2009–10 Paris Saint-Germain F.C. (Ladies) season: Difference between revisions

From formulasearchengine
Jump to navigation Jump to search
en>Yobot
m WP:CHECKWIKI error fixes / special characters in sortkey fixed using AWB (9440)
en>Cydebot
m Robot - Speedily moving category Paris Saint-Germain F.C. (Ladies) seasons to Category:Paris Saint-Germain Féminines seasons per CFDS.
 
Line 1: Line 1:
[[File:Waveguide-post-filter.JPG|thumb|right|alt=photo|'''Figure 1'''. Waveguide post filter: a band-pass filter consisting of a length of WG15 (a standard waveguide size for [[X band]] use) divided into a row of five [[coupled oscillation|coupled]] [[microwave cavity|resonant cavities]] by fences of three posts each. The ends of the posts can be seen protruding through the wall of the guide.]]
While everyone has heard of hypnosis, most peoples ideas of what it's all about come from seeing stage or street hypnotists, seeing it [http://tinyurl.com/ku6vjks uggs on sale] TV or in films, or perhaps reading about it in fiction books.<br><br>This leads to a lot of myths and misconceptions about hypnosis that can lead to people being nervous, or even afraid, of seeing a hypnotherapist for help. That's why I've compiled this list of the top 10 myths about hypnosis and hypnotherapy <br><br>1. Hypnosis is magic.<br>Hypnosis is not magic. It is not connected in any way to the occult or the paranormal. In fact hypnosis is a naturally occurring state that you have probably experienced many times, without even being aware of it. <br>2. Hypnosis is going to sleep. <br>The truth is that, while you are in a trance, you will still know what's going [http://tinyurl.com/ku6vjks uggs on sale] around you, just as when you are normally awake. If you are not aware of what's going [http://tinyurl.com/ku6vjks uggs on sale], at least at the subconscious level then you will be unlikely to get much benefit from the session. <br><br>3. I will forget what happened to me while I was in a trance. Typically you will have complete recall of everything that transpires while you are in a trance, unless of course you make a choice not to remember something. For example, if you regressed and experienced a painful or traumatic event from your past that you don't yet feel ready to deal with you may choose not to remember it until you are ready to deal with it.<br><br>4. I will not be in control when I am hypnotised.<br>You will always be in control. Hypnosis requires your total cooperation. If I, as a hypnotherapist, do anything, or suggest that you do anything, that you don't agree with, you will simply reject it, and maybe come out of the hypnotic state. <br>5. I could be force to give away secrets when I am in a trance. <br><br>Hypnosis is not a truth serum. You will only say what you want to say, you are in control the whole time. <br>6. I may get 'stuck' in a trance. <br>Anytime you want to come out of the trance you will be able to. While it's true that occasionally people don't want to come out of trance, it's only because they are enjoying the sensation so much and don;t want it to end. There are various ways to make sure that you come out of trance easily and safely.<br><br>In fact, even if I did nothing to 'wake you up' you would come out [http://tinyurl.com/ku6vjks uggs on sale] your own when you were ready, or when something happened that you needed to deal with. <br>7. I may be given Post Hypnotic Suggestions to do things I don't want to do. Post Hypnotic Suggestions are used in Hypnotherapy, often to directly address the problem being worked [http://tinyurl.com/ku6vjks uggs on sale], or to make going into a hypnotic state in later sessions much quicker and easier.<br><br>As with any [http://En.Wiktionary.org/wiki/suggestions suggestions] you will only accept those suggestions that you agree to and that are for your higher good. Any suggestions that you do not agree with will simply be rejected. <br>8. Very intelligent people can't be hypnotised. <br>Actually the reverse is normally true. Intelligent people normally find it easier to achieve a hypnotic state. <br><br>9. Strong willed people can't be hypnotised.<br>This one is only true if the person doesn't want to be hypnotised, in which case it doesn't really matter whether you are strong willed or not. If you don't want to be hypnotised then you won't be. <br>10. I've never experienced a hypnotic state before.<br>I hear this one a lot. Some people are totally convinced they have never been in a trance, when the truth is they will experience one regularly. <br><br>If you drive a car then you may remember a time when you have driven a familiar journey and when it's almost over you suddenly realise that you have no real recollection of the journey up to that point. Or maybe you have become so engrossed in a film, or book, that the rest of the world seems to fade away.<br><br>These are all [http://photo.net/gallery/tag-search/search?query_string=examples examples] of a naturally occurring light hypnotic state.
 
A '''waveguide filter''' is an [[electronic filter]] that is constructed with [[waveguide (electromagnetism)|waveguide]] technology. Waveguides are hollow metal tubes inside which an [[electromagnetic wave]] may be transmitted. Filters are devices used to allow signals at some frequencies to pass (the [[passband]]), while others are rejected (the [[stopband]]). Filters are a basic component of [[electronic engineering]] designs and have numerous applications. These include [[Selectivity (electronic)|selection]] of [[Signal (electrical engineering)|signals]] and limitation of [[Noise (electronics)|noise]]. Waveguide filters are most useful in the [[microwave]] band of frequencies, where they are a convenient size and have low [[insertion loss|loss]].  Examples of [[microwave filter]] use are found in [[satellite communications]], [[telephone network]]s, and [[television broadcasting]].
 
Waveguide filters were [[Technology during World War II|developed during World War II]] to meet the needs of [[radar]] and [[electronic countermeasure]]s, but afterwards soon found civilian applications such as use in [[microwave link]]s. Much of post-war development was concerned with reducing the bulk and weight of these filters, first by using new analysis techniques that led to elimination of unnecessary components, then by innovations such as dual-mode [[cavity resonator|cavities]] and novel materials such as [[ceramic resonator]]s.
 
A particular feature of waveguide filter design concerns the [[normal mode|mode]] of transmission. Systems based on pairs of [[Electrical conductor|conducting]] wires and similar technologies have only one mode of transmission. In waveguide systems, any number of modes are possible.  This can be both a disadvantage, as spurious modes frequently cause problems, and an advantage, as a dual-mode design can be much smaller than the equivalent waveguide single mode design. The chief advantages of waveguide filters over other technologies are their ability to handle high power and their low loss.  The chief disadvantages are their bulk and cost when compared with technologies such as [[microstrip]] filters.
 
There is a wide array of different types of waveguide filters. Many of them consist of a chain of coupled resonators of some kind that can be modelled as a [[ladder network]] of [[LC circuit]]s.  One of the most common types consists of a number of coupled [[resonant cavity|resonant cavities]].  Even within this type, there are many subtypes, mostly differentiated by the means of [[Coupling (electronics)|coupling]].  These coupling types include apertures,{{glosslink|aperture|w}} irises,{{glosslink|iris|x}} and posts. Other waveguide filter types include [[dielectric resonator]] filters, insert filters, finline filters, corrugated-waveguide filters, and stub filters. A number of waveguide components have [[filter theory]] applied to their design, but their purpose is something other than to filter signals. Such devices include [[impedance matching]] components, [[directional coupler]]s, and [[diplexer]]s. These devices frequently take on the form of a filter, at least in part.
 
==Scope==
 
The common meaning of ''waveguide'', when the term is used unqualified, is the hollow metal kind, but other waveguide technologies are possible.<ref>Gibilisco & Sclater, [https://www.google.co.uk/search?tbm=bks&q=%22The+term+waveguide+has+come+to+mean+a+hollow+metal+tube%22 page 913]</ref> The scope of this article is limited to the metal-tube type. The [[#Post-wall waveguide|post-wall waveguide]] structure is something of a variant, but is related enough to include in this article—the wave is mostly surrounded by conducting material. It is possible to construct [[waveguide (optics)|waveguides out of dielectric rods]],<ref>Yeh & Shimabukuro, page 1</ref> the most well known example being [[optical fibre]]s. This subject is outside the scope of the article with the exception that dielectric rod resonators are sometimes used ''inside'' hollow metal waveguides. [[Transmission line]]{{glosslink|tl|o}} technologies such as conducting wires and microstrip can be thought of as waveguides,<ref>Russer, pages 131–132</ref> but are not commonly called such and are also outside the scope of this article.
 
==Basic concepts==
 
===Filters===
In [[electronics]], [[Filter (signal processing)|filters]] are used to allow signals of a certain band of [[frequency|frequencies]] to pass while blocking others.  They are a basic building block of electronic systems and have a great many applications.  Amongst the uses of waveguide filters are the construction of [[duplexer]]s, [[diplexer]]s,{{glosslink|Dx|d}} and [[multiplexer]]s; [[Selectivity (electronic)|selectivity]] and [[Noise (electronics)|noise]] limitation in [[Receiver (radio)|receivers]]; and [[harmonic distortion]] suppression in [[transmitter]]s.<ref>Belov ''et al.'', page 147</ref>
 
===Waveguides===
[[Waveguide (electromagnetism)|Waveguide]]s are metal conduits used to confine and direct radio signals.  They are usually made of brass, but aluminium and copper are also used.<ref>Connor, page 52</ref>  Most commonly they are rectangular, but other [[Multiview orthographic projection#Cross-section|cross-sections]] such as circular or elliptical are possible.  A waveguide filter is a filter composed of waveguide components.  It has much the same range of applications as other filter technologies in electronics and radio engineering but is very different mechanically and in principle of operation.<ref>{{multiref|Hunter, page 201|Matthaei ''et al.'', page 243}}</ref>
 
The technology used for constructing filters is chosen to a large extent by the frequency of operation that is expected, although there is a large amount of overlap.  Low frequency applications such as [[audio electronics]] use filters composed of discrete [[capacitor]]s and [[inductor]]s.  Somewhere in the [[very high frequency]] band, designers switch to using components made of pieces of transmission line.{{glosslink|tl|p}}  These kinds of designs are called [[distributed element filter]]s.  Filters made from discrete components are sometimes called [[lumped element model|lumped element]] filters to distinguish them.  At still higher frequencies, the [[microwave]] bands, the design switches to waveguide filters, or sometimes a combination of waveguides and transmission lines.<ref>{{multiref|Hitchcock & Patterson, page 263|Bagad, pages 1.3–1.4}}</ref>
 
Waveguide filters have much more in common with transmission line filters than lumped element filters; they do not contain any discrete capacitors or inductors.  However, the waveguide design may frequently be equivalent (or approximately so) to a lumped element design.  Indeed, the design of waveguide filters frequently starts from a lumped element design and then converts the elements of that design into waveguide components.<ref>Matthaei ''et al.'', page 83</ref>
 
===Modes===
[[File:Selected modes.svg|thumb|alt=diagram|'''Figure 2.'''  The field patterns of some common waveguide modes]]
One of the most important differences in the operation of waveguide filters compared to transmission line designs concerns the mode of transmission of the [[electromagnetic wave]] carrying the signal.  In a transmission line, the wave is associated with electric currents on a pair of conductors.  The conductors constrain the currents to be parallel to the line, and consequently both the magnetic and electric components of the [[electromagnetic field]] are perpendicular to the direction of travel of the wave.  This [[transverse mode]] is designated TEM{{glosslink|TEM|l}} (transverse electromagnetic).  On the other hand, there are infinitely many modes that any completely hollow waveguide can support, but the TEM mode is not one of them. Waveguide modes are designated either TE{{glosslink|TE|m}} (transverse electric) or TM{{glosslink|TM|n}} (transverse magnetic), followed by a pair of suffixes identifying the precise mode.<ref>{{multiref|Connor, pages 52–53|Hunter, pages 201, 203|Matthaei ''et al.'', page 197}}</ref>
 
This multiplicity of modes can cause problems in waveguide filters when spurious modes are generated.  Designs are usually based on a single mode and frequently incorporate features to suppress the unwanted modes.  On the other hand, advantage can be had from choosing the right mode for the application, and even sometimes making use of more than one mode at once.  Where only a single mode is in use, the waveguide can be modelled like a conducting transmission line and results from transmission line theory can be applied.<ref>{{multiref|Hunter, pages 255–260|Matthaei ''et al.'', page 197}}</ref>
 
===Cutoff===
Another feature peculiar to waveguide filters is that there is a definite frequency, the [[Cutoff frequency#Waveguides|cutoff frequency]], below which no transmission can take place.  This means that in theory [[low-pass filter]]s cannot be made in waveguides.  However, designers frequently take a lumped element low-pass filter design and convert it to a waveguide implementation.  The filter is consequently low-pass by design and may be considered a low-pass filter for all practical purposes if the cutoff frequency is below any frequency of interest to the application.  The waveguide cutoff frequency is a function of transmission mode, so at a given frequency, the waveguide may be usable in some modes but not others.  Likewise, the [[guide wavelength]]{{glosslink|λg|h}} (λ<sub>g</sub>) and [[characteristic impedance]]{{glosslink|Z0|b}} (''Z''<sub>0</sub>) of the guide at a given frequency also depend on mode.<ref>{{multiref|Hunter, pages 201–202|Matthaei ''et al.'', page 197}}</ref>
 
===Dominant mode===
The mode with the lowest cutoff frequency of all the modes is called the dominant mode.  Between cutoff and the next highest mode, this is the only mode it is possible to transmit, which is why it is described as dominant.  Any spurious modes generated are rapidly attenuated along the length of the guide and soon disappear.  Practical filter designs are frequently made to operate in the dominant mode.<ref>{{multiref|Elmore & Heald, page 289|Mahmoud, pages 32–33}}</ref>
 
In rectangular waveguide, the TE<sub>10</sub>{{glosslink|TE10|q}} mode (shown in figure 2) is the dominant mode.  There is a band of frequencies between the dominant mode cutoff and the next highest mode cutoff in which the waveguide can be operated without any possibility of generating spurious modes.  The next highest cutoff modes are TE<sub>20</sub>,{{glosslink|TE20|r}} at exactly twice the TE<sub>10</sub> mode, and TE<sub>01</sub>{{glosslink|TE01|s}} which is also twice TE<sub>10</sub> if the waveguide used has the commonly used [[aspect ratio]] of 2:1.  The lowest cutoff TM mode is TM<sub>11</sub>{{glosslink|TM11|t}} (shown in figure 2) which is <math>\scriptstyle \sqrt 5</math> times the dominant mode in 2:1 waveguide.  Thus, there is an [[octave]] over which the dominant mode is free of spurious modes, although operating too close to cutoff is usually avoided because of phase distortion.<ref>{{multiref|Hunter, page 209, |Matthaei ''et al.'', page 198}}</ref>
 
In circular waveguide, the dominant mode is TE<sub>11</sub>{{glosslink|TE11c|u}} and is shown in figure 2.  The next highest mode is TM<sub>01</sub>.{{glosslink|TM01c|v}}  The range over which the dominant mode is guaranteed to be spurious-mode free is less than that in rectangular waveguide; the ratio of highest to lowest frequency is approximately 1.3 in circular waveguide, compared to 2.0 in rectangular guide.<ref>Matthaei ''et al.'', pages 198, 201</ref>
 
===Evanescent modes===
{{main|Evanescent wave}}
[[Evanescent wave#Evanescent-wave coupling|Evanescent mode]]s are modes below the cutoff frequency.  They cannot propagate down the waveguide for any distance, dying away exponentially.  However, they are important in the functioning of certain filter components such as irises and posts, described later, because energy is stored in the evanescent wave fields.<ref>Das & Das, page 112</ref>
 
==Advantages and disadvantages==
Like transmission line filters, waveguide filters always have multiple [[passband]]s, replicas of the lumped element [[prototype filter|prototype]].  In most designs, only the lowest frequency passband is useful (or lowest two in the case of [[band-stop filter]]s) and the rest are considered unwanted spurious artefacts.  This is an intrinsic property of the technology and cannot be designed out, although design can have some control over the frequency position of the spurious bands.  Consequently, in any given filter design, there is an upper frequency beyond which the filter will fail to carry out its function.  For this reason, true low-pass and [[high-pass filter]]s cannot exist in waveguide.  At some high frequency there will be a spurious passband or stopband interrupting the intended function of the filter.  But, similar to the situation with waveguide cutoff frequency, the filter can be designed so that the edge of the first spurious band is well above any frequency of interest.<ref>{{multiref|Lee, page 789|Matthaei ''et al.'', page 541|Sorrentino & Bianchi, page 262}}</ref>
 
The range of frequencies over which waveguide filters are useful is largely determined by the waveguide size needed.  At lower frequencies the waveguide needs to be impractically large in order to keep the cutoff frequency below the operational frequency.  On the other hand, filters whose operating frequencies are so high that the wavelengths are sub-millimetre cannot be manufactured with normal [[machine shop]] processes.  At frequencies this high, fibre-optic technology starts to become an option.<ref>{{multiref|Hunter, page 201|Eskelinen & Eskelinen, page 269|Middleton & Van Valkenburg, pages 30.26–30.28}}</ref>
 
Waveguides are a low-loss medium.  Losses in waveguides mostly come from [[Joule heating|ohmic]] dissipation caused by currents induced in the waveguide walls.  Rectangular waveguide has lower loss than circular waveguide and is usually the preferred format, but the TE<sub>01</sub> circular mode is very low loss and has applications in long distance communications.  Losses can be reduced by polishing the internal surfaces of the waveguide walls.  In some applications which require rigorous filtering, the walls are plated with a thin layer of gold or silver to improve surface [[electrical conductivity|conductivity]].  An example of such requirements is satellite applications which require low loss, high selectivity, and linear group delay from their filters.<ref>{{multiref|Belov ''et al.'', page 147|Connor, pages 6, 64|Hunter, page 230|Matthaei ''et al.'', page 243}}</ref>
 
One of the main advantages of waveguide filters over TEM mode technologies is the quality of their [[resonator]]s.  Resonator quality is characterised by a parameter called [[Q factor]], or just ''Q''.  The ''Q'' of waveguide resonators is in the thousands, orders of magnitude higher than TEM mode resonators.<ref>{{multiref|Sorrentino & Bianchi, page 691|Hunter, page 201}}</ref>  The [[electrical resistance|resistance]] of conductors, especially in wound inductors, limits the ''Q'' of TEM resonators.  This improved ''Q'' leads to better performing filters in waveguides, with greater stop band rejection.  The limitation to ''Q'' in waveguides comes mostly from the ohmic losses in the walls described earlier, but silver plating the internal walls can more than double ''Q''.<ref>Hunter, pages 201, 230</ref>
 
Waveguides have good power handling capability, which leads to filter applications in [[radar]].<ref>{{multiref|Belov ''et al.'', page 147|Bowen, page 114}}</ref>  Despite the performance advantages of waveguide filters, [[microstrip]] is often the preferred technology due to its low cost.  This is especially true for consumer items and the lower microwave frequencies.  Microstrip circuits can be manufactured by cheap [[printed circuit]] technology, and when integrated on the same printed board as other circuit blocks they incur little additional cost.<ref>{{multiref|Das & Das, page 310|Waterhouse, page 8}}</ref>
 
==History==
[[File:PSM V25 D738 John William Strutt Lord Rayleigh.jpg|thumb|right|upright=0.55|alt=likeness of|[[Lord Rayleigh]] first suggested waveguide transmission.]]
The idea of a waveguide for electromagnetic waves was first suggested by [[Lord Rayleigh]] in 1897.  Rayleigh proposed that a [[coaxial cable|coaxial transmission line]] could have the centre conductor removed, and waves would still propagate down the inside of the remaining cylindrical conductor despite there no longer being a complete electrical circuit of conductors.  He described this in terms of the wave reflecting repeatedly off the internal wall of the outer conductor in a zig-zag fashion as it progressed down the waveguide.  Rayleigh was also the first to realise that there was a critical wavelength, the cutoff wavelength, proportional to the cylinder diameter, above which wave propagation is not possible.  Waveguides were first developed, in a circular form, by [[George Clark Southworth]] and J. F. Hargreaves in 1932.<ref>Sarkar ''et al.'', pages 90, 129</ref>
 
The first [[analogue filter]] design which went beyond a simple single resonator was created by [[George Ashley Campbell]] in 1910 and marked the beginning of filter theory.  Campbell's filter was a lumped-element design of capacitors and inductors suggested by his work with [[loading coil]]s.  [[Otto Zobel]] and others quickly developed this further.<ref>Bray, page 62</ref>  Development of distributed element filters began in the years before World War II. A major paper on the subject was published by Mason and Sykes in 1937;<ref>{{multiref|Levy & Cohn, page 1055|See also Mason & Sykes (1937)}}</ref> a patent<ref>Mason, Warren P., "Wave filter", {{US patent|1781469}}, filed: {{Nowrap|25 June}} 1927, issued: {{Nowrap|11 November}} 1930.</ref> filed by Mason in 1927 may contain the first published filter design using distributed elements.<ref>Millman ''et al.'', page 108</ref>
 
[[File:Hans Bethe.jpg|thumb|left|upright=0.55|alt=photo|[[Hans Bethe]] developed waveguide aperture theory.]]
Mason and Sykes' work was focused on the formats of coaxial cable and [[balanced pair]]s of wires, but other researchers later applied the principles to waveguides as well. Much development on waveguide filters was carried out during World War II driven by the filtering needs of radar and [[electronic countermeasure]]s. A good deal of this was at the [[MIT Radiation Laboratory]] (Rad Lab), but other laboratories in the US and the UK were also involved such as the [[Telecommunications Research Establishment]] in the UK.  Amongst the well-known scientists and engineers at Rad Lab were [[Julian Schwinger]], [[Nathan Marcuvitz]], [[Edward Mills Purcell]], and [[Hans Bethe]].  Bethe was only at Rad Lab a short time but produced his aperture theory while there.  Aperture theory is important for waveguide cavity filters, which were first developed at Rad Lab.  Their work was published after the war in 1948 and includes an early description of dual-mode cavities by Fano and Lawson.<ref>{{multiref|Levy & Cohn, pages 1055, 1057|See also Fano and Lawson (1948)}}</ref>
 
Theoretical work following the war included the commensurate line theory of [[Paul Richards (scientist)|Paul Richards]]. Commensurate lines are networks in which all the elements are the same length (or in some cases multiples of the unit length), although they may differ in other dimensions to give different characteristic impedances.{{glosslink|Z0|a}} [[Richards' transformation]] allows any lumped element design to be taken "as is" and transformed directly into a distributed element design using a very simple transform equation.  In 1955 K. Kuroda published the transformations known as [[Kuroda's identities]].  These made Richard's work more usable in [[unbalanced line|unbalanced]] and waveguide formats by eliminating the problematic [[Series and parallel circuits|series]] connected elements, but it was some time before Kuroda's Japanese work became widely known in the English speaking world.<ref>{{multiref|Levy and Cohn, pages 1056–1057|See also Richards (1948)}}</ref>  Another theoretical development was the [[network synthesis filter]] approach of [[Wilhelm Cauer]] in which he used the [[Chebyshev approximation]] to determine element values.  Cauer's work was largely developed during World War II (Cauer was killed towards the end of it), but could not be widely published until hostilities ended.  While Cauer's work concerns lumped elements, it is of some importance to waveguide filters; the [[Chebyshev filter]], a special case of Cauer's synthesis, is widely used as a prototype filter for waveguide designs.<ref>{{multiref|Cauer ''et al.'', pages 3, 5|Mansour, page 166}}</ref>
 
Designs in the 1950s started with a lumped element prototype (a technique still in use today), arriving after various transformations at the desired filter in a waveguide form.  At the time, this approach was yielding [[fractional bandwidth]]s no more than about {{sfrac|1|5}}.  In 1957, Leo Young at [[Stanford Research Institute]] published a method for designing filters which ''started'' with a distributed element prototype, the stepped impedance prototype.  This filter was based on [[quarter-wave impedance transformer]]s of various widths and was able to produce designs with bandwidths up to an [[Octave (electronics)|octave]] (a fractional bandwidth of {{sfrac|2|3}}).  Young's paper specifically addresses directly coupled cavity resonators, but the procedure can equally be applied to other directly coupled resonator types.<ref>{{multiref|Levy & Cohn, page 1056|See also Young (1963)}}</ref>
[[File:Pierce cross-coupled filter.png|thumb|alt=drawing|'''Figure 3.'''  Pierce's waveguide implementation of a cross-coupled filter]]
The first published account of a cross-coupled filter is due to [[John R. Pierce]] at [[Bell Labs]] in a 1948 patent.<ref>{{multiref|Pierce, J. R., "Guided wave frequency range transducer", {{US patent|2626990}}, filed: 4 May 1948, issued: 27 January 1953.|See also Pierce (1949)}}</ref>  A cross-coupled filter is one in which resonators that are not immediately adjacent are coupled.  The additional [[Degrees of freedom (physics and chemistry)|degrees of freedom]] thus provided allow the designer to create filters with improved performance, or, alternatively, with fewer resonators.  One version of Pierce's filter, shown in figure 3, uses circular waveguide cavity resonators to link between rectangular guide cavity resonators.  This principle was not at first much used by waveguide filter designers, but it was used extensively by [[mechanical filter]] designers in the 1960s, particularly R. A. Johnson at [[Collins Radio Company]].<ref>Levy & Cohn, pages 1060–1061</ref>
 
The initial non-military application of waveguide filters was in the [[microwave link]]s used by telecommunications companies to provide the [[backbone network|backbone]] of their networks.  These links were also used by other industries with large, fixed networks, notably television broadcasters.  Such applications were part of large capital investment programs.  They are now also used in [[satellite communications]] systems.<ref>{{multiref|Hunter, page 230|Huurdeman, pages 369–371}}</ref>
 
The need for frequency-independent delay in satellite applications led to more research into the waveguide incarnation of cross-coupled filters.  Previously, satellite communications systems used a separate component for [[delay equalisation]].  The additional degrees of freedom obtained from cross-coupled filters held out the possibility of designing a flat delay into a filter without compromising other performance parameters.  A component that simultaneously functioned as both filter and equaliser would save valuable weight and space.  The needs of satellite communication also drove research into the more exotic resonator modes in the 1970s.  Of particular prominence in this respect is the work of E. L. Griffin and F. A. Young, who investigated better modes for the {{nowrap|12-14 GHz}} band when this began to be used for satellites in the mid-1970s.<ref>{{multiref|Levy & Cohn, pages 1061–1062|See also Griffin & Young (1978)}}</ref>
 
Another space-saving innovation was the [[dielectric resonator]], which can be used in other filter formats as well as waveguide.  The first use of these in a filter was by S. B. Cohn in 1965, using [[titanium dioxide]] as the dielectric material.  Dielectric resonators used in the 1960s, however, had very poor temperature coefficients, typically 500 times worse than a mechanical resonator made of [[invar]], which led to instability of filter parameters.  Dielectric materials of the time with better temperature coefficients had too low a [[dielectric constant]] to be useful for space saving.  This changed with the introduction of ceramic resonators with very low temperature coefficients in the 1970s.  The first of these was from Massé and Pucel using [[barium tetratitanate]]<ref group=note>Barium tetratitanate, BaTi<sub>4</sub>O<sub>9</sub> (Young ''et al.'', page 655)</ref> at [[Raytheon]] in 1972.  Further improvements were reported in 1979 by Bell Labs and [[Murata Manufacturing]].  Bell Labs' [[barium nonatitanate]]<ref group=note>Barium nonatitanate, Ba<sub>2</sub>Ti<sub>9</sub>O<sub>20</sub> (Nalwa, page 443)</ref> resonator had a dielectric constant of 40 and [[Q factor|''Q'']] of 5000–10,000 at {{nowrap|2-7 GHz}}.  Modern temperature-stable materials have a dielectric constant of about 90 at microwave frequencies, but research is continuing to find materials with both low loss and high permittivity; lower permittivity materials, such as [[zirconium stannate titanate]]<ref group=note>Zirconium stannate titanate, Zr<sub>1-''x''</sub>Sn<sub>x</sub>TiO<sub>4</sub> (Gusmano ''et al.'', page 690)</ref> (ZST) with a dielectric constant of 38, are still sometimes used for their low loss property.<ref>{{multiref|Levy & Cohn, pages 1062–1063|Nalwa, pages 525–526|See also:<br />Maasé & Pucel (1972)|Cohn (1965)}}</ref>
 
An alternative approach to designing smaller waveguide filters was provided by the use of non-propagating evanescent modes. Jaynes and Edson proposed evanescent mode waveguide filters in the late 1950s. Methods for designing these filters were created by Craven and Young in 1966. Since then, evanescent mode waveguide filters have seen successful use where waveguide size or weight are important considerations.<ref>Zhang, Wang, Li, and Lui (2008)</ref>
 
A relatively recent technology being used inside hollow-metal-waveguide filters is finline, a kind of planar dielectric waveguide.  Finline was first described by Paul Meier in 1972.<ref>{{multiref|Srivastava &Gupta, page 82|See also: Meier (1972)}}</ref>
 
===Multiplexer history===
[[File:John Robinson Pierce head.jpg|thumb|right|upright=0.55|alt=photo|[[John R. Pierce]] invented the cross-coupled filter and the contiguous passband multiplexer.]]
[[Frequency-division multiplexing|Multiplexers]] were first described by Fano and Lawson in 1948.  Pierce was the first to describe multiplexers with contiguous passbands.  Multiplexing using directional filters was invented by Seymour Cohn and Frank Coale in the 1950s.  Multiplexers with compensating [[immittance]] resonators at each junction are largely the work of E. G. Cristal and G. L. Matthaei in the 1960s.  This technique is still sometimes used, but the modern availability of computing power has led to the more common use of synthesis techniques which can directly produce matching filters without the need for these additional resonators.  In 1965 R. J. Wenzel discovered that filters which were singly terminated,{{glosslink|1term|k}} rather than the usual doubly terminated, were complementary—exactly what was needed for a diplexer.{{glosslink|Dx|c}}  Wenzel was inspired by the lectures of circuit theorist [[Ernst Guillemin]].<ref>{{multiref|Levy & Cohn, page 1065|See also:<br />Fano & Lawson (1948)|Pierce (1949)|Cristal & Matthaei (1964)|Wenzel (1969)}}</ref>
 
Multi-channel, multi-octave multiplexers were investigated by Harold Schumacher at Microphase Corporation, and his results were published in 1976.  The principle that multiplexer filters may be matched when joined together by modifying the first few elements, thus doing away with the compensating resonators, was discovered accidentally by E. J. Curly around 1968 when he mistuned a diplexer.  A formal theory for this was provided by J. D. Rhodes in 1976 and generalised to multiplexers by Rhodes and Ralph Levy in 1979.<ref>{{multiref|Levy & Cohn, pages 1064–1065|See also:<br>Schumacher (1976)|Rhodes (1976)|Rhodes & Levy (1979)}}</ref>
 
From the 1980s, planar technologies, especially microstrip, have tended to replace other technologies used for constructing filters and multiplexers, especially in products aimed at the consumer market.  The recent innovation of post-wall waveguide allows waveguide designs to be implemented on a flat substrate with low-cost manufacturing techniques similar to those used for microstrip.<ref>{{multiref|Levy & Cohn, page 1065|Xuan & Kishk, page 1}}</ref>
 
==Components==
[[File:Cauer lowpass.svg|thumb|alt=diagram|'''Figure 4.'''  Ladder circuit implementation of a lumped element low-pass filter]]
Waveguide filter designs frequently consist of two different components repeated a number of times.  Typically, one component is a resonator or discontinuity with a lumped circuit equivalent of an inductor, capacitor, or LC resonant circuit.  Often, the filter type will take its name from the style of this component.  These components are spaced apart by a second component, a length of guide which acts as an impedance transformer.  The impedance transformers have the effect of making alternate instances of the first component appear to be a different impedance.  The net result is a lumped element equivalent circuit of a ladder network.  Lumped element filters are commonly [[ladder topology]], and such a circuit is a typical starting point for waveguide filter designs.  Figure 4 shows such a ladder.  Typically, waveguide components are resonators, and the equivalent circuit would be [[LC circuit|LC resonators]] instead of the capacitors and inductors shown, but circuits like figure 4 are still used as [[prototype filter]]s with the use of a band-pass or band-stop transformation.<ref>Matthaei ''et al.'', pages 427–440</ref>
 
Filter performance parameters, such as stopband rejection and rate of transition between passband and stopband, are improved by adding more components and thus increasing the length of the filter.  Where the components are repeated identically, the filter is an [[image parameter filter]] design, and performance is enhanced simply by adding more identical elements.  This approach is typically used in filter designs which use a large number of closely spaced elements such as the [[waffle-iron filter]].  For designs where the elements are more widely spaced, better results can be obtained using a network synthesis filter design, such as the common Chebyshev filter and [[Butterworth filter]]s.  In this approach the circuit elements do not all have the same value, and consequently the components are not all the same dimensions.  Furthermore, if the design is enhanced by adding more components then all the element values must be calculated again from scratch.  In general, there will be no common values between the two instances of the design.  Chebyshev waveguide filters are used where the filtering requirements are rigorous, such as satellite applications.<ref name="Hunterpage">Hunter, page 230</ref><ref>Matthaei ''et al.'', pages 83–84</ref>
 
===Impedance transformer===
An impedance transformer is a device which makes an impedance at its output [[port (circuit theory)|port]] appear as a different impedance at its input port.  In waveguide, this device is simply a short length of waveguide.  Especially useful is the [[quarter-wave impedance transformer]] which has a length of λ<sub>g</sub>/4.  This device can turn [[capacitance]]s into [[inductance]]s and vice versa.<ref>Matthaei ''et al.'', pages 144–145</ref>  It also has the useful property of turning shunt-connected elements into series-connected elements and vice versa.  Series-connected elements are otherwise difficult to implement in waveguide.<ref>Matthaei ''et al.'', pages 595–596</ref>
 
===Reflections and discontinuities===
Many waveguide filter components work by introducing a sudden change, a discontinuity, to the transmission properties of the waveguide.  Such discontinuities are equivalent to lumped impedance elements placed at that point.  This arises in the following way: the discontinuity causes a partial reflection of the transmitted wave back down the guide in the opposite direction, the ratio of the two being known as the [[reflection coefficient]].  This is entirely analogous to a [[Reflections of signals on conducting lines|reflection on a transmission line]] where there is an established relationship between reflection coefficient and the impedance that caused the reflection.  This impedance must be [[electrical reactance|reactive]], that is, it must be a capacitance or an inductance.  It cannot be a resistance since no energy has been absorbed—it is all either transmitted onward or reflected.  Examples of components with this function include irises, stubs, and posts, all described later in this article under the filter types in which they occur.<ref>Montgomery ''et al.'', page 162</ref>
 
===Impedance step===
An impedance step is an example of a device introducing a discontinuity.  It is achieved by a step change in the physical dimensions of the waveguide.  This results in a step change in the characteristic impedance of the waveguide.  The step can be in either the [[E-plane]]{{glosslink|E-plane|f}} (change of height{{glosslink|height|j}}) or the [[H-plane]]{{glosslink|H-plane|g}} (change of width{{glosslink|height|i}}) of the waveguide.<ref>Das & Das, pages 134–135</ref>
 
==Resonant cavity filter==
 
===Cavity resonator===
A basic component of waveguide filters is the [[microwave cavity|cavity resonator]].  This consists of a short length of waveguide blocked at both ends.  Waves trapped inside the resonator are reflected back and forth between the two ends.  A given geometry of cavity will [[resonance|resonate]] at a characteristic frequency.  The resonance effect can be used to selectively pass certain frequencies.  Their use in a filter structure requires that some of the wave is allowed to pass out of one cavity into another through a coupling structure.  However, if the opening in the resonator is kept small then a valid design approach is to design the cavity as if it were completely closed and errors will be minimal.  A number of different coupling mechanisms are used in different classes of filter.<ref>{{multiref|Hunter, pages 209–210|Matthaei ''et al.'', page 243}}</ref>
 
The nomenclature for modes in a cavity introduces a third index, for example TE<sub>011</sub>.  The first two indices describe the wave travelling up and down the length of the cavity, that is, they are the transverse mode numbers as for modes in a waveguide.  The third index describes the [[longitudinal mode]] caused by the [[Interference (wave propagation)|interference pattern]] of the forward travelling and reflected waves.  The third index is equal to the number of half wavelengths down the length of the guide.  The most common modes used are the dominant modes: TE<sub>101</sub> in rectangular waveguide, and TE<sub>111</sub> in circular waveguide.  TE<sub>011</sub> circular mode is used where very low loss (hence high ''Q'') is required but cannot be used in a dual-mode filter because it is circularly symmetric.  Better modes for rectangular waveguide in dual-mode filters are TE<sub>103</sub> and TE<sub>105</sub>.  However, even better is the TE<sub>113</sub> circular waveguide mode which can achieve a ''Q'' of 16,000 at {{nowrap|12 GHz}}.<ref>{{multiref|Connor, pages 100–101|Levy & Cohn, page 1062}}</ref>
 
===Tuning screw===
Tuning screws are screws inserted into resonant cavities which can be adjusted externally to the waveguide.  They provide fine tuning of the [[resonant frequency]] by inserting more, or less thread into the waveguide.  Examples can be seen in the post filter of figure 1: each cavity has a tuning screw secured with [[jam nut]]s and [[thread-locking compound]].  For screws inserted only a small distance, the equivalent circuit is a shunt capacitor, increasing in value as the screw is inserted.  However, when the screw has been inserted a distance λ/4 it resonates equivalent to a series LC circuit.  Inserting it further it causes the impedance to change from capacitive to inductive, that is, the arithmetic sign changes.<ref>Montgomery ''et al.'', pages 168–169</ref>
 
===Iris===
[[File:Iris lumped equivalents.svg|thumb|alt=diagram|'''Figure 5.'''  Some waveguide iris geometries and their lumped element equivalent circuits]]
An iris is a thin metal plate across the waveguide with one or more holes in it.  It is used to couple together two lengths of waveguide and is a means of introducing a discontinuity.  Some of the possible geometries of irises are shown in figure 5.  An iris which reduces the width of a rectangular waveguide has an equivalent circuit of a shunt inductance, whereas one which restricts the height is equivalent to a shunt capacitance.  An iris which restricts both directions is equivalent to a parallel [[LC circuit|LC resonant circuit]].  A series LC circuit can be formed by spacing the conducting portion of the iris away from the walls of the waveguide.  Narrowband filters frequently use irises with small holes.  These are always inductive regardless of the shape of the hole or its position on the iris.  Circular holes are simple to machine, but elongated holes, or holes in the shape of a cross, are advantageous in allowing the selection of a particular mode of coupling.<ref>{{multiref|Bagad, pages 3.41–3.44|Matthaei ''et al.'', pages 232–242|Montgomery ''et al.'', pages 162–179}}</ref>
 
Irises are a form of discontinuity and work by exciting evanescent higher modes.  Vertical edges are parallel to the electric field (E field) and excite TE modes.  The stored energy in TE modes is predominately in the magnetic field (H field), and consequently the lumped equivalent of this structure is an inductor.  Horizontal edges are parallel to the H field and excite TM modes.  In this case the stored energy is predominately in the E field and the lumped equivalent is a capacitor.<ref>Montgomery ''et al.'', pages 162–179</ref>
 
It is fairly simple to make irises that are mechanically adjustable.  A thin plate of metal can be pushed in and out of a narrow slot in the side of the waveguide.  The iris construction is sometimes chosen for this ability to make a variable component.<ref>Bagad, page 3.41</ref>
 
===Iris-coupled filter===
[[File:Iris coupled filter.svg|thumb|alt=diagram|'''Figure 6.'''  Iris-coupled filter with three irises]]
An iris-coupled filter consists of a cascade of impedance transformers in the form of waveguide resonant cavities coupled together by irises.<ref name="Hunterpage" />  In high power applications capacitive irises are avoided.  The reduction in height of the waveguide (the direction of the E field) causes the electric field strength across the gap to increase and arcing (or dielectric breakdown if the waveguide is filled with an insulator) will occur at a lower power than it would otherwise.<ref>Montgomery ''et al.'', page 167</ref>
 
===Post filter===
[[File:Post filter.svg|thumb|alt=diagram|'''Figure 7.'''  Post filter with three rows of posts]]
Posts are conducting bars, usually circular, fixed internally across the height of the waveguide and are another means of introducing a discontinuity.  A thin post has an equivalent circuit of a shunt inductor.  A row of posts can be viewed as a form of inductive iris.<ref>{{multiref|Bagad, pages 3.41–3.44|Hunter, pages 220–222|Matthaei ''et al.'', pages 453–454}}</ref>
 
A post filter consists of several rows of posts across the width of the waveguide which separate the waveguide into resonant cavities as shown in figure 7.  Differing numbers of posts can be used in each row to achieve varying values of inductance.  An example can be seen in figure 1.  The filter operates in the same way as the iris-coupled filter but differs in the method of construction.<ref>{{multiref|Hunter, pages 220–228|Matthaei ''et al.'', page 540}}</ref>
 
===Post-wall waveguide===
{{main|Post-wall waveguide}}
A post-wall waveguide, or substrate integrated waveguide, is a more recent format that seeks to combine the advantages of low radiation loss, high ''Q'', and high power handling of traditional hollow metal pipe waveguide with the small size and ease of manufacture of planar technologies (such as the widely used microstrip format).  It consists of an insulated substrate pierced with two rows of conducting posts which stand in for the side walls of the waveguide.  The top and bottom of the substrate are covered with conducting sheets making this a similar construction to the [[triplate]] format.  The existing manufacturing techniques of [[printed circuit board]] or [[low temperature co-fired ceramic]] can be used to make post-wall waveguide circuits.  This format naturally lends itself to waveguide post filter designs.<ref>Xuan & Kishk, pages 1–2</ref>
 
===Dual-mode filter===
A dual-mode filter is a kind of resonant cavity filter, but in this case each cavity is used to provide two resonators by employing two modes (two polarizations), so halving the volume of the filter for a given order.  This improvement in size of the filter is a major advantage in aircraft [[avionics]] and space applications.  High quality filters in these applications can require many cavities which occupy significant space.<ref>Hunter, pages 255–260</ref>
 
==Dielectric resonator filter==
[[File:Dielectric resonator waveguide filter.svg|thumb|alt=diagram|'''Figure 8.'''  Dielectric resonator filter with three transverse resonators]]
 
Dielectric resonators are pieces of [[dielectric]] material inserted into the waveguide.  They are usually cylindrical since these can be made without [[machining]] but other shapes have been used.  They can be made with a hole through the centre which is used to secure them to the waveguide.  There is no field at the centre when the TE<sub>011</sub> circular mode is used so the hole has no adverse effect.  The resonators can be mounted coaxial to the waveguide, but usually they are mounted transversally across the width as shown in figure 8.  The latter arrangement allows the resonators to be tuned by inserting a screw through the wall of the waveguide into the centre hole of the resonator.<ref>{{multiref|Nalwa, page 525|Jarry & Beneat, page 10}}</ref>
 
When dielectric resonators are made from a high [[permittivity]] material, such as one of the [[barium titanate]]s, they have an important space saving advantage compared to cavity resonators.  However, they are much more prone to spurious modes.  In high-power applications, metal layers may be built into the resonators to conduct heat away since dielectric materials tend to have low [[thermal conductivity]].<ref>{{multiref|Nalwa, pages 525–526|Jarry & Beneat, page 10}}</ref>
 
The resonators can be coupled together with irises or impedance transformers.  Alternatively, they can be placed in a stub-like side-housing and coupled through a small aperture.<ref>{{multiref|Nalwa, pages 525–526|Jarry & Beneat, pages 10–12}}</ref>
 
===Insert filter===
[[File:E-plane insert filter.svg|thumb|alt=diagram|'''Figure 9.'''  Insert filter with six dielectric resonators in the E-plane.]]
In '''insert filters''' one or more metal sheets are placed longitudinally down the length of the waveguide as shown in figure 9.  These sheets have holes punched in them to form resonators.  The air dielectric gives these resonators a high ''Q''.  Several parallel inserts may be used in the same length of waveguide.  More compact resonators may be achieved with a thin sheet of dielectric material and printed metallisation instead of holes in metal sheets at the cost of a lower resonator ''Q''.<ref>Jarry & Beneat, page 12</ref>
 
===Finline filter===
'''Finline''' is a different kind of waveguide technology in which waves in a thin strip of dielectric are constrained by two strips of metallisation.  There are a number of possible topological arrangements of the dielectric and metal strips.  Finline is a variation of [[slot-waveguide]] but in the case of finline the whole structure is enclosed in a metal shield.  This has the advantage that, like hollow metal waveguide, no power is lost by radiation.  Finline filters can be made by printing a metallisation pattern on to a sheet of dielectric material and then inserting the sheet into the E-plane of a hollow metal waveguide much as is done with insert filters.  The metal waveguide forms the shield for the finline waveguide.  Resonators are formed by metallising a pattern on to the dielectric sheet.  More complex patterns than the simple insert filter of figure 9 are easily achieved because the designer does not have to consider the effect on mechanical support of removing metal.  This complexity does not add to the manufacturing costs since the number of processes needed does not change when more elements are added to the design.  Finline designs are less sensitive to manufacturing tolerances than insert filters and have wide bandwidths.<ref>{{multiref|Jarry & Beneat, page 12|Srivastava & Gupta, pages&nbsp;82–84}}</ref>
 
==Evanescent-mode filter==
It is possible to design filters that operate internally entirely in evanescent modes.  This has space saving advantages because the filter waveguide, which often forms the housing of the filter, does not need to be large enough to support propagation of the dominant mode.  Typically, an evanescent mode filter consists of a length of waveguide smaller than the waveguide feeding the input and output ports.  In some designs this may be folded to achieve a more compact filter.  Tuning screws are inserted at specific intervals along the waveguide producing equivalent lumped capacitances at those points.  In more recent designs the screws are replaced with dielectric inserts.  These capacitors resonate with the preceding length of evanescent mode waveguide which has the equivalent circuit of an inductor, thus producing a filtering action.  Energy from many different evanescent modes is stored in the field around each of these capacitive discontinuities.  However, the design is such that only the dominant mode reaches the output port; the other modes decay much more rapidly between the capacitors.<ref>{{multiref|Jarry & Beneat, pages 3–5|Golio, page 9.9}}</ref>
 
==Corrugated-waveguide filter==
[[File:corrugated filter.png|thumb|alt=diagram|'''Figure 10.'''  Corrugated waveguide filter with cutaway showing the corrugations inside]]
[[File:Corrugated filter section.svg|thumb|alt=diagram|'''Figure 11.''' Longitudinal section through a corrugated waveguide filter]]
'''Corrugated-waveguide filters''', also called '''ridged-waveguide filters''', consist of a number of ridges, or teeth, that periodically reduce the internal height of the waveguide as shown in figures 10 and 11.  They are used in applications which simultaneously require a wide passband, good passband matching, and a wide stopband.  They are essentially low-pass designs (above the usual limitation of the cutoff frequency), unlike most other forms which are usually band-pass.  The distance between teeth is much smaller than the typical λ/4 distance between elements of other filter designs.  Typically, they are designed by the image parameter method with all ridges identical, but other classes of filter such as Chebyshev can be achieved in exchange for complexity of manufacture.  In the image design method the equivalent circuit of the ridges is modelled as a cascade of [[Constant k filter#Derivation|LC half section]].  The filter operates in the dominant TE<sub>10</sub> mode, but spurious modes can be a problem when they are present.  In particular, there is little stopband attenuation of TE<sub>20</sub> and TE<sub>30</sub> modes.<ref>Matthaei ''et al.'', pages 380–390</ref>
 
===Waffle-iron filter===
{{main|waffle-iron filter}}
The waffle-iron filter is a variant of the corrugated-waveguide filter.  It has similar properties to that filter with the additional advantage that spurious TE<sub>20</sub> and TE<sub>30</sub> modes are suppressed.  In the waffle-iron filter, channels are cut through the ridges longitudinally down the filter.  This leaves a matrix of teeth protruding internally from the top and bottom surfaces of the waveguide.  This pattern of teeth resembles a [[waffle iron]], hence the name of the filter.<ref>Matthaei ''et al.'', pages 390–409</ref>
 
==Waveguide stub filter==
[[File:Waveguide stub filter.png|thumb|alt=diagram|'''Figure 12.'''  Waveguide stub filter consisting of three stub resonators]]
 
A [[Stub (electronics)|stub]] is a short length of waveguide connected to some point in the filter at one end and short-circuited at the other end.  Open-circuited stubs are also theoretically possible, but an implementation in waveguide is not practical because electromagnetic energy would be launched out of the open end of the stub, resulting in high losses.  Stubs are a kind of resonator, and the lumped element equivalent is an LC resonant circuit.  However, over a narrow band, stubs can be viewed as an impedance transformer.  The short-circuit is transformed into either an inductance or a capacitance depending on the stub length.<ref>{{multiref|Connor, pages 32–34|Radmanesh, pages 295–296}}</ref>
 
A waveguide stub filter is made by placing one or more stubs along the length of a waveguide, usually λ<sub>g</sub>/4 apart, as shown in figure 12.  The ends of the stubs are blanked off to short-circuit them.<ref>Ke Wu ''et al.'', page 612</ref>  When the short-circuited stubs are λ<sub>g</sub>/4 long the filter will be a [[band-pass filter]] and the stubs will have a lumped-element approximate equivalent circuit of  parallel resonant circuits connected in shunt across the line.  When the stubs are λ<sub>g</sub>/2 long, the filter will be a [[band-stop filter]].  In this case the lumped-element equivalent is series LC resonant circuits in shunt across the line.<ref>Matthaei ''et al.'', pages 595–596, 726</ref>
 
==Absorption filter==
'''Absorption filters''' dissipate the energy in unwanted frequencies internally as heat.  This is in contrast to a conventional filter design where the unwanted frequencies are reflected back from the input port of the filter.  Such filters are used where it is undesirable for power to be sent back towards the source.  This is the case with high power transmitters where returning power can be high enough to damage the transmitter.  An absorption filter may be used to remove transmitter [[spurious emission]]s such as [[harmonic]]s or spurious [[sideband]]s.  A design that has been in use for some time has slots cut in the walls of the feed waveguide at regular intervals.  This design is known as a '''leaky-wave filter'''.  Each slot is connected to a smaller gauge waveguide which is too small to support propagation of frequencies in the wanted band.  Thus those frequencies are unaffected by the filter.  Higher frequencies in the unwanted band, however, readily propagate along the side guides which are terminated with a matched load where the power is absorbed.  These loads are usually a wedge shaped piece of microwave absorbent material.<ref>Cristal, pages 182–183</ref>  Another, more compact, design of absorption filter uses resonators with a lossy dielectric.<ref>Minakova & Rud, page 1</ref>
 
==Filter-like devices==
There are many applications of filters whose design objectives are something other than rejection or passing of certain frequencies.  Frequently, a simple device that is intended to work over only a narrow band or just one spot frequency will not look much like a filter design.  However, a [[broadband]] design for the same item requires many more elements and the design takes on the nature of a filter.  Amongst the more common applications of this kind in waveguide are [[impedance matching]] networks, [[power dividers and directional couplers|directional couplers, power dividers, power combiners]], and [[diplexer]]s.  Other possible applications include [[multiplexer]]s, demultiplexers, [[negative-resistance amplifier]]s, and [[Bessel filter|time-delay networks]].<ref>Matthaei ''et al.'', pages 1–13</ref>
 
===Impedance matching===
[[File:Orthomode transducer.jpg|thumb|alt=photo|'''Figure 13'''.  An [[orthomode transducer]] (a variety of [[duplexer]]) incorporating stepped impedance matching]]
A simple method of impedance matching is [[stub matching]] with a single stub.  However, a single stub will only produce a perfect match at one particular frequency.  This technique is therefore only suitable for narrow band applications.  To widen the bandwidth multiple stubs may be used, and the structure then takes on the form of a stub filter.  The design proceeds as if it were a filter except that a different parameter is optimised.  In a frequency filter typically the parameter optimised is stopband rejection, passband attenuation, steepness of transition, or some compromise between these.  In a matching network the parameter optimised is the impedance match.  The function of the device does not require a restriction of bandwidth, but the designer is nevertheless forced to choose a bandwidth because of the ''structure'' of the device.<ref>{{multiref|Connor, pages 32–34|Matthaei ''et al.'', page 701}}</ref>
 
Stubs are not the only format of filter than can be used.  In principle, any filter structure could be applied to impedance matching, but some will result in more practical designs than others.  A frequent format used for impedance matching in waveguide is the stepped impedance filter.  An example can be seen in the duplexer{{glosslink|Dx|e}} pictured in figure 13.<ref>{{multiref|Das & Das, pages 131–136|Matthaei ''et al.'', Chapter 6 (pages 255–354)}}</ref>
 
===Directional couplers and power combiners===
[[File:Multi-hole waveguide coupler.png|thumb|alt=drawing|'''Figure 14.'''  A multi-hole waveguide coupler]]
Directional couplers, power splitters, and power combiners  are all essentially the same type of device, at least when implemented with [[Passivity (engineering)|passive]] components.  A directional coupler splits a small amount of power from the main line to a third port.  A more strongly coupled, but otherwise identical, device may be called a power splitter.  One that couples exactly half the power to the third port (a {{nowrap|3 dB}} coupler) is the maximum coupling achievable without reversing the functions of the ports.  Many designs of power splitter can be used in reverse, whereupon they become power combiners.<ref>Lee, page 193, 201</ref>
 
A simple form of directional coupler is two parallel transmission lines coupled together over a λ/4 length.  This design is limited because the [[electrical length]] of the coupler will only be λ/4 at one specific frequency.  Coupling will be a maximum at this frequency and fall away on either side.  Similar to the impedance matching case, this can be improved by using multiple elements, resulting in a filter-like structure.<ref>Matthaei ''et al.'', page 776</ref>  A waveguide analogue of this coupled lines approach is the [[Bethe-hole directional coupler]] in which two parallel waveguides are stacked on top of each other and a hole provided for coupling.  To produce a wideband design, multiple holes are used along the guides  as shown in figure 14 and a filter design applied.<ref>Ishii, pages 205–206, 212,213</ref>  It is not only the coupled-line design that suffers from being narrow band, all simple designs of waveguide coupler depend on frequency in some way.  For instance the [[rat-race coupler]] (which can be implemented directly in waveguide) works on a completely different principle but still relies on certain lengths being exact in terms of λ.<ref>Bagad, page 4.6</ref>
 
===Diplexers and duplexers===
 
A diplexer is a device used to combine two signals occupying different frequency bands into a single signal.  This is usually to enable two signals to be transmitted simultaneously on the same communications channel, or to allow transmitting on one frequency while receiving on another. (This specific use of a diplexer is called a duplexer.) The same device can be used to separate the signals again at the far end of the channel.  The need for filtering to separate the signals while receiving is fairly self-evident but it is also required even when combining two transmitted signals.  Without filtering, some of the power from source A will be sent towards source B instead of the combined output.  This will have the detrimental effects of losing a portion of the input power and loading source A with the [[output impedance]] of source B thus causing mismatch.  These problems could be overcome with the use of a {{nowrap|3 dB}} directional coupler, but as explained in the previous section, a wideband design requires a filter design for directional couplers as well.<ref>Maloratsky, pages 165–166</ref>
 
Two widely spaced narrowband signals can be diplexed by joining together the outputs of two appropriate band-pass filters.  Steps need to be taken to prevent the filters from coupling to each other when they are at resonance which would cause degradation of their performance.  This can be achieved by appropriate spacing.  For instance, if the filters are of the iris-coupled type then the iris nearest to the filter junction of filter A is placed λ<sub>gb</sub>/4 from the junction where λ<sub>gb</sub> is the guide wavelength in the passband of filter B.  Likewise, the nearest iris of filter B is placed λ<sub>ga</sub>/4 from the junction.  This works because when filter A is at resonance, filter B is in its stopband and only loosely coupled and vice versa.  An alternative arrangement is to have each filter joined to a main waveguide at separate junctions.  A decoupling resonator is placed λ<sub>g</sub>/4 from the junction of each filter.  This can be in the form of a short-circuited stub tuned to the resonant frequency of that filter.  This arrangement can be extended to multiplexers with any number of bands.<ref>Matthaei ''et al.'', pages 969–973</ref>
 
For diplexers dealing with contiguous passbands proper account of the [[crossover filter|crossover]] characteristics of filters needs to be considered in the design.  An especially common case of this is where the diplexer is used to split the entire spectrum into low and high bands.  Here a low-pass and a high-pass filter are used instead of band-pass filters.  The synthesis techniques used here can equally be applied to narrowband multiplexers and largely remove the need for decoupling resonators.<ref>{{multiref|Levy & Cohn, page 1065|Matthaei ''et al.'', pages 991–992}}</ref>
 
===Directional filters===
[[File:Waveguide directional filter.png|thumb|upright|alt=diagram|'''Figure 15'''.  A waveguide directional filter cut away to show the circular waveguide irises]]
A directional filter is a device that combines the functions of a directional coupler and a diplexer.  As it is based on a directional coupler it is essentially a four-port device, but like directional couplers, port 4 is commonly permanently terminated internally.  Power entering port 1 exits port 3 after being subject to some filtering function (usually band-pass).  The remaining power exits port 2, and since no power is absorbed or reflected this will be the exact complement of the filtering function at port 2, in this case band-stop.  In reverse, power entering ports 2 and 3 is combined at port 1, but now the power from the signals rejected by the filter is absorbed in the load at port 4.  Figure 15 shows one possible waveguide implementation of a directional filter.  Two rectangular waveguides operating in the dominant TE<sub>10</sub> mode provide the four ports.  These are joined together by a circular waveguide operating in the circular TE<sub>11</sub> mode.  The circular waveguide contains an iris coupled filter with as many irises as needed to produce the required filter response.<ref>Matthaei ''et al.'', pages 843–847</ref>
 
==Glossary==
{{gloss}}
<!-- Last used index letter=x -->
{{term|2={{glossback|aperture|w}}aperture}}
{{defn|An opening in a wall of a waveguide or barrier between sections of waveguide through which electromagnetic radiation can propagate.}}
{{term|2={{glossback|Z0|a|b}}characteristic impedance}}
{{defn|[[Characteristic impedance]], symbol ''Z''<sub>0</sub>, of a waveguide for a particular mode is defined as the ratio of the transverse electric field to the transverse magnetic field of a wave travelling in one direction down the guide. The characteristic impedance for air filled waveguide is given by,
:<math> Z_0 = \left \{ \begin{matrix} Z_\mathrm f \dfrac {\lambda_\mathrm g}{\lambda} & \text{(TE mode)} \\ \\ Z_\mathrm f \dfrac {\lambda}{\lambda_\mathrm g} & \text{(TM mode)} \end{matrix} \right .</math>
where ''Z''<sub>f</sub> is the [[impedance of free space]], approximately {{nowrap|377 Ω}}, λ<sub>g</sub> is the guide wavelength, and λ is the wavelength when unrestricted by the guide.  For a dielectric filled waveguide, the expression must be divided by {{sqrt|κ}}, where κ is the dielectric constant of the material, and λ replaced by the unrestricted wavelength in the dielectric medium.  In some treatments what is called characteristic impedance here is called the wave impedance, and characteristic impedance is defined as proportional to it by some constant.<ref>{{multiref|Connor, page 7|Matthaei ''et al.'', pages 197–198|Montgomery ''et al.'', page 162}}</ref>}}
 
{{term|2={{glossback|Dx|c|d|e}}diplexer, duplexer}}
{{defn|A diplexer combines or separates two signals occupying different passbands.  A duplexer combines or splits two signals travelling in opposite directions, or of differing polarizations (which may also be in different passbands as well).}}
 
{{term|2={{glossback|E-plane|f}}E-plane}}
{{defn|The E-plane is the plane lying in the direction of the transverse electric field, that is, vertically along the guide.<ref name="Meredithpage">Meredith, page 127</ref>}}
 
{{term|2={{glossback|λg|h}}guide wavelength}}
{{defn|[[Guide wavelength]], symbol ''λ''<sub>g</sub>, is the wavelength measured longitudinally down the waveguide.  For a given frequency, λ<sub>g</sub> depends on the mode of transmission and is always longer than the wavelength of an electromagnetic wave of the same frequency in free space. λ<sub>g</sub> is related to the cutoff frequency, ''f''<sub>c</sub>, by,
:<math> \lambda_\mathrm g = \frac {\lambda} {\sqrt{1- \left ( \frac {f_\mathrm c}{f} \right )^2}} </math>
where λ is the wavelength the wave would have if unrestricted by the guide.  For guides that are filled only with air, this will be the same, for all practical purposes, as the free space wavelength for the transmitted frequency, ''f''.<ref>Connor, page 56</ref>}}
 
{{term|2={{glossback|H-plane|g}}H-plane}}
{{defn|The H-plane is the plane lying in the direction of the transverse magnetic field (''H'' being the analysis symbol for [[magnetic field strength]]), that is, horizontally along the guide.<ref name="Meredithpage" />}}
 
{{term|2={{glossback|height|i|j}}height, width}}
{{defn|Of a rectangular guide, these refer respectively to the small and large internal dimensions of its cross-section. The polarization of the E-field of the dominant mode is parallel to the height.}}
 
{{term|2={{glossback|iris|x}}iris}}
{{defn|A conducting plate fitted transversally across the waveguide with a, usually large, aperture.}}
 
{{term|2={{glossback|1term|k}}singly terminated, doubly terminated}}
{{defn|A doubly terminated filter (the normal case) is one where the generator and load, connected to the input and output ports respectively, have impedances matching the filter characteristic impedance.  A singly terminated filter has a matching load, but is driven either by a low impedance voltage source or a high impedance current source.<ref>Matthaei ''et al.'', page 104</ref>}}
 
{{term|2={{glossback|TEM|l}}TEM mode}}
{{defn|Transverse electromagnetic mode, a transmission mode where all the electric field and all the magnetic field are perpendicular to the direction of travel of the electromagnetic wave.  This is the usual mode of transmission in pairs of conductors.<ref>{{multiref|Connor, page 2|Silver, pages 203–204}}</ref>}}
 
{{term|2={{glossback|TE|m}}TE mode}}
{{defn|Transverse electric mode, one of a number of modes in which all the electric field, but not all the magnetic field, is perpendicular to the direction of travel of the electromagnetic wave.  They are designated H modes in some sources because these modes have a longitudinal magnetic component.  The first index indicates the number of half wavelengths of field across the width of the waveguide, and the second index indicates the number of half wavelengths across the height.  Properly, the indices should be separated with a comma, but usually they are run together, as mode numbers in double figures rarely need to be considered.  Some modes specifically mentioned in this article are listed below.  All modes are for rectangular waveguide unless otherwise stated.<ref>Connor, pages 52–54</ref>}}
{{subterm|2={{glossback|TE01|s}}TE<sub>01</sub> mode}}
{{subdefn|A mode with one half-wave of electric field across the height of the guide and uniform electric field (zero half-waves) across the width of the guide.}}
{{subterm|2={{glossback|TE10|q}}TE<sub>10</sub> mode}}
{{subdefn|A mode with one half-wave of electric field across the width of the guide and uniform electric field across the height of the guide.}}
{{subterm|2={{glossback|TE20|r}}TE<sub>20</sub> mode}}
{{subdefn|A mode with two half-waves of electric field across the width of the guide and uniform electric field across the height of the guide.}}
{{subterm|2={{glossback|TE11c|u}}TE<sub>11</sub> circular mode}}
{{subdefn|A mode with one full-wave of electric field around the circumference of the guide and one half-wave of electric field along a radius.}}
 
{{term|2={{glossback|TM|n}}TM mode}}
{{defn|Transverse magnetic mode, one of a number of modes in which all the magnetic field, but not all the electric field, is perpendicular to the direction of travel of the electromagnetic wave.  They are designated E modes in some sources because these modes have a longitudinal electric component.  See TE mode for a description of the meaning of the indices.  Some modes specifically mentioned in this article are:}}
{{subterm|2={{glossback|TM11|t}}TM<sub>11</sub> mode}}
{{subdefn|A mode with one half-wave of magnetic field across the width of the guide and one half-wave of magnetic field across the height of the guide.  This is the lowest TM mode, since TM<sub>''m''0</sub> modes cannot exist.<ref>Connor, page 60</ref>}}
{{subterm|2={{glossback|TM01c|v}}TM<sub>01</sub> circular mode}}
{{subdefn|A mode with uniform magnetic field around the circumference of the guide and one half-wave of magnetic field along a radius.}}
 
{{term|2={{glossback|tl|o|p}}transmission line}}
{{defn|A transmission line is a signal transmission medium consisting of a pair of electrical conductors separated from each other, or one conductor and a common return path.  In some treatments waveguides are considered to be within the class of transmission lines, with which they have much in common.  In this article waveguides are not included so that the two types of medium can more easily be distinguished and referred.}}
{{glossend}}
 
==Notes==
{{reflist|group=note}}
 
==References==
{{Reflist|colwidth=23em}}
 
==Bibliography==
*Bagad, V. S., ''Microwave Engineering'', Technical Publications Pune, 2009 ISBN 81-8431-360-8.
*Belov, Leonid A.; Smolskiy, Sergey M.; Kochemasov, Victor N., ''Handbook of RF, Microwave, and Millimeter-wave Components'', Artech House, 2012 ISBN 1-60807-209-6.
*Bowen, Edward George, ''A Textbook of Radar'', Cambridge University Press, 1954 {{OCLC|216292853}}.
*Bray, John, ''Innovation and the Communications Revolution: From the Victorian Pioneers to Broadband Internet'', IEE, 2002 ISBN 0-85296-218-5.
*Cauer, E.; Mathis W.; Pauli, R., [http://www.cs.princeton.edu/courses/archive/fall03/cs323/links/cauer.pdf "Life and Work of Wilhelm Cauer (1900 – 1945)"], ''Proceedings of the Fourteenth International Symposium of Mathematical Theory of Networks and Systems (MTNS2000)'', Perpignan, June, 2000 {{OCLC|65290907}}.
*Connor, F. R., ''Wave Transmission'', Edward Arnold Ltd., 1972 ISBN 0-7131-3278-7.
*Cohn, S. B., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1122473 "Microwave filters containing high-''Q'' dielectric resonators"], G-MTT Symposium Digest, pages&nbsp;49–50, 5–7 May 1965.
*Cristal, Edward G., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1125631 "Analytical solution to a waveguide leaky-wave filter structure"], ''IEEE Transactions on Microwave Theory and Techniques'', volume 11, issue 3, pages 182–190, 1963.
*Cristal, Edward G.; Matthaei, G. L., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1125756 "A technique for the design of multiplexers having contiguous channels"], ''IEEE Transactions on Microwave Theory and Techniques'', volume 12, issue 1, pages&nbsp;88–93, 1964.
*Das, Annapurna; Das, Sisir K, ''Microwave Engineering'', Tata McGraw-Hill Education, 2009 ISBN 0-07-066738-1.
*Elmore, William Cronk; Heald, Mark Aiken, ''Physics of Waves'', Courier Dover Publications, 1969 ISBN 0-486-14065-2.
*Eskelinen, Harri; Eskelinen, Pekka, ''Microwave Component Mechanics'', Artech House, 2003 ISBN 1-58053-589-5.
*Fano, R. M.; Lawson, A. W., "Design of microwave filters", chapter 10 of Ragan, G. L. (ed.), ''Microwave Transmission Circuits'', McGraw-Hill, 1948 {{OCLC|2205252}}.
*Gibilisco, Stan; Sclater, Neil, ''Encyclopedia of Electronics'', Tab Professional and Reference Books, 1990 ISBN 0-8306-3389-8.
*Golio, Mike, ''Commercial Wireless Circuits and Components Handbook'', CRC Press, 2002 ISBN 1-4200-3996-2.
*Griffin, E. L.; Young, F. A., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?arnumber=1123783 "A comparison of four overmoded canonical narrow bandpass filters at 12 GHz"], ''Microwave Symposium Digest, 1978 IEEE-MTT-S International'', pages&nbsp;47–49.
*Gusmano, G.; Bianco, A.; Viticoli, M.; Kaciulis, S.; Mattogno, G.; Pandolfi, L., [http://onlinelibrary.wiley.com/doi/10.1002/sia.1389/abstract "Study of Zr<sub>1−x</sub>Sn<sub>x</sub>TiO<sub>4</sub> thin films prepared by a polymeric precursor route"], ''Surface and Interface Analysis'', volume 34, issue 1, pages&nbsp;690–693, August 2002.
*Hitchcock, R. Timothy; Patterson, Robert M., ''Radio-Frequency and ELF Electromagnetic Energies: A Handbook for Health Professionals'', John Wiley & Sons, 1995 ISBN 0-471-28454-8.
*Hunter, I. C., ''Theory and Design of Microwave Filters'', IET, 2001 ISBN 0-85296-777-2.
*Huurdeman, Anton A., ''The Worldwide History of Telecommunications'', Wiley-IEEE, 2003 ISBN 0-471-20505-2.
*Ishii, Thomas Koryu, ''Handbook of Microwave Technology: Components and devices'', Academic Press, 1995 ISBN 0-12-374696-5.
*Jarry, Pierre; Beneat, Jacques, ''Design and Realizations of Miniaturized Fractal Microwave and RF Filters'', John Wiley & Sons, 2009 ISBN 0-470-48781-X.
*Ke, Wu; Lei, Zhu; Vahldieck, Ruediger, "Microwave passive components", in Chen, Wai-Kei (ed.), ''The Electrical Engineering Handbook'', Academic Press, 2004 ISBN 0-08-047748-8.
*Lee, Thomas H., ''Planar Microwave Engineering'', pages&nbsp;585–618, Cambridge University Press, 2004 ISBN 0-521-83526-7.
*Levy, R.; Cohn, S. B., [http://ieeexplore.ieee.org/search/freesrchabstract.jsp?tp=&arnumber=1132817 "A History of microwave filter research, design, and development"], ''IEEE Transactions: Microwave Theory and Techniques'', pages 1055–1067, volume 32, issue 9, 1984.
*Mahmoud, S. F., ''Electromagnetic waveguides: Theory and Applications'', IEE, 1991 ISBN 0-86341-232-7.
*Maloratsky, Leo G., ''Integrated Microwave Front-ends with Avionics Applications'', Artech House, 2012 ISBN 1-60807-206-1.
*Mansour, R. R., "Three-dimensional cryogenic filters" in H. Weinstock, H.; Nisenoff, M., ''Microwave Superconductivity'', pages&nbsp;161–188, Springer, 2001 ISBN 1-4020-0445-1.
*Mason, W. P.; Sykes, R. A. [http://archive.org/stream/bellsystemtechni16amerrich#page/274/mode/2up "The use of coaxial and balanced transmission lines in filters and wide band transformers for high radio frequencies"], ''Bell System Technical Journal'', pages 275–302, volume 16, 1937.
*Massé, D. J.; Pucel, R. A., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1450668 "A temperature-stable bandpass filter using dielectric resonators"], ''Proceedings of the IEEE'', volume 60, issue 6, pages&nbsp;730–731, June 1972.
*Matthaei, George L.; Young, Leo; Jones, E. M. T., ''Microwave Filters, Impedance-Matching Networks, and Coupling Structures'', McGraw-Hill, 1964 {{LCCN|647937}}.
*Meier, Paul J., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1123053 "Two new integrated-circuit media with special advantages at millimeter wavelengths"], ''1972 IEEE GMTT International Microwave Symposium'', pages&nbsp;221–223, 22–24 May 1972.
*Meredith, Roger, ''Engineers' Handbook of Industrial Microwave Heating'', IET, 1998 ISBN 0-85296-916-3.
*Middleton, Wendy M.; Van Valkenburg, Mac Elwyn, ''Reference Data for Engineers: Radio, Electronics, Computers and Communications'', Newnes, 2002 ISBN 0-7506-7291-9.
*Millman, S. (ed.), ''A History of Engineering and Science in the Bell System: Communications Sciences (1925–1980)'', AT&T Bell Laboratories, 1984 ISBN 0-932764-06-1.
*Minakova, L. B.; Rud, L. A., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=4140391 "Natural-frequency approach to the synthesis of narrow-band waveguide absorption filters"], ''32nd European Microwave Conference, 2002'', 23–26 September 2002, Milan.
*Montgomery, Carol Gray; Dicke, Robert Henry; Purcell, Edward M., ''Principles of Microwave Circuits'', IEE, 1948 ISBN 0-86341-100-2.
*Nalwa, Hari Singh, ''Handbook of Low and High Dielectric Constant Materials and Their Applications'', Academic Press, 1999 ISBN 0-08-053353-1.
*Pierce, J. R., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1697958 "Paralleled-resonator filters"], Proceedings of the IRE, volume 37, pages&nbsp;152–155, February 1949.
*Radmanesh, Matthew M., ''Advanced RF and Microwave Circuit Design'', AuthorHouse, 2009 ISBN 1-4259-7244-6.
*Rhodes, J. D., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=4807418 "Direct design of symmetrical interacting bandpass channel diplexers"], ''IEE Journal on Microwaves, Optics and Acoustics'', volume 1, issue 1, pages&nbsp;34–40, September 1976.
*Rhodes, J. D.; Levy, R., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1129570& "A generalized multiplexer theory"], ''IEEE Transactions onMicrowave Theory and Techniques'', volume 27, issue 2, pages&nbsp;99–111, February 1979.
*Richards, Paul I., [http://ieeexplore.ieee.org/search/freesrchabstract.jsp?tp=&arnumber=1697618 "Resistor-transmission-line circuits"], ''Proceedings of the IRE'', volume 36, pages 217–220, February 1948.
*Russer, Peter, ''Electromagnetics, Microwave Circuits and Antenna Design for Communications Engineering'', Artech House, 2003 ISBN 1-58053-532-1.
*Sarkar, T. K.; Mailloux, Robert; Oliner, Arthur A.; Salazar-Palma, M.; Sengupta Dipak L., ''History of Wireless'', John Wiley & Sons, 2006 ISBN 0-471-78301-3.
*Schumacher, H. L., "Coax multiplexers: key to EW signal sorting", ''Microwave Systems News'', pages 89–93, August/September 1976 {{ISSN|0164-3371}}
*Silver, Samuel, ''Microwave Antenna Theory and Design'', IEE, 1949 ISBN 0-86341-017-0.
*Sorrentino, Roberto; Bianchi, Giovanni, ''Microwave and RF Engineering'', John Wiley & Sons, 2010 ISBN 0-470-66021-X.
*Srivastava, Ganesh Prasad; Gupta, Vijay Laxmi, ''Microwave Devices and Circuit Designs'', Prentice-Hall of India, 2006 ISBN 81-203-2195-2.
*Waterhouse, Rod, ''Microstrip Patch Antennas: A Designer's Guide'', Springer, 2003 ISBN 1-4020-7373-9.
*Wenzel, J. R., [http://ieeexplore.ieee.org/xpl/articleDetails.jsp?tp=&arnumber=1125922 "Application of exact synthesis methods to multichannel filter design"], ''IEEE Transactions on Microwave Theory and Techniques'', volume 13, issue 1, pages&nbsp;5–15, January 1965.
*Xuan, Hu Wu; Kishk, Ahmed A., ''Analysis and Design of Substrate Integrated Waveguide Using Efficient 2D Hybrid Method'', Morgan & Claypool, 2010 ISBN 1-59829-902-6.
*Yeh, C.; Shimabukuro, F. I., ''The Essence of Dielectric Waveguides'', Springer, 2008 ISBN 0-387-49799-4.
*Young, L., [http://ieeexplore.ieee.org/search/freesrchabstract.jsp?tp=&arnumber=1125629 "Direct-coupled cavity filters for wide and narrow bandwidths"], ''IEEE Transactions: Microwave Theory and Techniques'', volume MTT-11, pages 162–178, May 1963.
*Young, Soo Lee; Getsinger, W. J.; Sparrow, L. R., [http://ieeexplore.ieee.org/xpl/login.jsp?tp=&arnumber=1129696 "Barium tetratitanate MIC technology"], ''IEEE Transactions on Microwave Theory and Techniques'', volume 27, issue 7, pages&nbsp;655–660, July 1979.
*Zhang, Xianrong; Wang, Qingyuan; Li, Hong; Liu, Rongjun, [http://ieeexplore.ieee.org/xpl/login.jsp?tp=&arnumber=4540375 "Evanescent mode compact waveguide filter"], ''International Conference on Microwave and Millimeter Wave Technology, 2008 (ICMMT 2008)'', volume 1, pages 323–325, IEEE, 2008.
 
[[Category:Microwave technology]]
[[Category:Linear filters]]

Latest revision as of 23:27, 29 July 2014

While everyone has heard of hypnosis, most peoples ideas of what it's all about come from seeing stage or street hypnotists, seeing it uggs on sale TV or in films, or perhaps reading about it in fiction books.

This leads to a lot of myths and misconceptions about hypnosis that can lead to people being nervous, or even afraid, of seeing a hypnotherapist for help. That's why I've compiled this list of the top 10 myths about hypnosis and hypnotherapy

1. Hypnosis is magic.
Hypnosis is not magic. It is not connected in any way to the occult or the paranormal. In fact hypnosis is a naturally occurring state that you have probably experienced many times, without even being aware of it.
2. Hypnosis is going to sleep.
The truth is that, while you are in a trance, you will still know what's going uggs on sale around you, just as when you are normally awake. If you are not aware of what's going uggs on sale, at least at the subconscious level then you will be unlikely to get much benefit from the session.

3. I will forget what happened to me while I was in a trance. Typically you will have complete recall of everything that transpires while you are in a trance, unless of course you make a choice not to remember something. For example, if you regressed and experienced a painful or traumatic event from your past that you don't yet feel ready to deal with you may choose not to remember it until you are ready to deal with it.

4. I will not be in control when I am hypnotised.
You will always be in control. Hypnosis requires your total cooperation. If I, as a hypnotherapist, do anything, or suggest that you do anything, that you don't agree with, you will simply reject it, and maybe come out of the hypnotic state.
5. I could be force to give away secrets when I am in a trance.

Hypnosis is not a truth serum. You will only say what you want to say, you are in control the whole time.
6. I may get 'stuck' in a trance.
Anytime you want to come out of the trance you will be able to. While it's true that occasionally people don't want to come out of trance, it's only because they are enjoying the sensation so much and don;t want it to end. There are various ways to make sure that you come out of trance easily and safely.

In fact, even if I did nothing to 'wake you up' you would come out uggs on sale your own when you were ready, or when something happened that you needed to deal with.
7. I may be given Post Hypnotic Suggestions to do things I don't want to do. Post Hypnotic Suggestions are used in Hypnotherapy, often to directly address the problem being worked uggs on sale, or to make going into a hypnotic state in later sessions much quicker and easier.

As with any suggestions you will only accept those suggestions that you agree to and that are for your higher good. Any suggestions that you do not agree with will simply be rejected.
8. Very intelligent people can't be hypnotised.
Actually the reverse is normally true. Intelligent people normally find it easier to achieve a hypnotic state.

9. Strong willed people can't be hypnotised.
This one is only true if the person doesn't want to be hypnotised, in which case it doesn't really matter whether you are strong willed or not. If you don't want to be hypnotised then you won't be.
10. I've never experienced a hypnotic state before.
I hear this one a lot. Some people are totally convinced they have never been in a trance, when the truth is they will experience one regularly.

If you drive a car then you may remember a time when you have driven a familiar journey and when it's almost over you suddenly realise that you have no real recollection of the journey up to that point. Or maybe you have become so engrossed in a film, or book, that the rest of the world seems to fade away.

These are all examples of a naturally occurring light hypnotic state.