Microwave drill melts concrete

By Eric Smalley, Technology Research News, Anton paar

Cross a microwave oven with a nail and you get a drill that makes holes by melting the material in its path.

Researchers at Tel Aviv University in Israel have made a prototype microwave drill that can make a 2-millimeter-diameter, 2-centimeter-deep hole in concrete in about a minute.

The microwave drill is quiet, doesn't produce dust, and is less expensive to make than laser drills, said Eli Jerby, an associate professor of physical electronics at Tel Aviv University in Israel. Once the drill is optimized, it should work as fast as mechanical drills, he said. The principal behind the drill can be used to make other tools as well, he said.

The drill works on materials like ceramics, concrete and glass, which tend to absorb microwaves, but doesn't work on conductors like metals, which reflect microwaves, said Jerby. It could be used in production lines in the electronics, automobile, ceramics, glass and construction industries, and for geological applications, he said.

The drill bit serves as an antenna to focus microwave energy on a small spot under the surface of the material, and the intense microwave energy rapidly melts the material in this spot, said Jerby. Instead of spreading microwave energy over the entire chamber of a microwave oven, "we... compress the energy into a much smaller volume -- less than 1 cubic centimeter," said Jerby. The temperature within the small spot rapidly exceeds 1,500 degrees Celsius, which is hot enough to melt many materials.

This heating happens rapidly because of the thermal runaway effect, Jerby said. The effect begins when the focused microwaves, which are 122 millimeters long, produce a 1-millimeter hot spot. The intense heat modifies the properties of the material in the hot spot so that it absorbs the microwaves even more readily. This "further localizes the heating in an exponential growth, leading to an extremely high temperature in a small spot," said Jerby.

Once a hot spot is created, the microwave drill pushes into the softened material, and moves the hot spot forward, said Jerby. The bit moves through the material like a hot knife going through butter, he said.

The concept that led to the drill was the idea to produce a localized thermal runaway effect on purpose, said Jerby. The thermal runaway effect is usually undesirable, he said.

Useful applications of the effect are not limited to drilling, said Jerby. "Derivatives of the microwave drill could be used for cutting, nailing, jointing or just for local heating," he said. The researchers have used the drill to insert metal pins into hard materials, for example, said Jerby. "One can use the microwave drill to replace three successive operations -- drilling, nailing and gluing." Nailing is accomplished by leaving the bit in the material, and when the softened material cools and hardens, it glues the nail in place.


 


 


 



The main disadvantages of the device are that it could be a radiation hazard and it can produce heat stress effects like cracks, said Jerby. These make the drill appropriate only for professional workers and automatic production lines and not for do-it-yourselfers, he said.

Under normal operations, the radiation hazard posed by the drill is comparable to that of a microwave oven, Jerby added. "We [have] reached this safety level in our lab," he said.

The cost of a drill made from microwave oven components is comparable to the cost of a mechanical drill, said Jerby. However, it's more likely that commercially available microwave drills will be more expensive versions tailored for specific applications, he said.

The researchers' prototype is about 30 centimeters tall. Inside the drill is a 5 to 30 millimeter diameter cylinder encompassing the drill bit. The bits range from 0.3 to 6 millimeters in diameter.

The idea of the microwave drill appears sound and is relatively novel, said James R. Thomas, a professor of mechanical engineering at Virginia Polytechnic Institute. There are niche applications where it could be competitive with other types of drills, said Thomas. However, "it would be limited to relatively small holes and it will not be as precise as a laser drill," he said.

The researchers next steps are working on ways to prevent the drill from inducing thermal stress effects like cracks, dealing with safety issues, adapting the drill to make larger and smaller holes, studying materials-related effects produced by the drill, and developing derivative devices, said Jerby.

The microwave drill could be used in practical applications today, said Jerby. "We have working prototypes for ceramic materials, concrete and glasses," he said. "These can be industrialized within one to two years."


 



Jerby's research colleagues were Vladimir Dikhtyar, Oleg Actushev, and Uri Grosglick. They published the research in the October 18, 2002 issue of the journal Science. The research was funded by Tel Aviv University.

Timeline:   Now
Funding:   University
TRN Categories:   Materials Science and Engineering
Story Type:   News
Related Elements:  Technical paper, "The Microwave Drill," Science, October 18, 2002


 

Biological effects of electromagnetic radiation

Time to put the rumors to bed, you scared-i-cats… your cell phone, your microwave oven, and the neighborhood cell towers are not going to hurt you. Or cause another Godzilla attack. Sorry.

Disclaimer: this information is for your reference, and is not meant to stand up in court against attacks from some tricky lawyer!

Here's two relevant government bulletins on the subject:

Questions and Answers about Biological Effects and Potential Hazards of Radiofrequency Electromagnetic Waves

(by the FCC)

Assessment of Public Health Concerns Associated with Pave Paws Radar Installations
(by Massachusetts Department of Health)

Some of this info on this page was found in Introduction to Microwaves, by Fred E. Gardiol. Merci! Not bad for a French speaking Swiss!

There are people on jobs that get thousands of times higher exposure levels that the rest of us get. If microwave radiation was as deadly as some fools would have you believe, there would be a steady line of cell tower workers, radar operators, and RF lab technicians all taking an early dirt nap. There is no evidence that this is the case. Speaking of cases, how come television lawyer James R. Sokolove doesn't advertise he'll take on electromagnetic radiation liability cases? Maybe he's smarter than he looks. It would be a losing proposition.

Ionizing versus non-ionizing radiation

Microwave energy is non-ionizing electromagnetic radiation. Ionizing radiation messes up molecules, non-ionizing radiation merely heats them. You heat your body every day when you take a shower with more energy than your cell phone could ever muster. Ionizing radiation starts at ultraviolet frequencies (UV), and includes X-rays and gamma rays, as you go up the scale of deadliness. The effects of gamma rays are what changed Dr. Bruce Banner into The Amazing Hulk; if you are worried about gamma rays buy yourself a Geiger counter.

Here's a figure that we shamelessly stole from one of the government publications. Notice that ionizing radiation is higher in frequency than visible light. Microwave radiation is far below visible light; indeed, there is a huge band of infrared between the two. No one worries much about the infrared food warmer and its effect on cafeteria workers, do they?


Electromagnetic energy is carried by photons. The higher the frequency, the higher the energy in each photon. When a certain energy level is reached, the photon has enough energy to knock off electrons from molecules that it encounters. At this point it is called ionizing radiation. The critical energy level is 10 electron volts (eV). One Joule is 6.2x10E18 electron volts, so a single electron volt is immeasurably small. Here's how to calculate the energy of a photon, depending on its frequency:

E=h

h=Planck's constant = 6.626E-34 Joule-seconds

For the ISM band (2.45 GHz) where your microwave oven operates, energy of each photon is therefore 0.00001 electron volts. The power needed to ionize a molecule is one million times higher than this.

Sunlight is far higher in frequency than microwaves, it doesn't penetrate the body, so it is more dangerous at the same power level. Sunlight provides a power level of 100mW/cm2 during the summer months, mostly infrared, but with some visible and ultraviolet energy. The cause and effect of skin cancer is well known, it's the higher frequency UV light that is going to kill you. This is why sunblock that advertises "blocks UV rays" is a good thing!

Tanning beds do far more damage than microwaves. What is wrong with people, why do you need so badly to change your skin color? Please consider interracial marriage so your kids don't have to participate in this ridiculous and injurious passtime.

Microwave exposure levels

A safety factor of 10 applied to the solar radiation level has been widely adopted for RF radiation, the standard is 10 mW/cm2 maximum. This standard applies to continuous exposure; you can get whacked with higher power for short time with no permanent effects.

Microwave radiation exposure is often expressed in terms of incident power density, in mW/cm2. The following table shows the effects of exposure to certain power levels, without time limit:

Power level

Long-term effect on human body

Notes

0.01 mw/cm2

Nothing

  

0.1 mw/cm2

Nothing

  

1 mw/cm2

Nothing

  

5 mw/cms

Nothing

Accepted standard for microwave oven leakage

10 mw/cm2

Nothing

Accepted standard for maximum continuous exposure to radiated emissions (cell phones, etc.)

30 mw/cm2

You can feel heat

  

100 mw/cm2

Cataracts can be produced

Summer sunlight is at this level.

1000 mw/cm2

Pain is induced

  

5000 mw/cm2

Cooking commences

Set the timer, lasagna in five minutes!

Exposure to higher power levels has been shown to cause cataracts. How do we know this? During WWII, there were no guidelines for how much radiation a radar operator could take. Solders and sailors were exposed by radars. At a power level of 1W/cm2, pain is induced, so their medical problems were caused by radiation below that level. Thanks to these guinea pigs, we know where the limit is. If anyone has a good "radar burn" story they want to share, send it in and we'll send you a free gift!!!

Cell phones can muster 2 watts. If the antenna was right next to your head, and half the power went into your skull, maybe a power level of 100 mw/cm2 could be produced, but we doubt it. Certainly this power level will never heat up your eyeballs, unless you hold the antenna over your eye when you talk. If you suffer from cataracts, you didn't get them from Lucky Goldstar. Many modern phones pump up the power in areas of low signal quality, so if you really want to minimize your exposure, then consider 1. going outside when you are calling, 2. talking less, listening more, and 3. moving closer to the cell tower!

No study has proven that RF radiation at less than 10 mw/cm2 could cause cancer. It is possible that a cell phone were near enough to male genitals, temporary sterility could be caused. Unless you are talking on a phone it is not transmitting, so you shouldn't have to worry about proximity effects unless you are a circus contortionist. In any case, there is no shortage of people on the planet, so a little temporary sterility wouldn't be a bad thing.

Microwave ovens are allowed to leak 5 mw/cm2 at a distance of 2 inches away. Power level will drop off as the square of the distance, so 20 inches away it can be no more than 0.05 mw/cm2.

Cell towers will never put out anywhere near 10 mw/cm2 to pedestrians near by. Maximum effective radiated power (ERP) is on the order of 100 watts, but remember, ERP includes antenna gain; the actual power that is radiated is on the order of a few watts. By the time it reaches pedestrians, out at 100 meters for example, the power density is no more than 0.001 mw/cm.

Pacemakers are designed to handle 10 mw/cm2, no problemo.

 
 


 

Smith charts

http://www.microwaves101.com/encyclopedia/Smithchart.cfm#admittance

We've got our own Smith chart tutorial here, thanks to a fan from Florida, Mike Weinstein, who really knows this subject, and is a fine writer too. If anyone else wants to be a technical contributor on their favorite microwave subject, please contact us.

If you want to download a Smith chart in pdf or gif format, we have several different ones in our download area!

The Smith chart was developed by Philip Smith at Bell Telephone's Radio Research Lab during the 1930s. Be sure to check out our entry on Philip Smith in our Microwave Hall of Fame! Phil's wife still operates Analog Instruments, the company that sells the chart. Their snail mail address is:

Analog Instruments Company,
P.O. Box 950,
New Providence, NJ 07974,
(908) 464-4214.

A clickable index to our growing Smith chart page:

What's a Smith Chart?

Impedance, admittance

Which way is up and where's that short circuit?

"Yes sir!" and please don't flip me!

Normalization

How much is that stub worth?

Single-stub matching

Some Smith chart links

What's a Smith chart?

What is a Smith chart? It's really just a plot of complex reflection overlaid with an impedance and/or admittance grid referenced to a 1-ohm characteristic impedance. That's it! Transmission coefficient, which equals unity plus reflection coefficient, may also be plotted (see below). You can find books and articles describing how a Smith chart is a graphical representation of the transmission line equations and the mathematical reasons for the circles and arcs, but these things don't really matter when you need to get the job done. What matters is knowing the basics and how to use them, like always.

The Smith chart contains almost all possible impedances, real or imaginary, within one circle. All imaginary impedances from - infinity to + infinity are represented, but only positive real impedances appear on the "classic" Smith chart. Yes, it is possible to go outside the Smith chart "unity" circle, but only with an active device because this implies negative resistance.

One thing you give up when plotting reflection coefficients on a Smith chart is a direct reading of a frequency axis. Typically, plots that are done over any frequency band have markers calling out specific frequencies.

Why use a Smith chart?  It's got all those funny circles and arcs, and good ol' rectangular plots are much better for displaying things like VSWR, transmission loss, and phase, right? Perhaps sometimes a rectangular plot is better, but a Smith chart is the RF engineer's best friend! It's easy to master, and it adds an air of "analog coolness" to presentations, which will impress your friends, if not your dates! A master in the art of Smith-charting can look at a thoroughly messed up VSWR of a component or network, and synthesize two or three simple networks that will impedance-match the circuit in his head!

Impedance and admittance

A quick refresher on the basic quantities that have units of ohms or its reciprocal, Siemens (sometimes called by its former name, mhos), is helpful since many of them will be referenced below. We all think of resistance (R) as the most fundamental of these quantities, a measure of the opposition to current flow that causes a potential drop, or voltage, according to Ohms Law: V=I*R. By extension, impedance (Z) is the steady state AC term for the combined effect of both resistance and reactance (X), where Z=R+jX. (X=jwL for an inductor, and X=1/jwC for a capacitor, where w is the radian frequency or 2*pi*f.) Generally, Z is a complex quantity having a real part (resistance) and an imaginary part (reactance).

We often think in terms of impedance and its constituent quantities of resistance and reactance. These three terms represent "opposition" quantities and are a natural fit for series-connected circuits where impedances add together. However, many circuits have elements connected in parallel or "shunt" that are a natural fit for the "acceptance" quantity of admittance (Y) and its constituent quantities of conductance (G) and susceptance (B), where Y=G+jB. (B=jwC for a capacitor, and B=1/jwL for an inductor.) Admittances add together for shunt-connected circuits. Remember that Y=1/Z=1/(R+jX), so that G=1/R only if X=0, and B=-1/X only if R=0.

When working with a series-connected circuit or inserting elements in series with an existing circuit or transmission line, the resistance and reactance components are easily manipulated on the "impedance" Smith chart. Similarly, when working with a parallel-connected circuit or inserting elements in parallel with an existing circuit or transmission line, the conductance and susceptance components are easily manipulated on the "admittance" Smith chart. The "immittance" Smith chart simply has both the impedance and admittance grids on the same chart, which is useful for cascading series-connected with parallel-connected circuits.

Which way is up and where's that short circuit?

The most common orientation of the Smith chart places the resistance axis horizontally with the short circuit (SC) location at the far left. There's a good reason for this: the voltage of the reflected wave at a short circuit must cancel the voltage of the incident wave so that zero potential exists across the short circuit. In other words, the voltage reflection coefficient must be -1 or a magnitude of 1 at an angle of 180 degrees. Since angles are measured from the positive real axis and the real axis is horizontal, the short circuit location and horizontal orientation make sense. ("Voltage" is underlined above because the current reflection coefficient of a short circuit being +1 would place the short circuit location at the right end, but let's not go there.)

For an open circuit (OC), the reflected voltage is equal to and in phase with the incident voltage (reflection coefficient of +1) so that the open circuit location is on the right. In general, the reflection coefficient has a magnitude other than unity and is complex. For reasons we won't bore you with here, anywhere above the real axis is inductive (L) and anywhere below is capacitive (C).


"Yes sir!" and please don't flip me!

Can't remember which way to rotate the locus when moving along the transmission line?  Well, it's clockwise toward the generator because generals make you go like clockwork. Also keep in mind that moving "x" degrees along the line moves a point on the locus "2x" degrees on the chart because the reflected wave must transverse the round-trip distance moved (remember, it's the reflection coefficient). Alternately, you could remember that the impedance repeats itself every half wavelength along a uniform transmission line, so you must move one time around the chart to wind up at the same impedance. Of course, a physical line length has variable electrical length over a frequency band, so a fixed impedance will spread out to a locus when viewed through a connected transmission line. This is why it is always easier to obtain a wideband match when you're close to the device or discontinuity.

Many older RF engineers advocate reflecting through the origin to "convert" from impedance to admittance and vice versa. That's why you see the same axis labeled "INDUCTIVE REACTANCE OR CAPACITIVE SUSCEPTANCE" on the original Smith chart, for example. This can be confusing, you've got to do the flip, you need to remember what the grid currently represents, and SC, OC, L & C are moving targets! Why not just keep the reflection coefficient where it belongs and use the appropriate grid? We have computers, color printers, and immittance charts these days. (If you still like to do things manually and either can't deal with all those lines on an immittance chart or are color blind, use a transparency overlay and a blank piece of paper.)

Normalization

Moving along a uniform transmission line doesn't change the magnitude of the reflection coefficient or its radial distance plotted on the Smith chart. But what about when the impedance of the line changes, for example, when a quarter-wavelength transformer is used? Reflection coefficient (Gamma) is, by definition, normalized to the characteristic impedance (Z0) of the transmission line:

Gamma = (ZL-Z0) / (ZL+Z0)

where ZL is the load impedance or the impedance at the reference plane. Note that Gamma is generally complex. Likewise, the impedance (admittance) values indicated on the grid lines are normalized to the characteristic impedance (admittance) of the transmission line to which the reflection coefficient is normalized.

When Z0 changes just past the junction between two different transmission lines, so does the reflection coefficient. Determining the new impedance (admittance) is simple: multiply by the characteristic impedance (admittance) of the current line (this yields the unnormalized value), then divide by the characteristic impedance (admittance) of the new line to obtain the new renormalized value.

The new Gamma may be calculated with the formula above or graphically determined by drawing a line from the origin to the new renormalized value. This example ignores the effect of the step discontinuity encountered in physical (non-ideal) transmission lines, which typically introduces some shunt capacitance.

How much is that stub worth?

Transmission line stubs are essential for impedance matching, introducing small amounts of phase delay (in pairs to cancel reflections), biasing, etc. Are you sometimes unsure that a short-circuited stub that's less than a quarter wavelength is inductive, or whether a wide, low impedance stub in shunt with the main line has low or high Q? A smith chart can tell you these things and give you hard numbers in a jiffy.

For example, a short-circuited stub is just a short circuit seen through a length of transmission line. Place your pencil at the SC point on the chart and move clockwise toward the generator (at the other end of the stub) on the rim by an amount less than a quarter wavelength (180 degrees on the chart). This is in the inductive region; moving more than 180 degrees makes the stub input look capacitive. At exactly one-quarter wavelength, the impedance is infinite, an open circuit. You can do the same for an open-circuited stub by starting at the OC point on the chart.

The real power of the Smith chart comes into play for analysis over a frequency band. Suppose you want to know the susceptance variation of a 50-ohm short-circuited stub over a 3:1 band. This stub could be placed in shunt with the main line at the proper point to double-tune a series-resonant locus, for instance. (We'll cover double-tuning, a very powerful technique, in a future update.) Shown in the admittance chart below is a short-circuited stub that's one-eight wavelength long at the low end and thus is three-eighths wavelengths long at the high end of the 3:1 frequency band. The normalized susceptance varies from -1.0 siemens (inductive) at flow to zero (open circuit) at midband to +1.0 siemens (capacitive) at fhigh. Therefore, the unnormalized susceptance varies between ±1.0*Y0 siemens, where Y0 (=1/Z0) is the characteristic admittance of the stub. When the characteristic admittance (Y0) of the stub is the same as the main line, the normalized susceptance of the stub may be added to the normalized admittance of the load at each frequency to yield the normalized admittance of the parallel combination. When Y0 of the stub differs from that of the main line, renormalize the stub's susceptance by Y0 of the main line before adding.


Generally, the desired susceptance variation is other than ±0.02 siemens (±1.0*Y0), which a 50-ohm stub would provide in this example. Suppose a 50-ohm main line locus needs a normalized susceptance variation of only ±0.4 siemens instead of ±1.0 siemens. Achieve this simply by making the characteristic admittance of the stub equal to 0.4 times that of the main line or Y0=0.4*0.02=0.008 siemens. The stub is now a 125-ohm line (50/0.4) and its susceptance varies less over the band, so it has lower Q. Note that the unnormalized values are rarely needed, normalized values may be renormalized by the ratio of the characteristic impedances involved.

Next, consider a stub for changing the transmission phase of a main-line signal. We know that an open-circuited stub less than a quarter wavelength long retards the phase (adds phase delay), and this is readily seen on the Smith chart: Moving clockwise from the OC position, an open-circuited stub has a transmission coefficient (1 + Gamma) with a negative phase angle. Similarly, a short-circuited stub less than a quarter wavelength long will advance the phase. The following figure illustrates the phase delay of 50-ohm and 25-ohm open-circuited stubs in shunt with a 50-ohm main line. Note that the result is mismatched, which is why stubs should be added in pairs to cancel reflections. Also note that the amount of phase delay increases as the characteristic impedance of the stub decreases (a larger Y0 produces a larger unnormalized susceptance), which makes sense since a wider stub looks like a larger capacitor.


Single-stub matching

The ability to obtain a reasonable match over a frequency band depends upon the magnitude of the mismatch, the desired bandwidth, and the complexity of matching circuit. But at any one frequency any impedance mismatch can be perfectly matched to the characteristic impedance of the transmission line, as long as it's not on the rim of the chart (perfect reflection, |Gamma| = 1). And this always can be done with one stub that's less than a quarter-wavelength long. The technique is simple: move along the transmission line to rotate the mismatch to the unity resistance (conductance) circle and insert the appropriate type and length of stub in series (shunt) with the main line to move along this circle to the origin. If the far end of the stub is either a short or open circuit (or generally, any pure reactance), its input end is also a pure reactance (susceptance) so that it doesn't affect the resistance (conductance) component of the mainline impedance (admittance).

Since it's usually easier to add a stub in parallel with a transmission line, the example shown below uses an admittance chart because, at the attachment point, the resulting admittance is the sum of the stub's input susceptance and the main line admittance. First, the mismatched point is rotated around the origin until it reaches the unity conductance circle. Then, the characteristic impedance and length of the stub is chosen such that its input susceptance is equal and opposite to the main line susceptance indicated on the unity conductance circle. The example shows two cases: move toward the generator 39 degrees of line and add a short-circuited stub that provides 0.8 siemens normalized inductive susceptance, or move toward the generator 107 degrees of line and add an open-circuited stub that provides 0.8 siemens normalized capacitive susceptance.


There are an infinite number of possible solutions because, at one frequency, a stub of any characteristic impedance can provide the necessary normalized susceptance simply by adjusting its length. The differences show up when looking over a frequency band. For example, the stub's length may be increased by an integer multiple of half-wavelengths at a particular frequency and its input susceptance at this frequency will not change. But over a frequency band, the susceptance will vary considerably more than if the extra length had not been added. 

Some Smith chart links


 


 

Here are some links on Smith charts for anyone that wants additional info or needs a Smith chart clock!

http://www.sss-mag.com/smith.html

http://www.web-ee.com/primers/files/SmithCharts/smith_charts.htm

http://www.maxim-ic.com/appnotes.cfm/appnote_number/742

RFcafe.com offers an absolutely great Excel file download for plotting data on Smith charts.

Transmission line model

Click here to go to main page on transmission lines

Click here to go to separate page on characteristic impedance

Light, phase and group velocities (rewritten for December 2007!)

Click here to go to page on transmission line attenuation

Click here to go to page on characteristic impedance

New page for December 2007! This page was recently amalgamated from material from propagation constant, phase velocity and characteristic impedance pages.

The generalized lumped-element model of a transmission line can be used to calculate characteristic impedance, phase velocity, and both parts of the propagation constant (phase and attenuation). The model uses an infinitesimally small section of a transmission line with four elements as shown below. Here the series resistance, series inductance, shunt conductance and shunt capacitance are all normalized per unit length (denoted by the "prime" notation).


By the way, the transmission line model graphic can be downloaded in a Word document along with lots of other microwave schematic symbols, just visit our download page.

Let's examine the relationships between phase constant, frequency, phase velocity and wavelength. Recall that there are 2 radians in a wavelength, therefore the relationship between phase constant and wavelength is simply:


Here's phase constant as function of frequency:


Note that the phase constant is proportional to frequency. It also turns out that the expression SQRT(L'C') is the reciprocal of phase velocity of the transmission line. Here's a separate page on that topic! But for now, remember that it is always less than or equal to the speed of light in a vacuum, which is "approximately"2.99792458E+08 meters per second.

For completeness, here's some expressions for wavelength in terms of phase constant, or frequency:


The series impedance and shunt admittance of the structure are simply:


The general form for the propagation constant starts out as this simple expression:


Propagation constant of lossless transmission line

If the transmission line is lossless, then R' and G' terms in the propagation constant equation are zero. For the lossless case the lumped model reduces to:


If R' and G' terms in the propagation constant equation are zero, the attenuation constant is also zero. The general equation for propagation constant is neatly simplified:


In the case of a lossless transmission line, the propagation constant is purely imaginary, and is merely the phase constant times SQRT(-1):


Propagation constant of low-loss transmission line

The propagation constant equation does not easily separate into real and imaginary parts for and in the case where R' and G' are non-zero terms. But significant approximations can be made for "low-loss" transmission lines. For these approximations to hold, these conditions must be met:


What does low-loss mean here? Let's assume that the ratios in the above relations are held to 10%. We made a calculation for PTFE coax, 50 ohms, at 10 GHz. At the conditions described, R'=1500 ohms/meter, and G'=0.6 Siemens/meter. Both would result in losses of 130 dB/meter (or 0.13 dB/mm). This is a very lossy cable by lab standards. Although this is slightly oranges/apples comparison the condition is even lossier than what you might measure on a MMIC transmission line at X-band. Note that the condition scales with frequency, W-band signals can have ten times as much loss and still meet the condition. So the approximation holds for just about any transmission line, no worries!

Now on to the propagation constant equation. Approximations are made using the first two terms of a Taylor series expansion. We refer you to Pozar's excellent book if you want to study this. Here's the separated phase and attenuation constants.


Note that the phase constant is calculated exactly the same from way from capacitance and inductance per unit length, regardless if the transmission line is lossy or not.

You get a non-zero attenuation constant if either G' or R' in the transmission line model (above) are non-zero terms (when G' and R' are zero the transmission line is lossless and =0). The approximation of the attenuation constant under these conditions is calculated as:


In microwave engineering, we tend to separate the attenuation constant into different components. The mechanisms of series resistance and shunt conductance can be separated into two independent loss expressions:


The term alpha1 is actually the metal loss in a transmission line due to the skin depth effect. The term alpha2 can be further separated into loss due to dielectric loss tangent, and loss due to substrate conductivity.

We have a separate page on transmission lines loss that deals further with the topic of splitting up the attenuation constant, check it out!

Velocity of light in a transmission line

The velocity of light in a transmission line is often called the "phase velocity". We make a distinction because phase velocity can mean something very different when we discuss waveguide.

Velocity of light can be derived from the inductance and capacitance per unit length of a transmission line. Under the normal (los-loss) conditions of:


The velocity of light in the transmission line is simply:


For a TEM transmission line (coax, stripline) with air dielectric the velocity of light reduces to the constant "c" which is the velocity of light in a vacuum (2.997E8 maters/second).

Transmission line characteristic impedance

The general expression that defines characteristic impedance is:


Note that in its general form, characteristic impedance can be a complex number. Also note that it only becomes complex if either R' or G' are non-zero, which will give you a headache if you think about it too long. In practice we try to achieve nearly lossless transmission lines. For a low-loss transmission line, the following relationships will occur:


Then for all practical purposes we can ignore the contributions of R' and G' from the equation and end up with a nice scalar quantity for characteristic impedance. For lossless (or near loss-less) transmission lines the characteristic impedance equatin reduces to:


What are L' and C' to the lay person? L' is the tendency of a transmission line to oppose a change in current, while C' is the tendency of a transmission line to oppose a change in voltage. Characteristic impedance is a measure of the balance between the two. How do we calculate L' and C'? that depends on what the transmission line is. For example our page on coax give the coax equations.

Relationship of L' and C' to Z0 and VP

There are many situations where you need to know inductance per unit length and capacitance per unit length of a transmission line. Both can be calculated from the characteristic impedance and the propagation velocity of the wave in a transmission line. The key to solving these equations is that the propagation velocity of a transmission line is a very simple function of its capacitance and inductance per unit length:


Note than when you plug in inductance in Henries/meter and capacitance in Farads/meter, you are talking about one speedy wave, limited to be less than or equal to the velocity of light which is about 3x10E8 meters/second. From this equation and that for Z0 (above) you can arrive at the following for L' and C':


Now the truth comes out... for a TEM transmission line such as stripline or coax, and for a given dielectric material (which would correspond to Keff in the above equations), the inductance and capacitance per unit length don't change when you scale the geometry up and down. So all semirigid, 50 ohm PTFE-filled cables (and PTFE-filled stripline!) will have 94.8 pF/meter (28.9 pF/foot) capacitance and 237 nH/meter (72.2 nH/foot) inductance! Let's state that as a Microwaves101 Rule of Thumb:

For coax and stripline 50 ohm transmission lines that employ PTFE dielectric (or any dielectric material with dielectric constant=2), the inductance per foot of is approximately 70 nH, and the capacitance per foot is about 30 pF.


 


 


 


 

S-parameters

http://www.microwaves101.com/encyclopedia/sparameters.cfm

Click here to go to our page on the Smith chart

Click here to go to our page on reference planes

Click here to go to our page on network analyzer measurements

Click here to go learn about our our S-parameter Utilities spreadsheet

Click here to learn some basic network theory

Updated for March 2006! Welcome to our page on S-parameters... it's getting better and better, sorry for the slow start!

When you come down to it, there are really only a few things that separate a microwave engineer from a "normal" electrical engineer: knowledge of the Smith chart, S-parameters, transmission lines including waveguide, and decibels. Thankfully, these are all simple concepts and we'll help you master them right here at Microwaves101!

History of S-parameters

S-parameters refer to the scattering matrix ("S" in S-parameters refers to scattering). The concept was first popularized around the time that Kaneyuke Kurokawa of Bell Labs wrote his 1965 IEEE article Power Waves and the Scattering Matrix. Check him out in our Microwaves101 Hall of Fame! It helped that during the 1960s, Hewlett Packard introduced the first microwave network analyzers. We'll also admit that there are several papers that predate Kurokawa's from the 1950s, one good early work was written by E. M. Matthews, Jr., of Sperry Gyroscope Company, titled The Use of Scattering Matrices in Microwave Circuits. Also, Robert Collin's textbook Field Theory of Guided Waves, published 1960, has a brief discussion on the Scattering matrix. Collin's book is extensively annotated, including an author index, which reads like a Who's Who of electromagnetic theory for the first half of the twentieth century.

Introduction to S-parameters

Before we get into the math, let's define a few things you need to know about S-parameters.

The scattering matrix is a mathematical construct that quantifies how RF energy propagates through a multi-port network. The S-matrix is what allows us to accurately describe the properties of incredibly complicated networks as simple "black boxes". For an RF signal incident on one port, some fraction of the signal bounces back out of that port, some of it scatters and exits other ports (and is perhaps even amplified), and some of it disappears as heat or even electromagnetic radiation. The S-matrix for an N-port contains a N2 coefficients (S-parameters), each one representing a possible input-output path.

S-parameters are complex (magnitude and angle) because both the magnitude and phase of the input signal are changed by the network. Quite often we refer to the magnitude only, as it is of the most interest. Who cares how the signal phase is changed by an amplifier or attenuator? You mostly care about how much gain (or loss) you get. S-parameters are defined for a given frequency and system impedance, and vary as a function of frequency for any non-ideal network.

S-parameters refer to RF "voltage out versus voltage in" in the most basic sense. S-parameters come in a matrix, with the number of rows and columns equal to the number of ports. For the S-parameter subscripts "ij", j is the port that is excited (the input port), and "i" is the output port. Thus S11 refers to the ratio of signal that reflects from port one for a signal incident on port one. Parameters along the diagonal of the S-matrix are referred to as reflection coefficients because they only refer to what happens at a single port, while off-diagonal S-parameters are referred to as transmission coefficients, because they refer to what happens from one port to another. Here are the S-matrices for one, two and three-port networks:


Note that each S-parameter is a vector, so if actual data were presented in matrix format, a magnitude and phase angle would be presented for each Sij.

The input and output reflection coefficients of networks (such as S11 and S22) can be plotted on the Smith chart. Transmission coefficients (S21 and S12) are usually not plotted on the Smith chart.

Definition of S-parameters

S-parameters describe the response of an N-port network to voltage signals at each port. The first number in the subscript refers to the responding port, while the second number refers to the incident port. Thus S21 means the response at port 2 due to a signal at port 1. The most common "N-port" in microwaves are one-ports and two-ports, three-port network S-parameters are easy to model with software such as Agilent ADS, but the three-port S-parameter measurements are extremely difficult to perform with accuracy. Measure S-parameters are available from vendors for amplifiers, but we've never seen a vendor offer true three-port S-parameters for a even a simple SPDT switch (a three-port network).

Let's examine a two-port network. The incident voltage at each port is denoted by "a", while the voltage leaving a port is denoted by "b". Don't get all hung up on how two voltages can occur at the same node, think of them as traveling in opposite directions!


Now we can define the four S-parameters of the 2-port as:


See how the subscript neatly follows the parameters in the ratio (S11=b1/a1, etc...)? Here's the matrix algebraic representation of 2-port S-parameters:


If we want to measure S11, we inject a signal at port one and measure its reflected signal. In this case, no signal is injected into port 2, so a2=0; during all laboratory S-parameter measurements, we only inject one signal at a time. If we want to measure S21, we inject a signal at port 1, and measure the resulting signal exiting port 2. For S12 we inject a signal into port 2, and measure the signal leaving port 1, and for S22 we inject a signal at port 2 and measure its reflected signal.

Did we mention that all of the a and b measurements are vectors? It isn't always necessary to keep track of the angle of the S-parameters, but vector S-parameters are a much more powerful tool than magnitude-only S-parameters, and the math is simple enough either way.

S-parameter magnitudes are presented in one of two ways, linear magnitude or decibels (dB). Because S-parameters are a voltage ratio, the formula for decibels in this case is

Sij(dB)=20*log[Sij(magnitude)]

Remember that power ratios are expressed as 10xlog(whatever). Voltage ratios are 20xlog(whatever), because power is proportional to voltage squared.

The angle of a vector S-parameter is almost always presented in degrees (but of course, radians are possible).

Types of S-parameters

When we are talking about networks that can be described with S-parameters, we are usually talking about single-frequency networks. Receivers and mixers aren't referred to as having S-parameters, although you can certainly measure the reflection coefficients at each port and refer to these parameters as S-parameters. The trouble comes when you wish to describe the frequency-conversion properties, this is not possible using S-parameters.

Small signal S-parameters are what we are talking about 99% of the time. By small signal, we mean that the signals have only linear effects on the network, small enough so that gain compression does not take place. For passive networks, small-signal is all you have to worry about, because they act linearly at any power level.

Large signal S-parameters are more complicated. In this case, the S-matrix will vary with input signal strength. Measuring and modeling large signal S-parameters will not be described on this page (perhaps we will get into that someday)

Mixed-mode S-parameters refer to a special case of analyzing balanced circuits. We're not going to get into that either!

Pulsed S-parameters are measured on power devices so that an accurate representation is captured before the device heats up. This is a tricky measurement, and not something we're gonna tackle yet.

Other matrices

S-parameters are just one matrix that can fully describe a network. Other matrices include ABCD parameters, Y-parameters and Z-parameters. ABCD parameters are actually used "behind the scenes" in many calculations, because they are easily cascadable. By cascadable, we mean that if you want to simulate an attenuator followed by an amplifier, the S-parameter math will drive you insane, while the ABCD math involves nothing more than multiplication.

Health effects

Microwaves contain insufficient energy to directly chemically change substances by ionization, and so are an example of nonionizing radiation. The word "radiation" refers to the fact that energy can radiate, and not to the different nature and effects of different kinds of energy.
A great number of studies have been undertaken in the last two decades, most concluding they are safe. It is understood that microwave radiation at a level that causes heating of living tissue is hazardous (due to the possibility of overheating and burns) and most countries have standards limiting exposure, such as the Federal Communications Commission RF safety regulations.
Synthetic reviews of literature indicate the predominance of their safety of use.

Health effects

Microwaves contain insufficient energy to directly chemically change substances by ionization, and so are an example of nonionizing radiation. The word "radiation" refers to the fact that energy can radiate, and not to the different nature and effects of different kinds of energy.
A great number of studies have been undertaken in the last two decades, most concluding they are safe. It is understood that microwave radiation at a level that causes heating of living tissue is hazardous (due to the possibility of overheating and burns) and most countries have standards limiting exposure, such as the Federal Communications Commission RF safety regulations.
Synthetic reviews of literature indicate the predominance of their safety of use.

Microwave frequency bands

The microwave spectrum is usually defined as electromagnetic energy ranging from approximately 1 GHz to 1000 GHz in frequency, but older usage includes lower frequencies. Most common applications are within the 1 to 40 GHz range. Microwave frequency bands, as defined by the Radio Society of Great Britain (RSGB), are shown in the table below:

Microwave frequency bands Designation
Frequency range
L band
1 to 2 GHz
S band
2 to 4 GHz
C band
4 to 8 GHz
X band
8 to 12 GHz
Ku band
12 to 18 GHz
K band
18 to 26.5 GHz
Ka band
26.5 to 40 GHz
Q band
30 to 50 GHz
U band
40 to 60 GHz
V band
50 to 75 GHz
E band
60 to 90 GHz
W band
75 to 110 GHz
F band
90 to 140 GHz
D band
110 to 170 GHz (Hot)
The term P band is sometimes used for Ku Band. For other definitions see Letter Designations of Microwave Bands

Power

A microwave oven passes (non-ionizing) microwave radiation (at a frequency near 2.45 GHz) through food, causing dielectric heating by absorption of energy in the water, fats and sugar contained in the food. Microwave ovens became common kitchen appliances in Western countries in the late 1970s, following development of inexpensive cavity magnetrons.
Microwave heating is used in industrial processes for drying and curing products.
Many semiconductor processing techniques use microwaves to generate plasma for such purposes as reactive ion etching and plasma-enhanced chemical vapor deposition (PECVD).
Microwaves can be used to transmit power over long distances, and post-World War II research was done to examine possibilities. NASA worked in the 1970s and early 1980s to research the possibilities of using Solar power satellite (SPS) systems with large solar arrays that would beam power down to the Earth's surface via microwaves.
Less-than-lethal weaponry exists that uses millimeter waves to heat a thin layer of human skin to an intolerable temperature so as to make the targeted person move away. A two-second burst of the 95 GHz focused beam heats the skin to a temperature of 130 F (54 C) at a depth of 1/64th of an inch (0.4 mm). The United States Air Force and Marines are currently using this type of Active Denial System.

Remote Sensing

Radar uses microwave radiation to detect the range, speed, and other characteristics of remote objects. Development of radar was accelerated during World War II due to its great military utility. Now radar is widely used for applications such as air traffic control, navigation of ships, and speed limit enforcement.
Most radio astronomy uses microwaves.
Microwave imaging; see Photoacoustic imaging in biomedicine

Communication

Before the advent of fiber optic transmission, most long distance telephone calls were carried via microwave point-to-point links through sites like the AT&T Long Lines. Starting in the early 1950's, frequency division multiplex was used to send up to 5,400 telephone channels on each microwave radio channel, with as many as ten radio channels combined into one antenna for the hop to the next site, up to 70 km away.
Wireless LAN protocols, such as Bluetooth and the IEEE 802.11 specifications, also use microwaves in the 2.4 GHz ISM band, although 802.11a uses ISM band and U-NII frequencies in the 5 GHz range. Licensed long-range (up to about 25 km) Wireless Internet Access services can be found in many countries (but not the USA) in the 3.5–4.0 GHz range.
Metropolitan Area Networks: MAN protocols, such as WiMAX (Worldwide Interoperability for Microwave Access) based in the IEEE 802.16 specification. The IEEE 802.16 specification was designed to operate between 2 to 11 GHz. The commercial implementations are in the 2.5 GHz, 3.5 GHz and 5.8 GHz ranges.
Wide Area Mobile Broadband Wireless Access: MBWA protocols based on standards specifications such as IEEE 802.20 or ATIS/ANSI HC-SDMA (e.g. iBurst) are designed to operate between 1.6 and 2.3 GHz to give mobility and in-building penetration characteristics similar to mobile phones but with vastly greater spectral efficiency.
Cable TV and Internet access on coax cable as well as broadcast television use some of the lower microwave frequencies. Some mobile phone networks, like GSM, also use the lower microwave frequencies.
Microwave radio is used in broadcasting and telecommunication transmissions because, due to their short wavelength, highly directive antennas are smaller and therefore more practical than they would be at longer wavelengths (lower frequencies). There is also more bandwidth in the microwave spectrum than in the rest of the radio spectrum; the usable bandwidth below 300 MHz is less than 300 MHz while many GHz can be used above 300 MHz. Typically, microwaves are used in television news to transmit a signal from a remote location to a television station from a specially equipped van.

Microwave Sources

Vacuum tube based devices operate on the ballistic motion of electrons in a vacuum under the influence of controlling electric or magnetic fields, and include the magnetron, klystron, traveling wave tube (TWT), and gyrotron. These devices work in the density modulated mode, rather than the current modulated mode. This means that they work on the basis of clumps of electrons flying ballistically through them, rather than using a continuous stream.
A maser is a device similar to a laser, except that it works at microwave frequencies.

Frequency range

The microwave range includes ultra-high frequency (UHF) (0.3–3 GHz), super high frequency (SHF) (3–30 GHz), and extremely high frequency (EHF) (30–300 GHz) signals.
Above 300 GHz, the absorption of electromagnetic radiation by Earth's atmosphere is so great that it is effectively opaque, until the atmosphere becomes transparent again in the so-called infrared and optical window frequency ranges.

microwave definition

From Wikipedia, the free encyclopedia

Microwaves are electromagnetic waves with wavelengths shorter than one meter and longer than one millimeter, or frequencies between 300 megahertz and 300 gigahertz. (UHF, SHF, EHF)
Apparatuses and techniques may be described qualitatively as "microwave" when the wavelengths of signals are roughly the same as the dimensions of the equipment, so that lumped-element circuit theory is inaccurate. As a consequence, practical microwave technique tends to move away from the discrete resistors, capacitors, and inductors used with lower frequency radio waves. Instead, distributed circuit elements and transmission-line theory are more useful methods for design, analysis, and construction of microwave circuits. Open-wire and coaxial transmission lines give way to waveguides, and lumped-element tuned circuits are replaced by cavity resonators or resonant lines. Effects of reflection, polarization, scattering, diffraction, and atmospheric absorption usually associated with visible light are of practical significance in the study of microwave propagation. The same equations of electromagnetic theory apply at all frequencies.
While the name suggests a micrometer wavelength, it is better understood as indicating wavelengths very much smaller than those used in radio broadcasting. The boundaries between far infrared light, terahertz radiation, microwaves, and ultra-high-frequency radio waves are fairly arbitrary and are used variously between different fields of study. The term microwave generally refers to "alternating current signals with frequencies between 300 MHz (3×108 Hz) and 300 GHz (3×1011 Hz)."[1] (UHF, SHF, EHF)Both IEC standard 60050 and IEEE standard 100 define "microwave" frequencies starting at 1 GHz (30 cm wavelength).
Electromagnetic waves longer (lower frequency) than microwaves are called "radio waves". Electromagnetic radiation with shorter wavelengths may be called "millimeter waves", terahertz radiation or even T-rays. Definitions differ for millimeter wave band, which the IEEE defines as 110GHz to 300GHz while military radar definitions use 30-300GHz.