Techniques, vol.
6, iss. 4, pp. 369–373, October 1958.
Awang, Zaiki, Microwave Systems Design, Springer Science & Business Media, 2013
ISBN 981445124X.
Bahl, Inder J, Fundamentals of RF and Microwave Transistor Amplifiers, John Wiley & Sons,
2009 ISBN 0470462310.
Bahl, Inder J, Control Components Using Si, GaAs, and GaN Technologies, Artech House,
2014 ISBN 1608077128.
Bakshi, U A; Bakshi, A V, Antenna And Wave Propagation, Technical Publications, 2009
ISBN 8184317220.
Banerjee, Amal, Automated Electronic Filter Design, Springer, 2016 ISBN 3319434705.
Barrett, R M, "Etched sheets serve as microwave components", Electronics, vol. 25, pp. 114–
118, June 1952.
Barrett, R M; Barnes, M H, "Microwave printed circuits", Radio TV News, vol. 46, 16 September
1951.
Bhat, Bharathi; Koul, Shiban K, Stripline-like Transmission Lines for Microwave Integrated
Circuits, New Age International, 1989 ISBN 8122400523.
Borden, Brett, Radar Imaging of Airborne Targets, CRC Press, 1999 ISBN 1420069004.
Brittain, James E, "The introduction of the loading coil: George A. Campbell and Michael I.
Pupin" (https://www.jstor.org/stable/3102809), Technology and Culture, vol. 11, no. 1, pp. 36–
57, January 1970.
Chang, Kai; Hsieh, Lung-Hwa, Microwave Ring Circuits and Related Structures, John Wiley &
Sons, 2004 ISBN 047144474X.
Chen, L F; Ong, C K; Neo, C P; Varadan, V V; Varadan, Vijay K, Microwave Electronics:
Measurement and Materials Characterization, John Wiley & Sons, 2004 ISBN 0470020458.
Cohen, Nathan, "Fractal antenna and fractal resonator primer", ch. 8 in, Frame, Michael, Benoit
Mandelbrot: A Life In Many Dimensions, World Scientific, 2015 ISBN 9814366064.
Craig, Edwin C, Electronics via Waveform Analysis, Springer, 2012 ISBN 1461243386.
Doumanis, Efstratios; Goussetis, George; Kosmopoulos, Savvas, Filter Design for Satellite
Communications: Helical Resonator Technology, Artech House, 2015 ISBN 160807756X.
DuHamell, R; Isbell, D, "Broadband logarithmically periodic antenna structures" (https://doi.org/
10.1109/IRECON.1957.1150566), 1958 IRE International Convention Record, New York, 1957,
pp. 119–128.
Edwards, Terry C; Steer, Michael B, Foundations of Microstrip Circuit Design, John Wiley &
Sons, 2016 ISBN 1118936191.
Fagen, M D; Millman, S, A History of Engineering and Science in the Bell System: Volume 5:
Communications Sciences (1925–1980), AT&T Bell Laboratories, 1984 ISBN 0932764061.
Fano, R M; Lawson, A W, "Design of microwave filters", ch. 10 in, Ragan, G L (ed), Microwave
Transmission Circuits, McGraw-Hill, 1948 OCLC 2205252 (https://www.worldcat.org/oclc/2205
252).
Garg, Ramesh; Bahl, Inder; Bozzi, Maurizio, Microstrip Lines and Slotlines, Artech House, 2013
ISBN 1608075354.
Ghione, Giovanni; Pirola, Marco, Microwave Electronics, Cambridge University Press, 2017
ISBN 1107170273.
Grieg, D D; Englemann, H F, "Microstrip—a new transmission technique for the kilomegacycle
range" (https://doi.org/10.1109/JRPROC.1952.274144), Proceedings of the IRE, vol. 40, iss.
12, pp. 1644–1650, December 1952.
Gupta, S K, Electro Magnetic Field Theory, Krishna Prakashan Media, 2010
ISBN 8187224754.
Harrel, Bobby, The Cable Television Technical Handbook, Artech House, 1985
ISBN 0890061572.
Heaviside, Oliver, Electrical Papers, vol. 1, pp. 139–140, Copley Publishers, 1925
OCLC 3388033 (https://www.worldcat.org/oclc/3388033).
Heaviside, Oliver, "Electromagnetic induction and its propagation", The Electrician (https://dds.
crl.edu/crldelivery/19562), pp. 79–81, 3 June 1887 OCLC 6884353 (https://www.worldcat.org/o
clc/6884353).
Helszajn, J, Ridge Waveguides and Passive Microwave Components, IET, 2000
ISBN 0852967942.
Henderson, Bert; Camargo, Edmar, Microwave Mixer Technology and Applications, Artech
House, 2013 ISBN 1608074897.
Hilty, Kurt, "Attenuation measurement", pp. 422–439 in, Dyer, Stephen A (ed), Wiley Survey of
Instrumentation and Measurement, John Wiley & Sons, 2004 ISBN 0471221651.
Hong, Jia-Shen G; Lancaster, M J, Microstrip Filters for RF/Microwave Applications, John Wiley
& Sons, 2004 ISBN 0471464201.
Hunter, Ian, Theory and Design of Microwave Filters, IET, 2001 ISBN 0852967772.
Hura, Gurdeep S; Singhal, Mukesh, Data and Computer Communications: Networking and
Internetworking, CRC Press, 2001 ISBN 1420041312.
Ishii, T Koryu, Handbook of Microwave Technology: Components and devices, Academic
Press, 1995 ISBN 0123746965.
Janković, Nikolina; Zemlyakov, Kiril; Geschke, Riana Helena; Vendik, Irina; Crnojević-Bengin,
Vesna, "Fractal-based multi-band microstrip filters", ch. 6 in, Crnojević-Bengin, Vesna (ed),
Advances in Multi-Band Microstrip Filters, Cambridge University Press, 2015
ISBN 1107081971.
Johnson, Robert A, Mechanical Filters in Electronics, John Wiley & Sons Australia, 1983
ISBN 0471089192.
Johnson, Robert A; Börner, Manfred; Konno, Masashi, "Mechanical filters—a review of
progress" (https://doi.org/10.1109/T-SU.1971.29611), IEEE Transactions on Sonics and
Ultrasonics, vol. 18, iss. 3, pp. 155–170, July 1971.
Kumar, Narendra; Grebennikov, Andrei, Distributed Power Amplifiers for RF and Microwave
Communications, Artech House, 2015 ISBN 1608078329.
Lacomme, Philippe; Marchais, Jean-Claude; Hardange, Jean-Philippe; Normant, Eric, Air and
Spaceborne Radar Systems, William Andrew, 2001 ISBN 0815516134.
Lee, Thomas H, Planar Microwave Engineering, Cambridge University Press, 2004
ISBN 0521835267.
Levy, R; Cohn, S B, "A History of microwave filter research, design, and development" (https://d
oi.org/10.1109/TMTT.1984.1132817), IEEE Transactions: Microwave Theory and Techniques,
pp. 1055–1067, vol. 32, iss. 9, 1984.
Linkhart, Douglas K, Microwave Circulator Design, Artech House, 2014 ISBN 1608075834.
Magnusson, Philip C; Weisshaar, Andreas; Tripathi, Vijai K; Alexander, Gerald C, Transmission
Lines and Wave Propagation, CRC Press, 2000 ISBN 0849302692.
Makimoto, M; Yamashita, S, Microwave Resonators and Filters for Wireless Communication,
Springer, 2013 ISBN 3662043254.
Maloratsky, Leo G, Passive RF and Microwave Integrated Circuits, Elsevier, 2004
ISBN 0080492053.
Maloratsky, Leo G, Integrated Microwave Front-ends with Avionics Applications, Artech House,
2012 ISBN 1608072061.
Mason, Warren P, "Wave filter", U.S. Patent 2,345,491 (https://www.google.com/patents/US234
5491), filed 25 June 1927, issued 11 November 1930.
Mason, Warren P, "Wave transmission network", U.S. Patent 2,345,491 (https://www.google.co
m/patents/US2345491), filed 25 November 1941, issued 28 March 1944.
Mason, Warren P, "Electromechanical wave filter", U.S. Patent 2,981,905 (https://www.google.c
om/patents/US2981905), filed 20 August 1958, issued 25 April 1961.
Mason, W P; Sykes, R A, "The use of coaxial and balanced transmission lines in filters and
wide band transformers for high radio frequencies" (https://archive.org/stream/bellsystemtechni
16amerrich#page/274/mode/2up), Bell System Technical Journal, vol. 16, pp. 275–302, 1937.
Matthaei, G L, "Interdigital band-pass filters" (https://doi.org/10.1109/TMTT.1962.1125556), IRE
Transactions on Microwave Theory and Techniques, vol. 10, iss. 6, pp. 479–491, November
1962.
Matthaei, G L, "Comb-line band-pass filters of narrow or moderate bandwidth", Microwave
Journal, vol. 6, pp. 82–91, August 1963 ISSN 0026-2897 (https://www.worldcat.org/search?fq=x
0:jrnl&q=n2:0026-2897).
Matthaei, George L; Young, Leo; Jones, E M T, Microwave Filters, Impedance-Matching
Networks, and Coupling Structures McGraw-Hill 1964 OCLC 830829462 (https://www.worldcat.
org/oclc/830829462).
Meikle, Hamish, Modern Radar Systems, Artech House, 2008 ISBN 1596932430.
Milligan, Thomas A, Modern Antenna Design, John Wiley & Sons, 2005 ISBN 0471720607.
Misra, Devendra K, Radio-Frequency and Microwave Communication Circuits, John Wiley &
Sons, 2004 ISBN 0471478733.
Natarajan, Dhanasekharan, A Practical Design of Lumped, Semi-lumped & Microwave Cavity
Filters, Springer Science & Business Media, 2012 ISBN 364232861X.
Nguyen, Cam, Radio-Frequency Integrated-Circuit Engineering, John Wiley & Sons, 2015
ISBN 0471398209.
Ozaki, H; Ishii, J, "Synthesis of a class of strip-line filters" (https://doi.org/10.1109/TCT.1958.108
6441), IRE Transactions on Circuit Theory, vol. 5, iss. 2, pp. 104–109, June 1958.
Penn, Stuart; Alford, Neil, "Ceramic dielectrics for microwave applications", ch. 10 in, Nalwa,
Hari Singh (ed), Handbook of Low and High Dielectric Constant Materials and Their
Applications, Academic Press, 1999 ISBN 0080533531.
Polkinghorn, Frank A, "Oral-History: Warren P. Mason" (http://ethw.org/Oral-History:Warren_P._
Mason), interview no. 005 for the IEEE History Centre, 3 March 1973, Engineering and
Technology History Wiki, retrieved 15 April 2018.
Ramadan, Ali; Al-Husseini, Mohammed; Kabalan Karim Y; El-Hajj, Ali, "Fractal-shaped
reconfigurable antennas", ch. 10 in, Nasimuddin, Nasimuddin, Microstrip Antennas, BoD –
Books on Demand, 2011 ISBN 9533072474.
Richards, Paul I, "Resistor-transmission-line circuits" (https://doi.org/10.1109/JRPROC.1948.23
3274), Proceedings of the IRE, vol. 36, iss. 2, pp. 217–220, 1948.
Richtmeyer, R D, "Dielectric resonators" (https://doi.org/10.1063/1.1707320), Journal of Applied
Physics, vol. 10, iss. 6, pp. 391–397, June 1939.
Roer, T G, Microwave Electronic Devices, Springer, 2012 ISBN 1461525004.
Sharma, K K, Fundamental of Microwave and Radar Engineering, S. Chand Publishing, 2011
ISBN 8121935377.
Sheingold, L S; Morita, T, "A coaxial magic-T" (https://ieeexplore.ieee.org/document/1124845/),
Transactions of the IRE Professional Group on Microwave Theory and Techniques, vol. 1, iss.
2, pp. 17–23, November 1953.
Sisodia, M L; Raghuvanshi, G S, Basic Microwave Techniques and Laboratory Manual, New
Age International, 1987 ISBN 0852268580.
Taylor, John; Huang, Qiuting, CRC Handbook of Electrical Filters, CRC Press, 1997
ISBN 0849389518.
Tyrrell, W A, "Hybrid circuits for microwaves" (https://doi.org/10.1109/JRPROC.1947.233572),
Proceedings of the IRE, vol. 35, iss. 11, pp. 1294–1306, November 1947.
Vendelin, George D; Pavio, Anthony M; Rohde, Ulrich L, Microwave Circuit Design Using
Linear and Nonlinear Techniques, John Wiley & Sons, 2005 ISBN 0471715824.
Whitaker, Jerry C, The Resource Handbook of Electronics, CRC Press, 2000
ISBN 1420036866.
Zhurbenko, Vitaliy, Passive Microwave Components and Antennas, BoD – Books on Demand,
2010 ISBN 9533070838.

Retrieved from "https://en.wikipedia.org/w/index.php?title=Distributed-element_circuit&oldid=985623997"

This page was last edited on 27 October 2020, at 00:06 (UTC).

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using this
site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

You might also like