Antenna basics



Home | Glossary | Books | Links/Resources
EMC Testing | Environmental Testing | Vibration Testing




AMAZON multi-meters discounts AMAZON oscilloscope discounts


Before looking at the various antennas we need to look at some of the basics of antenna systems. In this section you will learn some of these basics. And while they will not make you a red-hot professional antenna engineer they will set you up well enough to understand this guide and others on amateur and hobbyist antennas. We will look at the matter of antenna radiation, antenna patterns, the symbols used to represent antennas, voltage standing wave ratio (VSWR), impedance, and various methods suitable for constructing wire antennas in the high-frequency (HF) and very high-frequency (VHF) regions of the spectrum.

ANTENNA SYSTEM SYMBOLS

ILL. 1; ILL. 2

Figures 2.1 and 2.2 show the various symbols used to represent antennas and grounds. The reason why there are so many variants is that there are differences from country to country, as well as different practices within any one country (especially between technical publishers). As for antenna symbols, I see the symbol in Ill. 1C more often in the USA, but Ill. 1B comes in for a close second. The supposedly correct symbol (endorsed by a professional society drawing standards committee) is that of Ill. 1A -- but it's only occasionally seen in the USA.



The situation for grounds is a little different because some differences reflect different forms of ground (although some of the differences also represent national or publisher differences). The ground in Ill. 2A is usually found representing a true earth ground, i.e. the wire is connected to a rod driven into the earth. The variant of Ill. 2B usually represents a chassis ground inside a piece of equipment. The symbol in Ill. 2C has two uses. One is to represent a common grounding point for different signals or different pieces of equipment. The other use is exactly the opposite: the triangle ground symbol often represents an isolated ground that has no direct electrical connection to the rest of the circuit, or with the earth. You will see this usage in medical devices. The grounds of Figures 2.2D and 2.2E are found mostly outside the USA.



From the time of Hertz and Marconi to the present, one thing has remained constant in wireless communications: radio waves travel, as if by magic, from a transmitting antenna to a receiving antenna (Ill. 3). Whether the two antennas are across the garden from each other, across continents and oceans, or on the Earth and the Moon, if there is not a transmitting antenna and at least one receiving antenna in the system then no communications can take place.

At one time, physicists believed that there must be some invisible medium for carrying the radio signal. But we now know that no such medium exists, yet radio waves travel even in outer space. Being electromagnetic waves, radio signals need no medium in order to propagate. If radio signals traveled only in the Earth's atmosphere, then we could make some guesses about a medium for carrying the wave, but space communications demonstrates that the atmosphere is not necessary (although it does affect radio signal propagation).

ILL. 3 ; ILL. 4 ; ILL. 5

Although there is no medium in which radio waves travel, it's useful to look at water waves for an analogy (even though imperfect). In Ill. 4 we see what happens when an object is dropped into a pool of water. A displacement takes place, which forms a leading wave that pushes out in concentric circles from the impact point. The situation in Ill. 4 represents a single pulse of energy, as if a transmitter fired a single burst of energy. Real transmitters send out wave trains that are analogous to cyclically bobbing the object up and down so that it goes in and out of the water (Ill. 5). The result is a continuous stream of identical waves propagating out from the 'transmitter' impact point. If another object is floating on the surface, say a cork or toy boat, then it will be perturbed as the wave passes. This is analogous to the receiver antenna.

The waves have an amplitude ('A'), which corresponds to the signal strength. They also have a wavelength (j), which corresponds to the distance traveled by the wave in one complete up-and-down cycle. In radio work, the wavelength is measured in meters (m), except in the microwave region where centimeters (cm) and millimeters (mm) make more sense. Wavelength can be measured at any pair of points on the wave that are identical: two peaks, two troughs, two zero crossings, as convenient in any specific case.

The number of cycles that pass a given point every second is the frequency of the wave. The classic measure of frequency was cycles per second (cps or c/s), but that was changed in 1960 by international consensus to the hertz (Hz), in honor of Heinrich Hertz. But since 1 Hz = 1 cps, there is no practical difference. The hertz is too small a unit for most radio work (although many of our equations are written in terms of hertz). For radio work the kilohertz (kHz) and megahertz (MHz) are used: 1 kHz = 1000 Hz, and 1 MHz = 1 000 000 Hz. Thus, a short-wave frequency of 9.75MHz is 9750 kHz and 9 750 000 Hz.

Wavelength and frequency are related to each other. The wavelength is the reciprocal of frequency, and vice versa, through the velocity constant. In free space, the velocity constant is the speed of light (c), or about 300 000 000 m/s. This is the reason why you often see '300' or its submultiples (150 and 75) in equations. When the frequency is specified in megahertz, then 300 000 000 becomes 300 for one wavelength. The half-wavelength constant is 150, and the quarter-wavelength constant is 75. The relationship is

lambda_meters = 300/F_MHz

INVERSE SQUARE LAW

When radio waves travel they become weaker by a relationship called the inverse square law. This means that the strength is inversely proportional to the square of the distance traveled (1/D^2). Ill. 6 shows how this works using the analogy of a candle. If the candle projects a distance r, all of the light energy falls onto square 'A'. At twice the distance 2r the light spreads out and covers four times the area (square 'B'). The total amount of light energy is the same, but the energy per unit of area is reduced to one-fourth of the energy that was measured at 'A.' This means that a radio signal gets weaker very rapidly as the distance from the transmitter increases, requiring ever more sensitive receivers and better antennas.

ILL. 6

THE ELECTROMAGNETIC WAVE

The electromagnetic (EM) wave propagating in space is what we know as a 'radio signal.' The EM wave is launched when an electrical current oscillates in the transmitting antenna (Ill. 7). Because moving electrical currents possess both electrical (E) and magnetic (H) fields, the electromagnetic wave launched into space has alternating E-field and H-field components. These fields are transverse (meaning they travel in the same direction) and orthogonal (meaning the E-and H-fields are at right angles to each other). When the EM wave intercepts the receiver antenna, it sets up a copy of the original oscillating currents in the antenna, and these currents are what the receiver circuitry senses.

ILL. 7; ILL. 8

The orthogonal E- and H-fields are important to the antenna designer. If you could look directly at an oncoming EM wave, you would see a plane front advancing from the transmitting antenna. If you had some magical dye that would render the E-field and H-field line of force vectors visible to the naked eye, then you would see the E-field pointing in one direction, and the H-field in a direction 90 deg. away (Ill. 8).

The polarization of the signal is the direction of the E-field vector. In Ill. 8 the polarization is vertical because the electric field vector is up and down. If the E-field vector were side-to-side, then the polarization would be horizontal. One way to tell which polarization an antenna produces when it transmits, or is most sensitive to when it receives, is to note the direction of the radiator element. If the radiator element is vertical, i.e. perpendicular to the Earth's surface, then it's vertically polarized. But if the radiator element is horizontal with respect to the Earth's surface, then it's horizontally polarized. Ill. 9 shows these relationships. In Ill. 9, two dipole receiver antennas are shown, one is vertically polarized (VD) and the other is horizontally polarized (HD). In Ill. 9A, the arriving signal is vertically polarized. Because the E-field vectors lines are vertical, they cut across more of the VD antenna than the HD, producing a considerably larger signal level. The opposite is seen in Ill. 9B. Here the E-field is horizontally polarized, so it's the HD antenna that receives the most signal.

The signal level difference can be as much as 20 dB, which represents a 10-fold decrease in signal strength if the wrong antenna is used.

DECIBELS (dB)

TABLE 1

ILL. 9

In the section above the term decibel (symbol 'dB') was introduced. The decibel is a unit of measure of the ratio of two signals: two voltages, two currents, or two powers. The equations for decibels take the logarithm of the ratio, and multiply it by a constant (10 for powers and 20 for voltages or currents). The use of decibel notation makes it possible to use ratios, such as found in gains and losses in electronic circuits, but use only addition and subtraction arithmetic. It is not necessary to be able to calculate decibels, but you should know that +dB represents a gain, and _dB represents a loss. The term '0 dB' means that the ratio of the two signals is 1:1 (neither gain nor loss). Some common ratios encountered in radio work include those listed in Table 2.1.

You can see that doubling a signal strength results in a+3 dB gain, while halving it produces a _3 dB loss. To put these figures into perspective, most S-meters on receivers use a scaling factor of 6 dB per S-unit (some use 3 dB per S-unit). Receiver designers tell us that a signal-to-noise ratio of 10 dB is necessary for 'comfortable listening,' while a signal-to-noise ratio of 3 dB is needed for barely perceptible but reliable communication for a listener who tries hard to hear what is being said.

LAW OF RECIPROCITY

Radio antennas obey a kind of law of reciprocity, i.e. they work the same on transmit as they do on receive. If an antenna has a certain gain and directivity on transmit, then that exact same pattern is seen in the receive mode.

Similarly, the feedpoint impedance, element lengths, spacings, and other issues are the same for both modes. An implication of this law is that receiver owners are able to translate the theory from discussions of transmitting antennas to their own needs (and vice versa).

Reciprocity does not necessarily mean that the same antennas are the best selection for both transmit and receive. For example, the use of gain on a transmit antenna is to increase the apparent signal power level at a distant point in a particular direction. Alternatively, some licensing authorities use antenna directivity to protect the coverage areas of relatively nearby stations that share the same or adjacent frequencies. To the receiver operator, there may be a reason to want the gain of the antenna to boost the received signal power to a useful level. There is, however, often a powerful argument for aiming an antenna such that the main sensitivity is not in the direction of the desired station. Instead, the receiver antenna owner may wish to position a null, i.e. the least sensitive aspect of the antenna, in the direction of an offending station in order to reduce its effect. The issue is, after all is said and done, the signal-to-noise ratio.

Another restriction is that there are several antennas that are fit for receive use, especially those for difficult installation situations, but are either technically unsuited or unsafe for transmitters except at the lowest of power levels. The issue might be impedance, VSWR, voltage arcing, or some other undesirable factor that occurs in transmit situations (even at moderate power levels such as 50-100W).

ANTENNA TYPES

Most antennas fall into either of two broad categories: Marconi or Hertzian.

The Marconi antennas (Ill. 10) are considered unbalanced with respect to ground because one side of the signal source (or receiver input load) is grounded. The Marconi antenna can be either vertical ( Ill. 10A) or at some angle ( Ill. 10B), including horizontal. The name derives from the types of antenna used by Guglielmo Marconi in his early radio experiments.

In his famous 1903 transatlantic communication, ol' 'Gug' used a single wire tethered to a kite aloft above the Newfoundland coast.

ILL. 10

The Hertzian antennas (Ill. 11) are balanced with respect to the ground, i.e. neither side of the oscillator or receiver load is grounded.

Dipoles fall into this category. Both Marconi and Hertzian antennas can be either vertically or horizontally polarized. However, most Marconi antennas are designed for vertical polarization and most Hertzian antennas for horizontal polarization. The long wire and vertical dipoles are obvious exceptions to that rule, however.

RECEIVER-ANTENNA INTERACTIONS

ILL. 11; ILL. 12

The antenna and the receiver (or the antenna and transmitter on the other end) form a system that must be used together. Either alone is not too useful. Ill. 12 shows an antenna connected to a receiver through a transmission line. For our practical purpose here, the receiver looks to the antenna and transmission line like a load resistor, RIN. On the positive half cycle ( Ill. 12A), the passing wave creates a current in the antenna-receiver system that flows in one direction (shown here as 'up'). When the oscillation reverses and becomes negative, the direction of current flow in the antenna-receiver reverses ( Ill. 12B). This is the mechanism by which the oscillating electromagnetic wave reproduces a signal in the input of the receiver - the signal that's then amplified and demodulated to recover whatever passes for 'intelligence' riding on the signal.

STANDING WAVE RATIO

The issue of the standing wave ratio (SWR) is of constant interest to radio enthusiasts. Some of the heat and smoke on this matter is well justified. In other cases, the perceived problems are not real.

Ill. 13 shows how the SWR comes into play in an antenna system.

In Ill. 13A, a single cycle of a signal is launched down a transmission line (it is called the 'incident' or' forward' wave). When it reaches the end of the line, if it's not totally absorbed by a load resistor or antenna, then it (or part of it) will be reflected back toward the source. This reflected wave is shown in Ill. 13B. The incident and reflected waves are both examples of travelling waves. The reflected wave represents power that's lost, and can cause other problems as well.

ILL. 13; ILL. 14

The situation in Figures 2.13A and 2.13B represent a single-cycle pulse launched down a transmission line. In a real radio system, the oscillations of the incident wave are constant ( Ill. 13C). When this situation occurs, then the reflected waves will interfere with following incident waves. At any given point, the amplitude of the wave is the algebraic sum of the interfering incident and reflected signals. The resultant caused by the interference of the incident and reflected waves is called a standing wave.

Ill. 14 shows what happens when continuous incident and reflected waves coexist on the same transmission line. In the case of Ill. 14A, the two waves coincide, with the resultant as shown. The waves begin to move apart in opposite directions in Ill. 14B, which causes the overall amplitude of the standing wave to decrease, but the location of the maximum and minimum points remain stable. The condition of 1808 out of phase between the two traveling waves results in a zero-amplitude standing wave, as shown in Ill. 14C. No current flows in this case. As the waves continue to move apart, the standing wave reappears, as shown in Ill. 14D. The final case, Ill. 14E, is the case with the traveling waves in-phase with each other (08 difference). Notice that the standing wave in Ill. 14E is the same amplitude as that of Ill. 14A, but is 1808 out of phase with it.

Ill. 14F shows the voltage along a transmission line when a VSWR higher than 1:1 is present. The antenna or load end is on the left, and the lengths along the transmission line are plotted in terms of wavelength. In the situation shown, the voltage is a minimum at the antenna end (marked '0').

It rises to a peak or 'anti-node' at a quarter wavelength, and then drops back to a minimum ('node') at the half-wavelength (j/2) point. The voltage then rises again to a peak at 3j/4, falling back to minimum at 1j. The current rises and falls in a similar manner, but the nodes and antinodes are offset from the voltage by 908 (quarter wavelength).

Note that the voltage is at a minimum and the current at a maximum at the same points that are integer multiples of a half wavelength. This situation has the effect of making it advisable to measure the VSWR at the transmitter end only when integer multiples of an electrical half wavelength of transmission line (note: the most valid measurement is made at the inter face of the transmission line and antenna feedpoint).

The plot shown in Ill. 14G is the same line system under the situation where the antenna impedance and transmission line impedance are matched. For obvious reasons, this line is said to be 'flat.'

Measures of SWR

There are several different methods for measuring the SWR of an antenna transmission line system. For example, one could measure the incident and reflected signal voltages along the line, producing a result of

VSWR = (V_1 + V_2) / (V_1 - V_2)

The method of Figures 2.14F and 2.14G can be used by comparing the maximum and minimum voltages:

VSWR = VMAX=VMIN. We can also measure the forward and reverse power levels to find the VSWR. A simple way to predict the VSWR is to compare the antenna feedpoint resistive impedance (ZL) to the transmission line characteristic impedance (Z0). The value of the VSWR is found from either

VSWR = Z0 / ZL (Z0 > ZL)

If, for example, we measure the antenna feedpoint impedance as being 25 ohms, and the antenna transmission line is 52 ohm coaxial cable, then the VSWR = 52=25 = 2:08:1.

We can find the resonant frequency of an antenna by finding the frequency at which the VSWR is a minimum. But that point may not be a VSWR of 1:1 unless the resistive component of the antenna feedpoint impedance is the same as the transmission line impedance, and the complex portion of the impedance is zero. To create this situation one might need an impedance matching device (see Section 12).

AN OLD MYTH REVIVED?

There are a number of myths that are widely held among radio communications hobbyists - and amateur radio is no less infested with some of these myths than others (CB, for example). Twenty-five years ago I worked in a CB shop in Virginia, and we kept hearing one old saw over and over again: you can 'cut your coax to reduce the VSWR to 1' (actually, they meant '1:1' but routinely called it '1'). Hoards of CBers have 'cut the coax' and watched the VSWR reduce to 1:1, so they can't be talked out of the error. I even know of one shop that kept 30 cm lengths of coaxial cable, with connectors on both ends, so they could insert them into the line at the transmitter in order to find the correct length that would reduce the VSWR to 1:1. What actually happens in that case is a measurement artifact that makes it appear to be true.

Of course, hams are superior to CBers and so don't believe that error, right? I would like to think so; but having been in both the CB and the amateur worlds, and 'Elmered' (mentored) more than a few CBers studying for amateur licenses, I have to admit that at least as many amateurs believe the 'cut the coax' error as CBers (sorry, fellows, but that's my observation).

The only really proper way to reduce the VSWR to 1:1 is to tune the antenna to resonance and then match the impedance. For a center-fed half wavelength dipole, or a bottom-fed quarter-wavelength vertical, the proper way to resonate the antenna is to adjust its length to the correct point. The formulas in the books and magazines only give approximate lengths - the real length is found from experimentation on the particular antenna after it's installed. Even commercial antennas are adjusted this way. On certain CB mobile antennas, for example, this trick is done by raising (or lowering) the radiator while watching the VSWR meter. On amateur antennas similar tuning procedures are used.

Even when the resonant point is found, the feed-point impedance may not be a good match to the transmission line. A VSWR will result in that case.

The impedance matching should be done between the far end of the transmission line (i.e. away from the receiver or rig) at the feed-point of the antenna. Antenna tuners intended for strictly coaxial cable are little more than line flatteners. They don't really 'tune' the antenna, but rather they reduce the VSWR looking into the transmission line so that the transmitter will work properly. If the antenna tuner is not a high-pass filter (as some are), then it will also provide some harmonic attenuation.

An approach used by many amateurs (including myself) is to connect an antenna-matching unit (tuner) at the output of the transmitter. For my Kenwood TS-430, I use either a Heath SA-2060A or an MFJ Differential Tuner to 'tune-out' the VSWR presented by my Hustler 4BTV and 23m of coaxial cable. But I don't even pretend to be tuning the antenna. The TS-430 is a solid-state rig, and the finals are, therefore, not terribly tolerant of VSWR, and will shut down with a high VSWR. The purpose of the antenna tuner is to reduce the VSWR seen by the transmitter - ignoring the actual antenna mismatch on the roof. The tuner also serves to reduce harmonics further, thereby helping to prevent television interference.

The best form of antenna tuner is one that both reduces the VSWR (for the benefit of the transmitter), and also resonates to the antenna frequency, preventing harmonics from getting out (a little secret is that many 'line flattener' antenna tuning units are actually variable high-pass filters, and must be used with a low-pass filter ahead of them if spurious signals are to be kept in abeyance).

Should you even worry about VSWR on a system? Or, more correctly stated, given that a 1:1 VSWR often requires a herculean effort to achieve, at what point do you declare the battle won and send the troops home? Some of the issues are:

. Transmitter heating of the transmission line due to power losses.

. Reduction of power from solid-state transmitters due to SWR shutdown circuitry.

. At kilowatt levels there may be excessive radio frequency voltage at nodes, which could lead to transmission line shorting.

. Loss of receiver sensitivity. The feedline loss is added directly to the receiver noise, so may degrade the matched noise figure of the receiver. This problem is especially severe in the VHF/UHF scanner bands.

Where these problems result in unacceptable performance, work hard to match the antenna to the feedline, and the feedline to the transmitter or receiver. As one of my correspondents said: 'The question then is to what extent does the problem impair the actually ability to use the receiver or transmitter successfully?' The same reader provided the guidelines below:

(1) The actual SWR on a feedline is 1:1 only if the load (e.g. an antenna) connected to the line is equal to the characteristic impedance of the line.

Adjusting an antenna to resonance will improve the SWR only if the impedance of the antenna at resonance is closer to the impedance of the feedline than before it was adjusted. If the impedance of the antenna at resonance is not equal to the impedance of the feedline, you will never get the SWR down to 1:1.

(2) The only possible places to measure the SWR with consistent accuracy are:

(a) at the load (b) at a distance of 0.5 electrical wavelength (accounting for the velocity factor of the line), or integer multiples thereof, from the load.

The latter is true because a half-wavelength long piece of feedline 'repeats' the impedance of whatever is connected to its far end; it's almost as if you are measuring right up at the feedpoint. As a consequence, many professional antenna installers go to pains to make the transmission line j/2 or an integer multiple of j/2. The VSWR and impedance looking into the line will reflect the situation at the feedpoint. Be aware that the velocity factor of the transmission line reduces the physical length required to make an electrical half wavelength.

(3) An easy way to check if your SWR measurements are being accurately made is as follows: Add a piece (j/8 to j/4 wavelength) of identical line to your feedline. Repeat the SWR measurement. If it's not the same, your SWR measurement is not accurate, i.e. you are not measuring the actual SWR.

(4) Know that a few things can upset SWR measurements, e.g. currents flowing on the outside of a coaxial feedline (a common situation). This can be caused by not using a BALUN when feeding a balanced antenna (e.g. a symmetric dipole) with an unbalanced (coaxial) feedline. It can also be induced by antenna currents onto a nearby coaxial feedline.

ANTENNA GAIN, DIRECTIVITY, AND RADIATION PATTERNS

Antennas don't radiate or receive uniformingly in all directions (and, indeed, you usually don't want them to!). The interrelated concepts of antenna gain and directivity are of much concern to antenna builders.

Gain refers to the fact that certain antennas cause a signal to seem to have more power than it actually can claim. The gain of the antenna can be expressed as either a multiple (e.g. a twofold increase in apparent power) or in decibel notation (e.g. a 3 dB increase). Both methods of expressing gain reflect the fact that a gain antenna seems to produce a signal that's stronger than the signal produced by some comparison antenna (e.g. a dipole). The apparent power produced by the antenna and transmitter combination is called the effective radiated power (ERP), and is the product of the transmitter power and the antenna gain. For example, if the antenna gain is 7 dB, the power is increased by a factor of five times. The ERP of a 100W transmitter is therefore 5 _ 100W = 500W.

On a receive-only antenna, the signal picked up by a gain antenna is louder than the same signal picked up on a non-gain antenna or the comparison antenna. Because of reciprocity, the gain is the same for both receive and transmit.

ILL. 15

Directivity refers to the fact that the radiation or reception direction is not constant. Certain directions are favored and others rejected. Indeed, this is how the antenna gets gain. After all, if there is an apparent increase in the signal power, there must be some phenomenon to account for it -- after all, antennas don't produce power. Ill. 15A shows how the gain and directivity are formed. Consider an isotropic radiator source located at point 'A' in Ill. 15A. An isotropic radiator is a perfectly spherical point source. The signal radiates outward from it in a spherical pattern. At some distance R we can examine the surface of the sphere. The entire output power of the transmitter is uniformly distributed over the entire surface of the sphere. But if the antenna is directive, then all of the power is focused onto a small area of the sphere with a vertical angle of _V and a horizontal area of _H. The same amount of energy exists, but now it's focused onto a smaller area, which makes it appear as if the actual power level was a lot higher in that direction.

The antenna pattern is a graphical way of showing the relative radiation in different directions (i.e. directivity). The most commonly seen pattern is the azimuthal or polar plot. That pattern is measured in the horizontal plane at all angles around the antenna, and is plotted graphically as if seen from above. It is a 'bird's eye view' of the antenna radiation characteristic. But antenna radiation patterns are actually three-dimensional, and the azimuthal plot is merely a slice of that three-dimensional pattern, as seen from but one perspective.

Figures 2.15B through 2.15F show how antenna patterns are developed.

The antenna in this case is a vertically polarized Hertzian radiator such as a half-wavelength dipole. It radiates poorly off the ends, and very well at angles perpendicular to the wire. The solid-pattern of Ill. 15B shows the three-dimensional view of what the radiation looks like. To find the azimuthal and elevation patterns for this antenna, we take slices out of the solid (Figures 2.15C and 2.15D), and plot the results (Figures 2.15E and 2.15F). Because the radiator is vertical (and perfect, I might add), it radiates equally well in all horizontal directions (see Ill. 15E), so the azimuthal or horizontal plot is termed omnidirectional. The elevation or vertical plane view ( Ill. 15F) has to take into account the minima found off the ends of the radiator, so plots as a fig ‘8.’ If the vertical radiator of Figures 2.15B-2.15F is rotated 908 to form a horizontal dipole, then the radiation patterns remain the same, but are reversed. The omnidirectional pattern of Ill. 15E becomes the elevation or vertical extent, while the fig '8' pattern of Ill. 15F is the horizontal or azimuthal extent.

ILL. 16

The patterns of Figures 2.15B-2.15F make a rather bold and unwarranted assumption: the antenna is perfect and is installed somewhere in free space. Since most antennas are not in the zone midway between the Earth and Mars, one has to account for the effects of being close to the Earth's surface. Ill. 16 shows the approximate patterns for horizontally polarized half-wavelength dipoles close to the ground. In Ill. 16 we see the azimuthal pattern. It is the figr '8,' but is pinched a bit. Notice that the maxima are perpendicular to the wire, and the minima (or 'nulls') are off the ends. One use for this type of pattern is to position the antenna so that it nulls out interfering signal or noise sources, making the overall signal-to noise ratio higher - and increasing listenability.

ILL. 17

Ill. 16B shows the elevation extent. This view is taken from a site perpendicular to the wire, and is essentially looking into one end of the three-dimensional '8' pattern. In other words, the directivity, or maxima, is in and out of the page. In free space, one would expect this pattern to be omnidirectional. But close to the Earth's surface, reflections from the surface add to the pattern, causing the pinching effect apparent in Ill. 16B. The shape shown is approximately that for a half wavelength above the ground. Other heights produce variants of this pattern.

The patterns shown in Ill. 17 are for a vertically polarized Marconi radiator, such as a quarter-wavelength, bottom-fed vertical. The azimuthal pattern, as seen from above, is shown in Ill. 17A. It is the omnidirectional pattern that one would expect. The elevation pattern in Ill. 17B represents a vertical slice taken from the 'doughnut' seen earlier. Note that the lobes are elevated, rather than being exactly on the horizon. The angle of these lobes is the angle of radiation, and affects the distance that an antenna will transmit to or receive from. A low angle of radiation is usually pre scribed for long-distance 'DX,' while a high angle of radiation produces skip to nearer regions.

Ill. 18 shows the conventions for specifying antenna patterns. The antenna is a dipole, with both relative electric field strength and relative powers plotted (both types of plot are done). The 0-2708 line is along the length of the dipole, while the 90-1808 line is perpendicular to the dipole.

The beamwidth of the antenna, which is specified in degrees, is measured as the angle between the half-power points, i.e. the points where the voltage field drops off to 70.7% of the maximum (along the 908 line), or to the 0.5 power point. These points are also called the '_3 dB' points. Recall that 3dB represents a power ratio of 1:2.

SOME ANTENNA CONSTRUCTION HINTS

ILL. 18

This section will help you think up some good ideas on how to erect the antenna. First, though, let us talk a little bit about antenna safety. It is difficult to say too much about this topic. Indeed, I mention it elsewhere in this guide as well as here. Wherever it seems prudent, I warn readers that antenna erection can be dangerous if you don't follow certain precautions.

Do it right, and the risk is mitigated and the job can be quite safe.

Safety first!

Before dealing with the radio and performance issues, let us first deal with safety matters - you don't want to be hurt either during installation, or during the next wind storm. Two problems present themselves: reliable mechanical installation and electrical safety.

Electrical safety note. Every year we read sad news in the magazines of a colleague being electrocuted while installing, or working on, an antenna. In all of these tragic cases, the antenna somehow came into contact with the electrical power lines. Keep in mind one dictum and make it an absolute: There is never a time or situation when it's safe to let an antenna contact the electrical power lines! None. Ever. BELIEVE THIS NO MATTER WHAT ANYONE SAYS TO THE CONTRARY.

This advice includes dipoles and long wires 'thrown over' supposedly insulated power mains lines, as well as antennas built from aluminum tubing. The excuse that the lines are insulated is nonsense. Old insulation crumbles on contact with even a thin wire antenna. Do not do it! The operant word is never! Consider a typical scenario involving a four-band trap vertical antenna made of tubing. It will be 5-8m tall (judging from adverts in magazines), and will be mounted on a roof, or mast, 4-10m off the ground. At my home in Virginia, a 7.6m tall trap vertical is installed atop a 4.6m telescoping television antenna mast. The total height above ground is the sum of the two heights: 7:6m þ 4:6m = 12:2m. The tip of the vertical is 12.2m above the ground. I had to select a location, on the side of my house, at which a 12.2m aluminum pole could fall safely. Although that requirement limited the selection of locations for the antenna, neither my father-in-law (who helped install the thing) nor myself was injured during the work session. Neither will a wind storm cause a shorted or downed power line if that antenna falls over.

In some jurisdictions, there might be legal limitations on antenna location. For example, some local governments in the USA have a requirement that the antenna be able to fall over and land entirely on your own property.

Before installing the antenna, check local building codes.

When installing a vertical, especially one that's not ground mounted, make sure that you have help. It takes at least two people to safely install the standard HF vertical antenna, and more may be needed for especially large models. If you are alone, then go and find some friends to lend an arm or two. Wrenched backs, smashed antenna (and house parts), and other calamities simply don't happen as often to a well-organized work party that has a sufficient number to do the job safely.

Mechanical integrity

The second issue in installing antennas is old-fashioned mechanical integrity. Two problems are seen. First, you must comply with local building regulations and inspections. Even though the courts in the USA seem to forbid local governments from prohibiting amateur radio activity (on grounds that it's Federal prerogative), local governments in the USA and elsewhere have a justifiable interest, and absolute right, to impose reason able engineering standards on the mechanical installation of radio antennas.

The second issue is that it's in your own best interest to make the installation as good as possible. View local regulations as the minimum acceptable standard, not the maximum; go one better. In other words, build the antenna installation like a brick outhouse.

Both of these mechanical integrity issues become extremely important if a problem develops. For example, suppose a wind or snow storm wrecks the antenna, plus a part of your house. The insurance company will not pay out (in most cases) if your local government requires inspections and you failed to get them done. Make sure the mechanical and /or electrical inspector (as required by US law) leaves a certificate or receipt proving that the final inspection was done. It could come in handy when disputing with the insurance company over damage.

A quality installation starts with the selection of good hardware for the installation. Any radio/television parts distributor who sells television antenna hardware will have what you need. I used Radio Shack stand-off brackets, ground pin, and a 5.8m telescoping mast. Wherever you buy, select the best-quality, strongest material that you can find. Opt for steel masts and brackets over aluminum, no matter what the salesperson behind the counter tells you. Keep in mind that, although salespeople can be knowledgeable and helpful, you, not they, are responsible for the integrity of the installation. In my own case, I found that the 5.8m mast was considerably sturdier at 4.6m than when fully up, so I opted to use less than the full length because the installation is un-guyed. Because I have never trusted the little cotter pin method of securing the mast at the slip-up height, I drilled a single hole through both bottom and slip-up segments (which telescope together), and secured the antenna mast with a 8mm stainless steel bolt.

The bolt was 'double nutted' in order to ensure that it did not come loose over time.

The television mast is set on a ground mounting pin/plate that's set into a 76 cm deep (local frost line regulations required only 71 cm) fencepost hole filled with concrete. The top end of the mast was secured to the roof overhang of the house (see Ill. 23). That overhang was beefed up with 5 _ 20 cm kiln-dried lumber that was bolted between two, 60 cm center, roof rafters. I felt it necessary to do that because the roof is only plywood, and the gutter guard is only 2:5 _ 20 cm lumber (and is old), and the soffits are aluminum. There was not enough strength to support a 12m lever arm, whipping around in a 35 knot wind.

Wind can be a terrible force, especially when acting on the 'sail area' of the antenna through a 7.5-12m lever arm. A shabby installation will tear apart in wind, causing the antenna to be damaged, damage to the house, and destruction of the installation. That is why I recommend 'brick outhouse' construction methods. Over the 33 years I have been in amateur radio, I have seen a lot of verticals toppled over. Except for a few shabby models that were so poorly built that they should not have been on the market in the first place, all of these failed installations were caused by either poor installation design or poor-quality materials.

Antenna erection methods

In this section we will look at some methods for erecting wire antennas. It is recommended that you also consult Section 3 to find out how to make connections to wire antennas and the usual fittings (end insulators, center insulators, BALUN transformers, etc.). Section 3 also deals with the type of wire used for antenna construction. The ideas in this section are not the be all-end-all discussion, but rather points of departure of representative methods. Your own innate intelligence can determine out other methods suitable for your own situation (keeping safety foremost in your plans).

Ill. 19 shows one of the most basic methods for installing a wire antenna. A mast is installed on the roof of the house, or on the side of the building close to the roof. Alternatively, the mast could be replaced by some other support on the house. The other support is a convenient nearby tree. It could also be another building or a mast erected on the ground. The antenna is supported between two end insulators, that are in turn supported by ropes connected to the mast and the tree. For a simple random length Marconi antenna, as shown here, the direction of the wire can be up, down, or horizontal, depending on the convenience of the situation (horizontal is preferred but not always easily attained). The down-lead to the receiver or antenna tuner is run through a convenient window. There are special straps and fittings for making the connection under the window (see your radio dealer).

The installation shown in Ill. 20 is a half-wavelength dipole, although it could represent a variety of doublet antennas. As with the previous case, the supports can be masts, a building, or a convenient tree. The installation is very much like that in Ill. 19. The goal in making this installation is to make the antenna horizontal, with the feed line coming away from the antenna center at a right angle for as far as is practical.

The installation in Ill. 21 is for a dipole, but it must be a small dipole (either inductively loaded or for a high frequency), or built over a very wide roof. Both end supports are on the roof, or on the walls at either end of the roof. The coaxial cable comes away at a right angle, and then passes through a window. A vertically polarized dipole is shown in Ill. 22.

ILL. 20; ILL. 21; ILL. 22

Figures 2.23A and 2.23B show methods for installing quad loops and delta loops, respectively. As above, the end supports can be a tree, a mast, or a building. It is necessary to support both the top and bottom of loops. The weight of the BALUN transformer and coaxial cable is not sufficient to keep the antenna from whipping around in the wind, and possibly doing some damage to either itself or other structures (not to mention the odd passerby).

The bottom support can be either separate ropes to the end supports ( Ill. 23A), or to a stake in the ground ( Ill. 23B). Make sure that the stake is secure against pulling out under the stress of wind blowing against the sail area of the antenna.

Now, let us talk about wire and connections.

Next: Antennas: Wire, connection, grounds, etc.

Prev:

top of page   Home



Home | Glossary | Books | Links/Resources
EMC Testing | Environmental Testing | Vibration Testing

Updated: Thursday, 2014-11-20 2:06 PST