Coll. - Antennas Fundamentals, Design, Measurement, 3rd Edn (Standard) PDF
Coll. - Antennas Fundamentals, Design, Measurement, 3rd Edn (Standard) PDF
Coll. - Antennas Fundamentals, Design, Measurement, 3rd Edn (Standard) PDF
DESIGN, MEASUREMENT
This book is dedicated to my wife Beverly Benson
Long, for our enduring love and her world-wide
work in the promotion of mental health and the
prevention of mental disorders.
ANTENNAS: FUNDAMENTALS,
DESIGN, MEASUREMENT
THIRD EDITION
LAMONT V. BLAKE
MAURICE W. LONG
No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form
or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise, except as
permitted under Sections 107 or 108 of the 1976 United Stated Copyright Act, without either the prior
written permission of the Publisher, or authorization through payment of the appropriate per-copy fee
to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400,
fax (978) 646-8600, or on the web at copyright.com. Requests to the Publisher for permission should be
addressed to the Publisher, SciTech Publishing, Inc., 911 Paverstone Drive, Suite B, Raleigh, NC 27615,
(919) 847-2434, fax (919) 847-2568, or email [email protected].
The publisher and the author make no representations or warranties with respect to the accuracy or
completeness of the contents of this work and specifically disclaim all warranties, including without
limitation warranties of fitness for a particular purpose.
This book is available at special quantity discounts to use as premiums and sales promotions, or for
use in corporate training programs. For more information and quotes, please contact the publisher.
Dedication ii
Preface xv
Chapter 1 Electromagnetic Waves 1
1.1 Characteristics of Electromagnetic Waves 2
1.1.1 Wave Velocity 2
1.1.2 Frequency and Wavelength 3
1.1.3 Space–Time Relationships 4
1.1.4 Polarization 6
1.1.5 Rays and Wavefronts 7
1.1.6 Spherical Waves and the Inverse-Square Law 8
1.1.7 Field Intensity and Power Density 10
1.1.8 Decibel (Logarithmic) Expression of Attenuation 12
1.1.9 Absorption 13
1.2 Radio-Wave Optical Principles 15
1.2.1 Refraction 15
1.2.2 Reflection 18
1.2.3 Principle of Images 21
1.2.4 Interference 22
1.2.5 Huygens’ Principle: Diffraction 25
1.3 Radiation and Reception 26
1.3.1 Requirements for Radiation 28
1.3.2 Reception and Reciprocity 30
1.4 Environmental Wave-Propagation Effects 31
1.4.1 The Earth Environment 31
1.4.2 Beyond-the-Horizon Propagation 32
1.4.3 Reflection from the Earth’s Surface 35
1.4.4 Variations in Earth’s Reflection Coefficient 39
1.4.5 Wave Nomenclature 40
1.4.6 Transionospheric Propagation 41
1.4.7 The Radio-Transmission Equation 42
References 44
Problems and Exercises 44
v
vi Contents
Appendices
A. Maxwell’s Equations 427
References 431
B. Polarization Theory 433
B.1 Elliptically Polarized Waves Resolved into Linearly
Polarized Components 433
B.2 Elliptically Polarized Waves Resolved into
Circularly Polarized Components 435
xiv Contents
Index 483
About the Author 503
Supplemental Materials (SM)
The publisher has posted a list of supplemental materials (SM) to an accompanying
website at:
www.scitechpub.com/blakelong3.htm
Within the text of the book you will find references to the specific SM sections that relate
to the material being covered. A computer icon ( ) is used in the margin to further
identify sections that refer to the SM.
SM Table of Contents
SM 1.0 Reflection Coefficients for Flat, Smooth Surfaces
SM 1.1. Reflection Coefficients-Flat, Smooth Sea
SM 1.2. Reflection Coefficients-Flat, Smooth Land
is used most often, with MATLAB being the second most popular choice. Often, original
MATLAB users, after switching to Mathcad, have expressed appreciation for their intro-
duction to Mathcad for its relative ease of use and intuitive qualities.
Because of the wide differences in student backgrounds, and after considering the
available textbooks, Blake’s Antennas, Second Edition, was selected for adoption because
of its superior readability. Due to the book’s age, the selection required preparing
and distributing materials for updating and expanding the text and adding appendices.
Consequently, the present book evolved into one that retains the benefits of Blake’s
second edition but expands the subject material suitably for a senior or graduate level
textbook.
Background Assumed
Most antenna textbooks are written for students proficient with vector calculus and begin
with the use of Maxwell’s equations in the development of antenna theory. Such books
often do not meet the needs of many students and practicing engineers who, because of
their backgrounds or personal interests, desire a more direct path for assimilating antenna
fundamentals and their connection to application topics of antenna engineering. Although
antenna theory is founded on Maxwell’s equations, understanding their concepts does
not require advanced mathematics. At the beginning of each antenna course, the revising
author (MWL) uses Appendix A, Maxwell’s Equations, to address the key “postulates”
of Maxwell and provide a brief introduction to, or review of, the essential equations.
Thus, Maxwell equations are discussed with the goal of expressing their meaning in
words. Then, the concepts of displacement current, interdependence of changing electric
and magnetic fields, and wave propagation are described, and thus Maxwell’s equations
are underscored as “the ultimate truth” but thereafter considered outside the scopes of
antenna design, performance analysis, and measurements.
Organization
This book was prepared with the intention of providing a comprehensive antenna text
that can be readily understood by persons with undergraduate educations in engineering,
science, or technology. The chapter titles follow:
Chapters 1 through 6 cover, generally, the physics and technology of antennas and
include such subjects as wave propagation, reflection, refraction, diffraction, transmission
and reception, basic radiators, antenna arrays, reflector antennas, and lenses.
Chapter 7 discusses antenna properties and analysis techniques not addressed in other
chapters. Its range of topics is wide, and includes techniques for providing wide band-
widths, multiple polarizations, low receiver noise, and extremely low sidelobes. In addi-
tion, direction-finding antennas and mechanical beam scanners are addressed. Finally
discussed are synthetic-aperture antennas, geometrical theory of diffraction (GTD),
method of moments (MoM), and fractals.
Chapter 8 treats electronically steered arrays, whereas Chapter 5 is focused on fixed
beam arrays. In other words, chapter 8 stresses array concepts specific to beam movement
made possible with fast, wide-dynamic-range digital components and cheap computer
memory, along with continued improvements in high-speed switches and phase shifters.
Chapter 9 includes a broad coverage of antenna measurement techniques and equip-
ment. Subjects include radiating near fields as well as far field patterns and pattern sta-
tistics, compact ranges, and near-field measurements. Included also is a comprehensive
treatment of antenna noise, noise temperature, noise figure, and system signal-to-noise
ratios.
There are problems at the end of each chapter, and answers to the odd numbered
problems are included in a section near the book’s end. Appendices provide technical
depth to the chapters, appropriate for a senior or first graduate level antenna course. The
appendix titles follow:
The files also include a supplemental chapter in PDF on the creation of antenna radiation
field graphics using Mathcad. It was prepared by student Aaron Loggins as one of three
project assignments in a one-semester antenna course.
Files can be downloaded from the publisher’s web page for this book:
www.scitechpub.com/blakelong3.htm
Acknowledgements
Permissions to use the contents of Antennas, 2nd Edition, by Lamont V. Blake, now
deceased, were provided by Barbara Blake, Lamont Blake’s daughter, and other Blake
family members and are gratefully appreciated. This third edition could not have been
written otherwise because it was built upon an easily read, well-written text based on a
solid technical foundation. Therefore, it could be readily expanded to provide a senior
or graduate level textbook suitable for students with widely different academic back-
grounds, including persons with limited or no computer programming experience.
Two important and closely related tasks were accomplished by Dr. Donald G. Bodnar
in connection with “Chapter 9 – Measurements.” First, he completed a technical review
of an early version of the chapter, and he then wrote Sec. 9.4, a major section titled “Near
Field Antenna Measurements.” That section is copyrighted by MI Technologies, Inc.,
Don Bodnar’s company.
Appreciation is acknowledged to Aaron Loggins for letting me use his classroom
project paper as a PDF file that discusses the creation of 3-D graphics with Mathcad.
A major and generally thankless task of pursuing a penetrating technical edit of each
chapter and appendix was accomplished by Dr. Edward B. Joy, and it was performed
with record-breaking speed. Ed found and corrected not only accidental and careless
errors, but he also underscored and made suggestions for correcting more substantive
oversights.
Special thanks are due to Dr. Anatoliy Boryssenko of the University of Massachusetts
for his expertise in checking the Mathcad files, offering helpful suggestions, and
then rewriting them into MATLAB scripts. He did so under very tight deadlines. Dr.
Boryssenko has also graciously offered additional files from his personal collection to
further enhance the data set of the publisher’s web page.
A major contribution to this book was made by Dr. Randy J. Jost of Utah State Uni-
versity, as an advisor to SciTech Publishing, by reviewing and suggesting additions to
my early book writing plans. One of those suggestions was to include the files that contain
Preface xxi
a number of antenna radiation problems and their computer solutions. Inclusion of the
CD in the Deluxe Edition that contains Mathcad, version 14, software results from the
initiative of Dudley Kay, Founder of SciTech Publishing, and the cooperation of Para-
metric Technology Corporation, the owner of Mathcad. Permission by Parametric Tech-
nology Corporation to use screenshots of computer images from Mathcad software,
included in the file on creating 3-D graphics is gratefully acknowledged.
There have been a number of persons who have made significant editorial improve-
ments and others who have simply expressed an interest in an updated edition of Lamont
Blake’s Antennas becoming available. Some of these include Gerald Oortman of Lock-
heed Martin, Marietta, Georgia; Professor Charles Bachman of the Southern Polytechnic
State University; Dr. Andrew Peterson of the School of Electrical and Computer Engi-
neering, Georgia Institute of Technology; James Gitre of Motorola; Michael Havrilla of
the Air Force Institute of Technology; and Rickey Cotton (deceased), Mark Mitchell, and
Dr. Charles Ryan (retired) of Georgia Tech Research Institute.
I thank Phyllis Hinton of Georgia Tech Research Institute, who has, over the years,
brightened my days when she sketches a figure I need or somehow helps me find my
way through the ends and outs of Microsoft Word.
Appreciation is expressed here to Dudley Kay, Susan Manning, and Robert Lawless
of SciTech Publishing who, during the preparation of this book, have demonstrated an
enthusiasm for producing quality textbooks.
Maurice W. Long
Atlanta, Georgia
[email protected]
CHAPTER 1
Electromagnetic Waves
This book begins with an elementary discussion of electromagnetic wave theory, which
is basic to an understanding of antennas. Readers who are quite familiar with the princi-
ples of electromagnetic waves may prefer to begin with the second (or third) chapter, but
even they may find this introductory chapter a handy review of the subject and a useful
reference source.
Electromagnetic waves in space are the basis of radio transmission over great distances
without direct wire connection between the transmitting and receiving points. At the
transmitting and receiving stations, radio signals exist in the form of high-frequency
alternating currents in conductors and in electronic amplifying devices. Between the
transmitter and receiver they exist as electromagnetic waves in space. Antennas, the
subject of this book, are the devices that act as go-betweens.
At the transmitting station the antenna is energized by the electrical currents generated
in the transmitter, and it converts the energy into the form of an electromagnetic field. It
“launches” the waves into space. At the receiving station the antenna captures energy
from the arriving field, and it converts the field variations into current and voltage replicas
of those at the transmitter (though of much smaller amplitude).
Current and voltage in conductors are always accompanied by electric and magnetic
fields in the adjoining region of space, and in a sense it is incorrect to speak of “convert-
ing” electrical energy from the form of current and voltage into the form of an electro-
magnetic field, and vice versa. In a practical sense, however, the distinction is made. In
the one case the fields are bound to the conductors in which the current flows; in the
other they are “free.”
The picture of electromagnetic waves presented here is considerably simplified and
necessarily leaves unanswered some questions that may be disturbing to the reader with
an inquiring mind. Often these questions can be answered, but only through the use of
rather sophisticated mathematical and physical concepts, based on James Maxwell’s
equations, which he published in 1873 (see Appendix A). Bear in mind that most of the
ideas and principles to be discussed have this background, even though it is not necessary
to look deeply into Maxwell’s equations for the purposes of this book.
1
2 Electromagnetic Waves
v
λ= (1–1)
f
As has been noted, v may have different values in different propagation media. When
the wave velocity in free space (vacuum) c is used in (1–1), the resulting value of l is
the free-space wavelength, sometimes denoted by l0.
The electromagnetic spectrum covers an enormous range of frequencies, including
cosmic-ray radiation with frequencies in excess of 1020 Hz. Radio frequencies, their
designations, and free-space wavelengths are given in Table 1-1, that is derived from the
reference IEEE Standard 211-1997. However, the radio waves that may be detected and
amplified, from a practical point of view, are from about 10 kHz to 100 GHz (IEEE
Standard 100–1992, p. 1059).
The International Telecommunications Union (ITU), an organization of the United
Nations, makes the general guidelines for the assignment and use of frequencies.
Generally, the regulations of individual nations follow those of the ITU. For the United
States, separate authorities for use by federal governmental and nonfederal governmental
frequency usages are assigned to the National Telecommunications and Information
Agency and the Federal Communications Commission, respectively. The ITU Radio
Regulations allocates the frequencies between 9 kHz and 275 GHz into frequency bands,
according to forty-two types of radio services usage—radio amateur, FM broadcast,
television, and so on—and location within three regions of the world.
A governmental operating license, with restrictions on transmit power and waveform,
is ordinarily required for radio transmit equipment, but there are some frequency bands
where operations are permitted for low-power, short-range operations. Ah Yo and Emrick
(2007) is a useful information source on frequency bands available for amateur (“ham”),
commercial, and military.
Some of the better-known frequency allocations within the United States include:
• AM (amplitude modulation) broadcast: 535–1705 kHz
• FM (frequency modulation) broadcast: 88–108 MHz
• Television in 6 MHz bandwidth numbered channels:
• 2–6, 54–60 MHz, . . . , 82–88 MHz
• 7–13, 174–180 MHz, . . . , 210–216 MHz
• 14–36, 470–476 MHz, . . . , 600–608 MHz
Note: Following the current website on U.S. frequency allocations, www.ntia.
doc.gov/osmhome/allocchrt.pdf, the region 614–698 MHz (without channel
numbers) is allocated to TV. The frequencies 608–806 MHz were previously
allocated for TV channels 37–69.
• GPS (global positioning satellite): 1227.6 MHz (military); 1575.42 MHz and
1227.6 (civilian); telemetry on 2227.5 MHz
Radar band designations following IEEE Standard 521-1984, page 8, are given in
Table 1-2. Although standard for radar, these designations are used in a broader electron-
ics community. The region 1 GHz (30 cm) to 30 GHz (1 cm) is usually called the micro-
wave region; and, contrary to Table 1-2, the region 30 GHz (10 mm) to 300 GHz (1 mm)
is usually called the millimeter wavelength region. Although millimeter-wave radars are
becoming more widely used, most radars operate within the microwave region. Some
operate at frequencies of a few MHz and a few even use infrared (8 × 1011–4 × 1014 Hz)
and visible light (4 × 1014–7.5 × 1014 Hz) frequencies.
There are differences between Tables 1-1 and 1-2. In Table 1-1 (which is for nonradar
usage), UHF is the frequency range 300 MHz–3 GHz, and the designations SHF and
EHF cover the range 3–300 GHz. Although Table 1-2 is widely used, there is not uni-
versal agreement on its letter designations and frequencies.
instant of time there is a sinusoidal variation in space along the direction of propagation,
with spatial period (wavelength) l = v/f meters, where v is the velocity of propagation
in meters per second [from (1–1)]. In terms of a cartesian coordinate system (rectangular
coordinates x, y, z), if the electric field E of the wave is represented by vectors parallel
to the x-axis and the wave is propagating in free space in a direction parallel to the z-axis,
the magnetic field H will be represented by vectors parallel to the y-axis, as shown in
Fig. 1–1. These space–time relationships for a plane wave are expressed by (1–2) and
(1–3):
2π z
Ex ( z, t ) = E0 sin ⎛⎜ 2π ft − + φ ⎞⎟ (1–2)
⎝ λ ⎠
2π z
H y ( z, t ) = H0 sin ⎛⎜ 2π ft − + φ ⎞⎟ (1–3)
⎝ λ ⎠
The notation Ex(z, t) indicates that Ex is a vector parallel to the x-axis and has a magnitude
that depends on the values of the variables z and t. The parameter E0 is the maximum
value, defined as the amplitude of the wave. Note that E0 is the value that Ex(z, t) attains
when |sin(2pft − 2pz/l + f)| = 1, which in turn will occur periodically at time intervals
of T = 1/2f at a fixed point and at z-intervals of l/2 (half-cycle and half-wavelength
intervals). Sometimes the root-mean-squared (rms) value, rather than the peak value, is
used to characterize the amplitude of a sinusoidal oscillation. The parameter f is the
initial phase angle of the wave; that is, at t = 0 and z = 0, Ex(z, t) has the value E0 sin f.
Similar statements apply to Hy(z, t). Figure 1–1 portrays these relationships
schematically.
As shown, both the electric and magnetic components of the wave are “in phase” in
space, that is, their maxima and minima occur for the same values of z. They are also in
6 Electromagnetic Waves
E0
x
Hy
H0
Ex
E(f)
FIGURE 1–1.
Spatial relationships of a plane electromagnetic wave in free space.
phase in time, at a fixed value of z. However, they are both directed at right angles to
each other and to the direction of propagation, a relationship that they always bear to
each other in free-space propagation. The designation plane wave means that the pattern
shown, although described as existing only along the z-axis, actually exists everywhere
in space, the wave vectors at any point (x, y, z) being exactly like those at the point (0,
0, z). At a fixed value of z there is no variation of the field in the x- and y-directions, that
is, in an xy-plane at the point z; hence the name plane wave. (Not all electromagnetic
waves are plane. A plane wave is an idealization never perfectly realized, but in practice
waves may often be considered locally plane, with small error and with great simplifica-
tion of mathematical description.)
The motion of the wave may be visualized by imagining that the entire set of field
vectors, not only those shown but also those at all other values of x and y, is moving in
unison in the positive z-direction at velocity c = 3 × 108 meters per second. An observer
at a fixed point would see a sinusoidal time variation of both E and H. On the other hand,
if he could somehow (magically) “freeze” the motion and take measurements of E and
H along the z-axis he would observe the pattern in Fig. 1–1.
1.1.4. Polarization
The plane wave shown in Fig. 1–1 is linearly polarized; that is, the electric vector has a
particular direction in space for all values of z, in this case the x-axis direction. The wave
is therefore said to be polarized in the x-direction. In actual space above the earth, if the
Characteristics of Electromagnetic Waves 7
electric vector is vertical or lies in a vertical plane, the wave is said to be vertically
polarized; if the E-vector lies in a horizontal plane, the wave is said to be horizontally
polarized. (It is conventional to describe polarization in terms of the E-vector.)
The initial polarization of a radio wave is determined by the antenna (and its orienta-
tion) that launches the waves into space. The polarization desired, therefore, is one of
the factors entering into antenna design. In some applications a particular polarization is
preferable; in others it makes little or no difference.
Electromagnetic waves are not always linearly polarized. In circular polarization,
from the viewpoint of a fixed observer, the electric vector appears to be rotating with a
screw motion about the z-axis (direction of propagation), making one full turn for each
rf cycle. In further analogy with a screw thread the rotation may be clockwise or coun-
terclockwise, corresponding to right-hand-circular and left-hand-circular polarizations.
A circularly polarized wave results when two linearly polarized waves are combined—
that is, if they are simultaneously launched in the same direction from the same antenna—
provided that the two linear polarizations are at right angles to each other and their phase
angles [the angle f in (1–2) and (1–3) differ by 90 degrees or p/2 radians]. The right-
hand or left-hand rotation depends on whether the phase difference is plus or minus. For
truly circular polarization it is necessary also that the two linearly polarized components
be of equal amplitude. If they are of different amplitudes, elliptical polarization results
(see Figs. B-1 through B-4 of Appendix B).
The polarization is random when there is no fixed polarization or pattern of polariza-
tion-variation that is repetitive along the z-axis, an effect present in light waves emitted
from an incandescent source (e.g., the sun or an electric light bulb). It is seldom observed
in manmade radio emissions, but such would result if two independently random sources
of radio noise (used in radio and radar military countermeasures, or “jamming”) are con-
nected to right-angle-polarized elements of a single antenna.
Linear polarization is the most commonly employed by far. One application for cir-
cular polarization is in communications between earth and space, to mitigate the effects
of polarization rotation caused by the ionosphere (see sec. 7.3.1).
Wavefront D
Wavefront C
Wavefront B
RC
RB RD
Source
Rays
d
Circular arc
tangent to
cube face
FIGURE 1–2.
Point-source wavefronts and rays in free space.
10 Electromagnetic Waves
[P will denote total power (watts), and lower case p will denote the power density of the
wave (watts per square meter).]
By similar reasoning the power density pC at the greater distance of wavefront C
will be
Pt
pC = watts per square meter (1–5)
4π RC2
This value is obviously smaller than the power density at wavefront B, since RC is greater
than RB. Thus the power density decreases as the distance from the source increases.
What is the law of this decrease? It may be found by dividing (1–4) by (1–5),
2
pB RC (1–6)
=
pC RB
which shows that the power density is inversely proportional to the square of the distance
from the source. This is the celebrated inverse-square law of radiation, observed experi-
mentally for electromagnetic waves in free space or in limited regions whose character-
istics approximate the uniformity of free space.
In deriving this result it was assumed that the source radiates isotropically—uniformly
in all directions. Actually this is not a necessary assumption, because the same result is
obtained if the source radiates preferentially in certain directions, as occurs with direc-
tional antennas. It is necessary, however, to assume that the velocity of electromagnetic
propagation is the same in all outward directions from the source. This assumption
permits the distance to the wavefront from the source at any instant to be equal in all
directions, corresponding to the geometrical definition of a sphere.
A propagation medium is called isotropic if the propagation velocity is the same in
all directions. The inverse-square law, therefore, is the result both of the spherical spread-
ing of the wavefronts in an isotropic propagation medium and of the law of conservation
of energy.
The inverse-square law is based on the distance to the wavefront from a point source
at any instant being equal in all directions, corresponding to the geometrical definition
of a sphere. However, this is true only within the far field of an antenna, because an
antenna is of finite size and therefore it is not located solely at a point. The far field exists
beyond a minimum separation distance of an antenna, which depends on antenna dimen-
sions and wavelength. (sec. 3.2.5).
power, but this quantity is usually of little interest. The average power density over an
rf cycle is ordinarily desired, and, just as in computing power in a-c circuits, it is obtained
by multiplying the effective values of E and H, equal to 1 2 times the amplitudes, or
0.707E0 and 0.707H0. Hence, the average power density p may be expressed as (1–7)
that follows:
where E0 and H0 are the amplitudes as in (1–2) and (1–3). E0 is expressed in volts per
meter, H0 in amperes per meter to give p in watts per square meter.
Just as voltage and current in circuits are related through the resistance by Ohm’s law,
the electric and magnetic intensities are related by the characteristic wave impedance of
space. In a lossless propagation medium this impedance is equal to the square root of
the ratio of its magnetic permeability m to its electric permittivity e, as given by (1–8)
that follows:
Z s = μ / ε ohms (1–8)
In a vacuum m has the value 1.26 × 10−6 henrys per meter, and e is 8.85 × 10−12 farad
per meter. (These values are customarily denoted m0 and e0.) Consequently Zs is about
377 ohms (actually 120p ohms) in free space, a value also applicable in air. Hence in
these media
E2
p= = 377 H 2 watts per square meter (1–9)
377
where E and H are the effective (rms) values, equal to 0.707E0 and 0.707H0, in volts per
meter and amperes per meter, respectively. This also means that the magnetic intensity
can be expressed by (1–10) that follows:
E
H= amperes per meter (1–10)
377
for any wave propagating in free space or air; that is, E and H are related through this
expression, and specifying one of them is equivalent to specifying both. Ordinarily,
therefore, only the electric intensity is specified.
If (1–9) is applied to the inverse-square law, the result is
E B RC
= (1–11)
EC RB
which states that the electric intensity is inversely proportional to the first power of the
distance from the source (subject to the same stipulations that apply to the inverse-square
law in its original form).
12 Electromagnetic Waves
Equations (1–6) and (1–11) are different ways of showing how the electromagnetic
wave is attenuated with increasing distance from the source. Equation (1–6) expresses
the attenuation in terms of the power-density ratio, and (1–11) expresses attenuation in
terms of the electric-intensity ratio.
2 2
p1 R2 2R
= = 1 = 4
p2 R1 R1
and
E1 R2
= =2
E2 R1
p1
10 log = 10 log 4 = 10 (0.60206 ) = 6 dB
p2
or
E1
20 log = 20 log 2 = 20 (0.30103) = 6 dB
E2
Thus the same result is obtained for the attenuation in decibels from either the power-density
or the electric-intensity ratio.
Attenuation due to the spherical spreading of the wave—that is, as expressed by
the inverse-square law—is sometimes called the space attenuation of the wave. As
can be shown from (1–6) or (1–11) by applying the definition of the decibel, the space
Characteristics of Electromagnetic Waves 13
attenuation in decibels is twenty times the logarithm of the distance ratio. As already
shown, if the distance is doubled, the wave is attenuated by 6 dB (20 log 2); if it is tripled,
the attenuation is 20 log 3, or 9.5 dB; if the distance is increased by a factor of 10, the
attenuation is 20 dB; and so on.
Since it is the distance ratio rather than the actual distance change that determines the
space attenuation, a given distance change has a greater effect at points close to a source
than it has far from the source. For example, two points differing by 1 km in distance
from the source will show a 6-dB power-density ratio if they are 1 km and 2 km from
the source, respectively; but if they are 100 km and 101 km from the source, the attenu-
ation is only 0.09 dB, a negligible amount. This fact further supports the validity of
regarding a spherical wave within a limited region of space at a considerable distance
from the source, virtually as a plane wave; for it is a property of a plane wave that the
power density and electric intensity do not change as the wave progresses.
In some antenna considerations, the ratios of the field intensities at different points in
the vicinity of the antenna may be fairly large for relatively short separations of the points
in question. However, at short distances from an antenna the fields may not obey the
inverse-square law power relationship. Therefore, it is to be underscored that the inverse-
square law relationship is applicable only to distances from an antenna that are large
enough to be within the antenna far field (sec. 3.2.5).
1.1.9. Absorption
In addition to space attenuation, which is always present (even though it may be ignored
over short distances far from the source of the waves), there may sometimes be attenua-
tion due to absorption of power by the propagation medium. This does not occur in a
vacuum, but it will occur in a medium that contains material particles that interact with
the waves. At some frequencies, for example, certain gases of the earth’s atmosphere
(oxygen and water vapor) cause absorption. This occurs slightly in the VHF region and
becomes significant over long transmission paths in the UHF region and above. Unlike
space attenuation, attenuation due to absorption does not depend on the distance from
the source but only on the total distance traveled by the wave. That is to say, attenuation
due to absorption is the same over a one-kilometer path whether the path is 1 km or
100 km from the source of the waves.
Absorption in the atmosphere is significant only over appreciable distances, measured
in kilometers, and so it is not a factor in antenna design, except in the very indirect sense
that it may dictate what frequency or frequencies are chosen for a particular application.
The earth’s ionosphere also absorbs waves at some frequencies; this absorption may be
significant in the high HF and low VHF regions. At frequencies above the low UHF the
ionosphere is completely transparent to radio waves ordinarily, and below the HF region
it acts like a reflecting barrier; that is, low-frequency waves do not penetrate the iono-
sphere appreciably.
Certain materials are capable of absorbing radio waves very strongly. Waves traveling
in these materials will be attenuated greatly within a short distance, of the order of cen-
timeters or meters. Sometimes such materials are used in antenna design to suppress
14 Electromagnetic Waves
p1
= 100.1γ ( R2 − R1 ) (1–13)
p2
The factor 0.1 in the exponent is there because decibels are ten times the logarithm of
the power ratio. Note that the mathematical form of (1–13) for attenuation due to absorp-
tion is different from (1–6), which is for space attenuation.
The total decibel attenuation AdB over this path is the sum of the space attenuation
and the absorption attenuation, which is
R
AdB = 20 log 2 + γ ( R2 − R1 ) dB (1–14)
R1
Thus the space attenuation depends on the ratio R2/R1, whereas the absorption attenuation
depends on the difference R2 − R1. Therefore close to the source the space attenuation
predominates, but at large distances from the source the absorption becomes more impor-
tant, if g has a nonzero constant value.
In the earth’s atmosphere (troposphere and ionosphere), however, the situation is
somewhat more complicated because g does vary from point to point; it is a function of
altitude (air density) in the atmosphere, and in the ionosphere it is a function of the
electron density. In such cases the absorption attenuation of a wave over the propagation
path from R1 to R2 is expressed mathematically by the following integral:
R2
AdB = ∫ γ ( R ) dR dB (1–15)
R1
Radio-Wave Optical Principles 15
1.2.1. Refraction
When a wave passes from one region to another in which the wave speed is slower or
faster, refraction occurs. Refraction takes place either when the two propagation regions
are separated by a sharp boundary or when the wave velocity varies gradually and
approximately linearly over a region that is large compared with the wavelength. Refrac-
tion at a plane boundary is illustrated in Fig. 1–3. It is to be noted that reflection also
occurs at the boundary, and it is discussed in the section 1.2.2 that follows.
As shown, a plane wavefront AB in medium I is directed toward the boundary surface.
The direction is defined by the angle of incidence q1, which the rays make with the
normal (perpendicular) to the surface. This wavefront is shown at the exact instant that
its lower edge A has reached the boundary surface.
16 Electromagnetic Waves
sin θ1 v1
= (1–16)
sin θ 2 v2
This ratio is also commonly expressed in terms of the refractive indexes of the two media.
The refractive index is defined as the ratio of the speed of the wave in vacuum, c = 3 × 108
meters per second, to the speed v in the actual medium. Thus the refractive index of medium
I is n1 = c/v1 and that of medium II is n2 = c/v2. Consequently,
sin θ1 n2
= (1–17)
sin θ 2 n1
1
v= (1–18)
µε
where m is the magnetic permeability and e is the electric permittivity. The index of
refraction (i.e., refractive index), therefore, is given by
Radio-Wave Optical Principles 17
c με
n= = (1–19)
v μ0 ε 0
where m0 and e0 are the permeability and permittivity of a vacuum. These values are
m0 = 1.26 × 10−6 henry per meter and e0 = 8.85 × 10−12 farad per meter. Then, by using
(1–18) and (1–19), and the numerical values of m0, and e0, c is attained as follows:
1
c= 3 × 108 meters per second
(1.26 × 10 ) (8.85 × 10 )
−6 −12
as previously stated. The permeability of most common dielectrics is the same as that of
a vacuum (i.e., m = m0). Thus, for most lossless dielectrics, the index of refraction n is
the ratio ε ε 0 , where e/e0 is the relative permittivity er. Therefore, from (1–16), one
may usually write
sin θ1 ε
= r2 (1–20)
sin θ 2 ε r1
To account for the conduction current and ohmic losses in a nonideal dielectric, er is
expressed as e′r − je″r, where e′r is the relative permittivity related to the displacement
current (Appendix A) and e″r is the relative permittivity related to the conduction current.
The reader may recognize that er is a complex number (for a review, see Appendix C).
Another name for e′r is dielectric constant, and for low-loss dielectrics (where e″r << e′r )
the dielectric constant (i.e., e′r ) is ordinarily assumed equal to er. Thus, (1–20) is usually
a valid approximation for most dielectrics used with antennas, but it is actually exact
only if er2/er1 is replaced by e′r2/e′r1.
Dielectric constants of typical substances used at radio frequencies range from 1 to
about 10, although a few special materials with higher values are available. Generally
the low-loss materials (low absorption coefficients) tend also to have low dielectric con-
stants. The value er = e′r − je″r = 1 applies to air (approximately) and vacuum (exactly).
When the angle of incidence of a wave is zero (q1 = 0), it is apparent from (1–16) to
(1–18) that the angle of refraction must also be zero (q2 = 0), regardless of the indexes
of refraction. That is, an incident plane wave whose rays are normal to the plane surface
between two propagation media will not be refracted (meaning that its direction of
propagation will not be changed; however, its forward speed will of course be
changed).
In Fig. 1–3 it is assumed that the wave goes from a medium of higher speed into one
of slower speed. If the order of speeds is reversed, the behavior of the rays and wavefronts
is reversed. That is, the directions of the rays in Fig. 1–3 may be reversed and everything
else will remain unchanged, except that q2 becomes the angle of incidence and q1 the
angle of refraction.
Refraction will also occur when a curved (e.g., spherical) wavefront is incident on a
plane interface between two media, when a plane wavefront is incident on a curved
18 Electromagnetic Waves
1.2.2. Reflection
In the discussion of refraction at a plane boundary (Fig. 1–3) it was tacitly implied that
the wave is totally refracted—that all the wave power incident on the surface in medium
I passes into medium II. Actually this is not true. Some of the power is reflected. Reflec-
tion refers to the formation of an additional new wavefront that travels upward from the
surface into medium I so that in general the incident wave is split up into two waves,
Radio-Wave Optical Principles 19
Γ = Γ e jφr (1–21)
where fr is the phase angle of the reflection coefficient. If the incident and reflected
electric intensities are E1 and E3, with individual phase angles f1 and f3, then G is found
as follows:
E3 E3 e jφ3 E
Γ= = jφ1
= 3 e j (φ3 −φ1 ) (1–22)
E1 E1 e E1
Thus |G | in (1–21) is given in (1–22) by the ratio of the amplitudes of the waves, and
fr in (1–21) is given in (1–22) by the difference of their phase angles.
The ratio of the power densities of the reflected and incident waves is given by |G |2,
since the power density is proportional to the square of the electric intensity (1–9).
The fraction of the total incident power transmitted through the surface (refracted) is
expressed by the power transmission coefficient T (usually called the transmission coef-
ficient). Since the Law of Conservation of Energy requires that the total of the reflected
and transmitted power shall equal the incident power, it is readily deduced that
20 Electromagnetic Waves
G 2+T =1 (1–23)
Note that T does not allow computing the electric intensity or power density of the
transmitted wave. It expressed a relationship between the total power incident on and
transmitted through the surface, just as |G |2 expresses a similar relationship for the inci-
dent and reflected power.
When the conductivity of medium II (Fig. 1–5) is very high, most of the incident wave
is reflected and very little is transmitted. In fact, for a perfect conductor |G | = 1 and T =
0. For nonconducting dielectric materials both G and T are functions of the angle of
incidence of the waves, and of the polarization, as well as of the dielectric constants of
the two media.
If medium II has some conductivity but is not an excellent conductor, waves penetrat-
ing the surface are absorbed, that is, they set up currents that are converted to heat in
the resistance of the material. In this case it is customary to express the fraction of power
that penetrates the surface in terms of an absorption coefficient, A, rather than a transmis-
sion coefficient. Equation (1–23) holds for this situation, with T replaced by A.
When the reflecting surface is curved rather than plane, the curvature of the reflected
wave is different from that of the incident wave; however, when a curved wavefront is
reflected from a plane surface, the curvature of the reflected wave is the same as that of
the incident wave. Plane and curved metallic reflecting surfaces have important applica-
tions in the design of high-frequency antennas.
Reflection may also occur from a surface that is irregular or rough rather than smooth.
Such a surface may destroy the shape of the wavefront. The reflected wave in this case
is scattered in a random fashion, so that the reflected amplitude or power density in a
given direction is not exactly predictable as it is for smooth surfaces. This phenomenon
is called diffuse reflection, whereas reflection from a smooth surface is said to be specular
(mirror-like).
Diffuse reflection is not characterized by a reflection coefficient containing a phase
angle, as is specular reflection. The random nature of diffuse reflection results in a phase
angle that varies unpredictably at different regions of the surface. Diffuse reflection,
therefore, is expressible only in terms of a power reflection coefficient, Rd.
Surfaces may also be semi-rough. For such surfaces the degree of roughness is not
sufficient to destroy the shape of the reflected wavefront completely; there is a mixture
of diffuse reflection and specular reflection. In such cases a total power reflection coef-
ficient, R, may be expressed in terms of the diffuse-power-reflection coefficient, Rd, and
the specular-reflection coefficient |G |. Since the total reflected power must be the sum
of the diffuse and specular components, the relation for R is
R = Rd + |G |2 (1–24)
When the incident wave is reflected partly specularly and partly diffusely and is also
partly transmitted, R may be substituted for |G |2 in (1–23).
For a given degree of actual roughness of a surface the values of Rd and |G | vary with
the angle of incidence qi of the wave before reflection. At normal incidence (qi = 0°),
Radio-Wave Optical Principles 21
λ
d< (1–25) Spherical
8cos θ i wavefronts,
apparent
source S2 P2
where l is the wavelength of the incident Spherical
wave and d is the depth of the surface wavefront,
irregularities, as shown in Fig. 1–6. Here source S1
the specular component of the reflected S1
qi
wave appears to be reflected from a plane h1 qi qr
intermediate between the highest and
lowest portions of the surface, whereas h2 P1
Reflecting surface
the diffuse reflection is indicated by ray qr
lines going off in all directions from each S2 qi = qr
Image wavefront,
point of the surface. from S2 h2 = h1
S1P1 = S2P1
FIGURE 1–7.
1.2.3. Principle of Images The principle of images.
When a spherical wavefront is reflected
from a plane surface, it retains its spherical shape, and its direction is changed as in a
plane wave. The exact behavior of the reflected wavefronts may be predicted by the
principle of images, as illustrated in Fig. 1–7. In that figure, the reflected wavefront
appears to be originating at S2, although it actually originates at S1 and is specularly
reflected from the plane surface. The virtual source S2 is located a distance h2 below
the reflecting surface equal to the actual height h1 of the true source S1 above the
surface and on a line from S1 perpendicular to the surface. Therefore the distance from
the actual source to a point P2, which is S1P1 + P1P2, is the same as the distance
from the virtual source S2P1 + P1P2. The virtual source S2 is called the image of the
actual source S1. This is the familiar optical principle of the mirror, according to which
an observer at P2, looking along the line P2P1, sees an image of an object located at
S1. The object appears to be at S2, although the light waves are actually coming
from S1; they are reflected from the surface in such a way that they seem to be coming
from S2.
22 Electromagnetic Waves
1.2.4. Interference
So far the optical principles considered have been in the category of geometric optics,
which means essentially that the discussion has centered on the behavior of rays and
wavefronts. Additional principles are necessary for problems in the realm of physical
optics.
Electromagnetic waves in free space, and in many other media, are subject to the
important principle of linear superposition on the basis of which interference effects can
be analyzed. Interference occurs when two or more electromagnetic waves exist simul-
taneously at the same point in space. The principle of linear superposition states that the
total electric and magnetic intensities at the point are the complex (amplitude and phase)
vector sums of the individual complex wave vectors.
This may seem to be a simple and natural fact. However, linear superposition does
not always hold in every physical situation, and sometimes it does not hold for electro-
magnetic waves. Certain types of propagation media (not commonly encountered, fortu-
nately) have nonlinear properties, so that a field of 2 volts per meter superimposed on a
field of 3 volts per meter (parallel vectors and instantaneous values) does not result in 5
volts per meter. But in ordinary media, including air and vacuum, the principle of linear
superposition does hold. (Even for the medium of air an exception may be noted, when
the voltage exceeds the value at which “breakdown” occurs.)
When the superposition principle does apply, it applies to the instantaneous values of
vectors, added vectorially. The principle of vector addition is illustrated in Fig. 1–8. (The
reader is assumed to have encountered this principle previously, and it is illustrated here
primarily for reference purposes.)
The vectors to be added are E1 and E2, which may be thought of as instantaneous
values of electric intensity at a point in space, described by equations of the type of (1–2).
Their directions in space differ by the angle a, as shown. (Their directions may differ
this much because their polarizations differ by this amount, for example; or one wave-
front may have been tilted the amount a with respect to the other by a refractive or
reflective process.)
The magnitude of the sum of the two vectors E1 and E2 is
In this equation the letters E1, E2, E3 denote the magnitudes (lengths) of the vectors E1,
E2, E3. The equation gives no information
concerning the direction of E3, which E1
Some special cases are worth noting. Referring to Fig. 1–8, if E1 = E2, it may be seen
that g = a/2 = d. In addition, if E1 = E2, and by using a standard trigonometric formula,
(1–26) becomes
α
E3 = E1 2 + 2 cos α = 2 E1 cos (1–28)
2
2π r
E1 (r , t ) = E01 sin 2π ft − + φ1 (1–29)
λ
2π s
E2 ( s, t ) = E02 sin 2π ft − + φ2 (1–30)
λ
in accordance with (1–2), except that the letters r and s are used here for distances mea-
sured along the propagation paths. This implies that the two waves may have traveled
different distances before arriving at the point under consideration from the reference
points at which r = 0 and s = 0 and at which the phase angles were f1 and f2. Moreover,
the amplitudes of the two interfering waves E01 and E02 are not necessarily the same.
The interference problem thus described has numerous antenna applications.
To avoid unnecessary complication of the equations in further discussion of this
problem, the following standard abbreviations are used:
2π f = ω (1–31)
2π
=β (1–32)
λ
The quantity w (omega) is known as the angular frequency, and b (beta) is the phase
constant; w expresses the angular rate (in radians) at which the phase angle of the wave
24 Electromagnetic Waves
at a fixed point in space changes in time, and b is the rate of phase change with distance
at a fixed instant of time. The customary units for w are radians per second, and for b,
radians per meter; that is, t and l are ordinarily measured in these units. Thus a time
period T = 1/f corresponds to 2p radians, or 360° of phase angle, as does also a distance
x = l. (T = 1 cycle and l = 1 wavelength.)
If this notation is used and the principle of superposition is applied, it is apparent that
when E1 and E2 are parallel vectors at the same point in space the resultant intensity is
No distance variable is indicated for E3, since it expresses the intensity at a single point
only. Therefore E3 is a sinusoidally varying electric intensity with an amplitude and phase
angle
E03 = E01
2
+ E02
2
+ 2 E01E02 cos [φ1 − φ2 − β (r − s )] (1–35)
and
Although these equations appear complicated, they are actually simple in the sense
that if numerical values of all the quantities on the right-hand sides are known numerical
values of E03 and f3 are easily calculated.
It is easily shown that (1–35) and (1–36) are essentially the same as (1–26) and (1–27),
if a is equated to [f1 + f2 − b(r − s)] and if the angle (f1 − br) is arbitrarily taken to be
zero, which is permissible, for it amounts only to choosing a particular origin for the
coordinate system. Therefore, as is well known in a-c circuit theory, the same geometric
construction (Fig. 1–8) can also be applied in this case.
When E1 and E2 are different both in direction and phase (other than 180°) the case
is more complicated and is not treated in detail here. The resultant electric intensity is a
space vector that undergoes an elliptical variation in its direction and amplitude in the
general fashion described for an elliptically polarized wave (see Appendix B), except
that the plane of the ellipse is not perpendicular to the direction of propagation unless
E1 and E2 are propagating in the same direction.
The space–time relationships that have been described are often confusing to
those who are encountering them for the first time. They become clear only after famil-
iarity has been gained by working problems and by thinking about them at some length.
Some helpful exercises are given at the end of the chapter. It is worthwhile to achieve
Radio-Wave Optical Principles 25
familiarity with these ideas, because the interference phenomenon is the basic principle
of most directional antennas, and it has other applications as well.
A Obstacle
Shadow
region
B
C
radiation. Radiation results when high-frequency electric currents flow under suitable
conditions. The detailed mathematical description of the relationship between electric
currents and their associated fields can be derived from Maxwell’s equations. These are
four partial differential equations (as usually formulated), and on them the whole struc-
ture of electromagnetic theory is based. The practical successes of this theory place it
among the most impressive of man’s scientific achievements.
28 Electromagnetic Waves
Equations (1–2) and (1–3) are solutions of Maxwell’s equations for the simplest situ-
ation, that of a wave in free space at a great distance from the radiating source—in short,
a plane wave. Equations 1–2 and 1–3 give no information on how radiation occurs, but
Maxwell’s equations do give this information. They state that whenever electric current
flows, a magnetic field is set up in the surrounding space. Any variation of this magnetic
field will result in creation of an electric field. The magnetic field will vary, of course, if
the current varies; therefore, with alternating current the magnetic field will be continu-
ously varying and will therefore continuously generate an electric field. Because the two
fields always exist together—one cannot exist without the other unless it is nonvary-
ing—the combination is called the electromagnetic field. The time variations of the field
components will be related to the nature of the variations of the current; if it is sinusoidal,
the fields will vary sinusoidally at the same frequency.
Moreover, the varying electric field, according to Maxwell’s equations, generates a
magnetic field. This exchange between the two fields provides the mechanism by which
waves may be propagated in space at the “speed of light.” The fields are a form of energy,
and therefore electromagnetic wave propagation represents a transport of energy outward
from the radiating source. The energy is supplied by the electric power source that is the
origin of the current.
Not all of the field surrounding a current-carrying conductor results in propagation of
waves outward into space. Some of the energy of these fields is returned to the conduc-
tor; it is temporarily stored in the fields, which are related to reactive effects—that is,
inductive and capacitive effects. The total field consists of two components—the induc-
tion field and the radiation field. The induction field is confined to a fairly local region
near the conductors; the radiation field may propagate to great distances, although its
strength will decrease with distance, both by spherical spreading of the wavefront
(inverse-square law) and by absorption if there is any. This radiation process will be
described in further detail for the case of a dipole radiator, in sec. 4.1.
accelerations may be due to collision with other particles, interaction with electric and
magnetic fields, or to energy-level transitions of electrons in atomic orbits.
region, but ordinarily the field strength from this effect is too small to be of practical
value for radio communication. Thus, “line-of-sight” radio propagation from an antenna
of height h, Fig. 1–12, would be limited to the horizon distance, dh, assuming that the
receiving point is close to the earth’s surface.
The spherical earth model is frequently used to estimate the propagation path of waves
over the earth (Appendix D). With this model, the effects of the earth’s curvature and
refraction (bending) in the atmosphere are estimated by assuming the propagation is
along straight lines and that the earth has an “effective” radius of aE. For representing
average or “typical” atmospheric effects, aE is usually taken to be 8,493 km, which is
4/3 times the 6,370 km value geophysicists conventionally assume for the actual earth
radius. With this model, the radar horizon dh, that is, the distance from antenna at height
h to the tangent point on a smooth earth of radius aE, is
From (1–37a), by neglecting the h2 term (which is ordinarily much less than 2aEh) and
letting aE equal 8,494 km (the 4/3 earth radius), one gets (1–37b)
dh = 1.07do = 1.15dg
Therefore, under normal conditions, atmospheric refraction extends the distance to the
radio (and radar) horizon dh beyond the geometric horizon dg, and dh is at a slightly
greater distance than the optical horizon do (the one observed visually).
between Poldhu, England, and stations he established in Nova Scotia and on Cape Cod
(Kraus, 1988, p. 5). At VLF, the lower edge of the ionosphere behaves virtually like a
perfect conducting surface, and vertically polarized waves at these frequencies may be
propagated around the earth as if they were confined between two perfectly conducting
spherical shells, the inner shell being the earth’s surface. This is known as VLF wave-
guide-mode propagation; and it provides effective, narrow bandwidth, and reliable round-
the-world communication. However, extremely high power and a large transmitting
antenna are required.
By the early 1920s, the use of “wireless” had progressed into the present-day AM
broadcast band (535–1605 kHz). At these higher frequencies vertically polarized waves
may propagate to some distance beyond the line-of-sight horizon by means of a surface
wave. However, this surface wave provides only a moderate extension of the line-of-sight
limitation. Surface wave propagation depends on the fact that a conducting surface
tends to “guide” a wave polarized perpendicularly to it, so that the wavefront remains
approximately perpendicular and thus maintains a propagation direction parallel to the
surface.
However, the guiding effect is only partial when there is but one guiding surface,
rather than two as in the VLF waveguide mode. The surface wave partially escapes from
the guiding surface as it progresses, so that the method is not effective beyond a distance
that depends on the frequency. It also depends on the conductivity of the earth, being
greater over the ocean, for example, than over dry land. Usable field strengths are attain-
able at distances ranging from about 150 kilometers (at the higher end of this frequency
range to thousands of kilometers at the lower end, although here the VLF waveguide
mode is probably partially operating. The surface wave is the principal mechanism of
beyond-the-horizon broadcast-band propagation in the daytime and for propagation to
moderate distances at night. The long-distance reception observed at night is due to iono-
spheric reflection.
The principal method of reaching a point around the curve of the earth consists of two
steps, namely, propagation by a path that first goes upward to reach a point high above
the earth and then downward into what
would be the shadow region for direct-
path propagation. This process is illus- P2
trated in Fig. 1–13. The transmission from
point P1 on the earth’s surface reaches
point P3 by first going to P2, high above Direct-path
the earth, then down again to P3. As shadow region
ionosphere. This layer of electrically charged particles (actually, several layer-like regions
at different altitudes) acts as a reflector of radio waves at the lower frequencies—generally
from about 10 MHz downward, although the exact upper limit varies with time of day and
other factors. The ionization is produced by the sun’s radiation and so is most intense in
the daytime; however, it persists to some degree through the night so that ionospheric
propagation is possible throughout the twenty-four hours at some frequencies. For any
particular path there is likely to be an approximately optimum frequency at a particular
time of day. Round-the-world propagation is possible by means of multiple reflections.
The ionosphere extends from about 60 to more than 300 kilometers above the earth. (It is
a reflecting region rather than a reflecting surface; it is in fact more logical to regard the
“reflection” as actually a refraction process, at least at the higher frequencies.)
The ionosphere continues to be a valuable means of long-distance radio propagation;
but as it is subject to the vagaries of solar activity and is restricted to the lower frequen-
cies, it has limited usefulness. Other means have therefore been sought, and found in the
years following World War II.
Tropospheric scatter propagation is useful for distances up to a few hundred kilometers
at frequencies in the VHF and UHF. In this method the waves are “scattered” downward
from irregularities of the atmospheric dielectric constant. The exact nature of these scat-
tering regions is a subject of some controversy. Scattering may be regarded as a diffrac-
tion process, semireflective in effect. Because the scattered field strength is not a very
large percentage of the transmitted field, large antennas and high transmitter powers are
required. Nevertheless, the scatter method has proved to be advantageous for fixed-point
communication links, largely because of the freedom from ionospheric variability effects
and because the distances involved are too short for good ionospheric reflection. It has
the further important advantage of permitting the use of frequencies too high for iono-
spheric reflection, thus relieving the overcrowded conditions in the ionospheric-reflection
portion of the radio spectrum.
Scattering from ionospheric irregularities and from the ionized trails of meteors have
also been found to be practical methods. Ionospheric scatter is limited to the frequency
region of about 25 to 60 MHz; it is effective up to about 1,500 km. Meteor scatter can
be employed from 6 to 75 MHz also for distances up to about 1,500 km.
Because scatter methods are subject to limitations of distance and frequency useage,
National Aeronautical and Space Administration (NASA) began investigating satellite
communications with passive, metallized reflecting spheres (http://msl.jpl.nasa.gov/
QuickLooks/echoQL.html). The first of these satellites, Echo 1, was placed in orbit at
a height of about 1,500 km in 1960; it was a reflecting sphere 100 feet (30.5 m) in
diameter, with an orbital period of about 2 hours. Echo 2, 135 feet (41.1 m) in
diameter, was launched in January 1964. Reflection via these satellites required large
directional antennas that followed the satellite in orbit. Echo 1 reflected 960 and
2,390 MHz signals; and Echo 2 reflected 162 MHz signals. Echo 1 successfully redi-
rected transcontinental and intercontinental telephone, radio, and television signals.
Echo 2 continued the passive communications experiments, and also investigated the
dynamics of large spacecraft and was used for global geometric geodesy. Although NASA
abandonded passive communications in favor of active satellites following Echo 2, the
Environmental Wave-Propagation Effects 35
Echo systems demonstrated several ground station and tracking technologies later used
by active systems.
Active satellites receive signals from the earth, and then retransmit them at a different
frequency to another location. These satellites contain receiving and transmitting equip-
ment, including, of course, receiving and transmitting antennas. There are about 150
communications satellites operational today, and they use geostationary, Molnya, or low
polar circular orbits (http://en.wikipedia.org/wiki/Communications_satellite).
Geostationary satellites appear stationary as viewed from the rotating earth; and they
continue to operate above the equator, at an altitude of about 35,000 km. Although the
path lengths are long, an important advantage is that the earth-based antennas do not
have to track a rapidly moving satellite. However, since geostationary satellites remain
over the equator, they are not well suited to high latitude (far North or far South) cover-
age. Molnya orbits are highly inclined, providing good Northern coverage. A major use
of Molnya satellites is for telephony and TV over Russia.
A low earth orbit typically is circular and about 400 kilometers above the earth’s
surface, and time to revolve around the earth is about 90 minutes. Thus, they change their
position relative to a fixed ground position quickly, and thus a large number of satellites
are needed for uninterrupted connectivity. However, discontinuous coverage is possible
by storing data received while passing over one part of earth and transmitting it later while
passing over another part. Low earth orbiting satellites are closer to the ground than geo-
stationary satellites and consequently their equipment costs are less. Thus, relative costs
involve trades between the number of satellites and their individual costs.
For special purposes reflection from the moon’s surface has been used as a communi-
cation method at VHF. This method has the disadvantage, of course, of being limited to
the times that the moon is above the horizon; it also requires very high power and large
antennas.
At some frequencies, usually VHF and higher, unusual atmospheric conditions some-
times result in abnormally great downward bending of ray paths, owing to an extreme
effect of the type illustrated by Fig. 1–4. Under such circumstances waves may be propa-
gated in a path that follows the curve of the earth, hence into the normal shadow region.
This effect is called superrefraction or sometimes “trapping” or “ducting.” It usually
occurs in warm climates when unusual moisture and temperature conditions prevail and
is not sufficiently reliable or predictable for most beyond-the-horizon propagation appli-
cations. However, under very unusual conditions propagation by this method beyond
1,500 km has been observed, and 150-km paths are not unusual.
direction and the surface, called the grazing angle). At larger grazing angles semispecular
reflection may occur, as shown in Fig. 1–6. For smooth earth surfaces, effects of specular
reflection may be observed at frequencies up to at least 35 GHz.
Reflection from a plane surface was illustrated in Fig. 1–5, and the principle of images
for specular reflection was shown in Fig. 1–7. The reflection that occurs from the earth’s
surface may be understood in terms of these concepts. Although the earth is approxi-
mately spherical, the region of reflection may be regarded as virtually plane if the antenna
is located within about a hundred meters of the earth’s surface.
Waves reflected from the earth’s surface are important because they interfere with
waves propagated in a direct path, as illustrated in Fig. 1–14. A transmitting antenna is
assumed to be located at Sl, at a height h above a perfectly reflecting plane surface; and
therefore its image is located at S2, a distance 2 h from Sl. A direct wave travels to a
receiving point at P2 following a straight-line path (except for the slight downward cur-
vature due to atmospheric refraction, which may be disregarded here). A reflected wave
also arrives at P2 via the path S1P1 + P1P2. However, in accordance with the image prin-
ciple, the reflected wave may be regarded as originating at S2 and proceeding to P2 via
the straight-line path S2P1 + P1P2. The elevation angle of point P2 as viewed from Sl is
designated q. Notice that the grazing angles y, whether entering or exiting point P1, are
equal in accordance with the principles of optics (Fig. 1–5).
Since the two waves arriving at P2 are of exactly the same frequency, being from the
same source, they will interfere in accordance with (1–35) and (1–36). Whether the inter-
ference is constructive or destructive depends on the phase difference. Since the lengths of
the two wave paths are not the same, there will in general be a path difference d and a con-
sequent phase difference equal to bd radians (b = 2p/l). At some points in space the two
waves will reinforce each other; at other points they will tend to cancel so that there will
be regions of zero or near-zero field strength. This effect can have serious consequences;
for example, it can create “blind” regions in space where signals cannot be detected.
Figure 1–14 is used to analyze a relatively simple multipath propagation problem that
provides the results of Fig. 1–15. Applicable assumptions follow:
a. The wavelength is long enough that the surface is smooth, in accordance with
the Rayleigh criterion [equation (1–25)]. Ordinarily, even in the presence of
the earth’s surface irregularities and undulations, this assumption is valid for
VHF (30–300 MHz) and lower frequencies.
b. The distance S1P2 between antenna and the observation point is short enough
that the earth’s curvature is neglected.
c. The antenna elevation pattern is sufficiently broad that the amplitude and
phase of the radiated waves are the same at every elevation angle q.
d. The reflection coefficient amplitude and phase delay are constant, correspond-
ing to |G| = 1 and f = −p radians, respectively. This is a valid approximation
for horizontal polarization and a reflecting surface that is sea water or another
good conductor.
e. The paths S1P2 and S2P2 are long enough, relative to 2 h, so that they are
effectively parallel.
Environmental Wave-Propagation Effects 37
P2
th
Direct- wave pa
S1
q
th
e pa
av
d-w
e cte
Refl
y y
2h Reflecting Surface
P1
q1
2hsinq1
S2
FIGURE 1–14.
Geometry of interference between direct-path and reflected-path waves.
Elevation
For purposes of the analysis, we denote angle, q
the electric intensity of the direct-wave 20°
as Ed and the reflected wave as Er.
Then, based on the above assumptions 10°
and Fig. 1–14, the electric field E result-
ing from the phasor sum of the direct-
Horizontal distance
wave and the reflected-wave fields at P2
follows: FIGURE 1–15.
Interference-pattern field-strength contours
E = Ed + Er = Ed + Ed Ge − jβδ relative to free-space contour, for h/l = 1.44
= Ed 1 + G jφ e − jβδ (1–38) and G = −1.
Here G is the complex (amplitude and phase) reflection coefficient of the earth’s surface,
and bd is the phase delay (in radians) caused by the difference d of the lengths in the
38 Electromagnetic Waves
direct- and reflected-wave paths. It can be seen in (1–38) that the amplitude and phase
of G are expressed by |G | and e jf, respectively. Also, recall that b = 2p/l. Thus, the
magnitude of E may be expressed as
Note that within the second absolute-value brackets of (1–39) there is a sum of two
phasors, separated in angle by f−bd radians. Consequently, by using the law of cosines
or other means, one finds the magnitude of E to be
E = Ed 1 + Γ 2 + 2 Γ cos(φ − βδ ) (1–40)
Then, by using the assumptions |G | = 1 and f = −p of item d in the list above and a half-
angle formula, (1–40) becomes
βδ
E = 2 Ed 1 + cos(−π − βδ ) = 2 Ed 1 − cos(βδ ) = 2 Ed sin (1–41)
2
where, as previously, b = 2p/l. Note that changes in the path length difference d cause
|E | to vary between 2|Ed| and zero.
From Fig. 1–14 it can be seen that d equals 2hsinq1 requires parallel S1-P2 and S2-P2
paths, but the angle q1 is generally not known. However, in many circumstances, the
path S2-P2 is approximately parallel to path S1-P2 because S1-P2 is very much larger than
height h. Then, q1 is approximately q, where q is the more likely known angle which is
between the direct-wave path and the local horizontal at the antenna. Thus, when q ≈ q1,
the path length difference d is approximately 2hsinq and then
2π
δ = 4π sin θ
h
βδ = (1–42)
λ λ
Hence for a given value of h/l (antenna height expressed in wavelengths) the interfer-
ence becomes a function of the elevation angle q only. A final simplification of (1–38)
is then possible from (1–40) and (1–42):
2π h sin θ
E = 2 Ed sin (1–43)
λ
Application of this formula to the particular value of h/l = 1.44 results in the field-
strength pattern shown in Fig. 1–15.
The contours, or lobes, represent surfaces of constant field strength. The origin of the
coordinate system is the location of the transmitting antenna. Elevation-angle markers
Environmental Wave-Propagation Effects 39
As now discussed, the Brewster angle effect causes large variations in |G | and f. The
Brewster angle is the grazing angle at which the reflection coefficient is zero, and it
occurs only for V POL and a lossless dielectric. Instead, for land and sea, which are
imperfect dielectrics, |G | for V POL has a non-zero minimum at the so-named pseudo-
Brewster angle ypB. Accordingly, |G | is much larger for grazing angles both smaller and
larger than ypB. At grazing angles larger than ypB, f becomes close to 0°, f is approxi-
mately −90° at ypB, and f approaches −180° as y approaches zero. Thus, f will be very
near −180° for very small grazing angles, even with V POL and poorly conducting
ground. Additionally, |G | increases and approaches unity for grazing angles y that
approach zero. Therefore, an interference null exists at small grazing angles for both H
and V polarizations—however, as now discussed, below 1 MHz, the V POL nulls near
the horizontal direction will be increasingly filled as the frequency is reduced.
The pseudo-Brewster angle ypB is at considerably smaller grazing angles for seawater
than for land, and it becomes smaller for both land and sea with increasingly longer
wavelengths. Interesting, the movement of ypB closer to zero grazing permits acceptable
V POL performance (without an interference null) down to very small grazing angles at
low frequencies. For smooth planar surfaces at 1 MHz, ypB is about 0.2° and 3° for sea
and typical land, respectively. At 100 kHz, the corresponding ypB values are about 0.07°
and 1° for sea and land, respectively. Natural surface-slope undulations and other irregu-
larities diminish the destructive interference and thereby further enhance signal ampli-
tudes at the near-zero grazing angles. Therefore, at the lower frequencies it is possible
to place a vertical-polarization radiator close to the ground and obtain acceptably strong
radiation in the horizontal direction. This phenomenon permits the effective use of V
POL for frequencies less than about 1 MHz.
Although the interference of direct and reflected waves creates undesirable minima in
the radiation pattern, it also creates maxima in field strengths. Since the earth is not a
perfect reflector, the amplitude of the reflected wave will be less than that of the direct
wave, with the result that the field strength will be greater than zero at the minima and
less than twice the free-space value at the maxima. As discussed in more detail in Appen-
dix D, the earth’s surface roughness and curvature reduces |G |, and thus the pattern nulls
are partly filled and the maxima reduced. The reader will recall that, in accordance with
the Rayleigh criterion, equation (1–25), surfaces become rougher (electromagnetically)
with increases in the radio wave frequency.
1.4.6. Transionospheric
Propagation
Now that radio transmission to and from Ionosphere
various outer-space locations is com-
monplace, the effect of the ionosphere on
waves passing through it is of impor-
Antenna Ray path
tance, whereas at one time the ionosphere
was of interest primarily because of its
ability to reflect radio waves back to
earth. Naturally, transmission through
Earth
the ionosphere requires the use of a fre-
quency well above the frequencies that FIGURE 1–17.
are reflected from it; generally, frequen- Refraction in transionospheric propagation.
cies above 10 MHz are necessary.
A wave that is above the critical reflec-
tion frequency may nevertheless be appreciably refracted in passing through the iono-
sphere, as illustrated in Fig. 1–17. As shown, the wave path is at first bent downward as
it enters the ionosphere from below. Then, after reaching the region of densest ionization,
it is bent back upward, and eventually emerges with nearly its original direction. If the
direction of the antenna beam at the earth were used to determine the direction of a source
of incoming waves, this bending effect would cause some error in the determination.
Therefore it is important to be aware of the effect. If the ionosphere electron-density
height profile is known, the magnitude of the effect can be calculated. Generally it is not
an appreciable effect above 100 MHz and is less at night than in the daytime. A similar
statement applies to the absorption losses that occur in the ionosphere.
A more important effect in transionospheric propagation below a few hundred MHz
is the Faraday rotation of the polarization of a wave. This rotation occurs as a result of
42 Electromagnetic Waves
the combined effect of the charged particles (electrons) of the ionosphere and of the
earth’s magnetic field. Because of the variability of the effect, it is not feasible to attempt
to predict the exact amount of rotation at low frequencies, where the total rotation may
even be many multiples of 360 degrees. This rotation is important because, if linearly
polarized antennas are used for both transmission and reception, the received polarization
will sometimes be correct and sometimes incorrect for optimum reception. If the polar-
ization of the signal is at right angles to that which the antenna is intended to receive, in
principle no reception will result.
Several remedies are available, however. One is to use either circular polarization, or
two linear polarizations at right angles to each other, so that one or the other of them
will at all times receive a signal. Another remedy is to use a linearly polarized receiving
antenna whose polarization direction can be varied to agree with the polarization of the
incoming wave. The third remedy is to operate at a high enough frequency so that no
appreciable rotation occurs—above 1,000 MHz, or possibly a bit lower. The present
tendency for space communication and other space transmissions is to use frequencies
between about 1 and 10 GHz, although avoidance of Faraday rotation is only one of
several reasons for using this frequency region.
Pt Gt
pr = (1–44)
4π R 2
30 Pt Gt
Er = (1–45)
R
Environmental Wave-Propagation Effects 43
When the path between the transmitter and receiver is not in “free space,” a modifica-
tion of these equations is required. This is done through use of the pattern-propagation
factor F. Following the IEEE definition (IEEE Standard 686–1997, p. 19), F is a scalar
and it is the ratio of the field strength that is actually present at a point in space to that
which would have been present if free-space propagation had occurred with the antenna
beam directed toward the point in question. Accordingly,
Er (actual )
F= (1–46)
Er ( free space)
By merely rearranging this definition, (1–45) which gives Er(free space) can be converted
to a nonfree-space equation, as follows:
F 30 Pt Gt
Er (actual ) = FEr ( free space ) = (1–47)
R
Similarly, the equation for power density pr under nonfree-space conditions becomes
Pt Gt F 2
pr = (1–48)
4π R 2
Ordinarily, the gain of an antenna varies with direction of radiation, but the gain Gt in
(1–47) and (1–48) is the gain in a specific direction of observation. However, for use of
the above IEEE definition of F, one is reminded that, in a nonfree-space environment, the
actual field strength at a point may depend on gain in directions other than that toward the
point in question. For example, in accordance with Fig. 1–14, waves radiated downward
from an antenna may reflect from the earth and be summed at a point in space, both in
phase and amplitude, with waves arriving directly from the antenna. Therefore, with the
definition of F, antenna gain versus direction must be considered when more than one
direction is involved, unless the antenna gain is the same in the various directions.
Effective receive aperture area Ar, and its relationships with maximum receive antenna
gain Gr and received power Pr, is discussed in chapter 3, sec. 3.4. Received power Pr
equals prAr, assuming the receiving antenna and the incident wave are polarization
matched, where pr is the power density incident on the area Ar, and Ar = Grl2/4p. Then,
by using these relationships and (1–48), a useful relationship for received power
follows:
Pt Gr Gt λ 2 F 2
Pr = (1–49)
(4π )2 R 2
The reflection-interference pattern calculation, exemplified by Fig. 1–15, can be used to
determine F, for the case where the antenna gain is independent of the elevation angle.
Accordingly,
44 Electromagnetic Waves
E
F= (1–50)
Ed
since Ed is the free-space field strength. With E expressed in terms of (1–43), (1–50)
becomes
2π h sin θ
F = 2 sin (1–51)
λ
In this case the maximum value of F is obviously 2 (when the sine function has its
maximum numerical values ±1), and the minimum value is zero. In free space, of course,
F = 1.
Appendix D and the accompanying website discuss calculations of the pattern-propa-
gation factor for smooth and rough land and sea surfaces, under assumptions that the
earth is either flat or spherical.
References
Ah Yo, D. M. K., and R. Emrick, “Frequency Bands for Military and Commercial
Applications,” ch. 2 in J. L. Volakis (ed.), Antenna Engineering Handbook, 4th ed.,
McGraw-Hill, 2007.
Blake, L. V., “Tropospheric Absorption Loss and Noise Temperature in the Frequency
Range 100–10,000 Mc”, IRE Transactions on Antennas and Propagation, AP-10,
January 1962, p. 101.
IEEE Standard 100–1992, The New IEEE Standard Dictionary of Electrical and Electronic
Terms, 1992.
IEEE Standard 211-1997, The IEEE Standard Definitions of Terms for Radio Wave
Propagation, December 1997, p. 27.
IEEE Standard 686–1997, IEEE Standard Radar Definitions, 25 March 1998.
Kraus, J. D., Antennas, 2nd ed., 1988.
3. The rms electric intensity of a wave in free space is 2.7 × 10−3 volt per meter.
(a) What is its power density in watts per square meter? (b) What is the rms
magnetic intensity in amperes per meter?
(b) If the initial point in part (a) is at a distance R1 = 100 km from the source
of the waves, so that the wavefront must now be regarded as spherical (and
spreading) rather than plane, and if a second point is at a distance R2 = 400 km
from the source in the same direction, what is the total attenuation in decibels
due to both the spreading and the absorption? (Note: Because the medium is
assumed to be of infinite extent, the wave propagation will follow free-space
laws except for the absorption; Eq. 1–14 applies.)
5. A plane wave in free space is incident upon the plane surface of a dielectric
material at an incidence angle of 30 degrees. The dielectric constant of the mate-
rial e′r = 4. By what angle does the direction of propagation change as the wave
passes into the dielectric medium? (That is, what is the difference between the
angle of incidence and the angle of refraction?)
8. A radio station has a total radiated power Pt = 10,000 watts, and an antenna gain
Gt = 30. If free-space propagation is assumed, what is the rms electric intensity
of the radiated field at a distance R = 100 km (105 meters), expressed in volts
per meter?
9. A shipboard transmitting antenna radiates Pt = 1,000 watts and has a power gain
Gt = 100. Its height above the sea surface is such that at a certain elevation angle
the propagation factor is F = 2. An airplane at a range R = 50 km (5 × 104 meters)
is at an altitude corresponding to this elevation angle. What power density exists
at its receiving antenna due to the transmission from the ship?
46 Electromagnetic Waves
Transmission Lines
This chapter discusses transmission lines. Transmission lines must be addressed in the
study of antennas for three reasons: (1) A transmission line virtually always connects an
antenna to a transmitter or receiver, and is often regarded as a part of the antenna system;
(2) in some types of antennas transmission-line elements are integral parts of the antenna;
and (3) the principles of transmission lines are applicable to understanding some aspects
of antenna theory.
The material of this chapter, like that of chapter 1, is intended for review and refer-
ence. A complete treatise on transmission lines would provide material for an entire book.
Therefore only the basic aspects of the subject will be covered here, with emphasis on
antenna applications.
47
48 Transmission Lines
L R1 L R1 L R1 L R1
Input C R2 C R2 C R2 C R2 Output
x=1
FIGURE 2–3.
Equivalent-circuit representation of a transmission line.
arrangement of many lumped-circuit elements (of short length) as shown in Fig. 2–3. In
other words, small lumped circuit elements are used to connote the infinitesimally small
amounts of inductance, capacitance and resistance that are distributed uniformly along a
transmission line.
Each identical circuit section in this representation corresponds to a unit length of the
line, that is, L is the inductance per unit length of the conductors, C is the capacitance
between the conductors per unit length, R1 is the series resistance of the conductors per
unit length, and R2 is the shunt or “leakage” resistance per unit length. For this repre-
sentation of a line to be valid, it is necessary to assume that the “unit of length” is a very
small fraction of the total line length, so that there will be very many of the identical
lumped unit-length sections. More specifically, it is necessary that the unit of length be
chosen small compared to a wavelength on the line, defined in the same way as for
electromagnetic waves in space:
v
λ= meters (2–1)
f
where v is the velocity of propagation of the voltage and current “waves” along the line,
meters per second, and f is the frequency in Hertz of the applied voltage.
These assumptions are necessary to justify certain steps in the analysis; but after the
analysis is made it is found that the results involve ratios of the unit-length circuit values.
For this reason, in application of the results, L, C, R1, and R2 can be expressed as values
for unit of length, for example, henrys per meter (H/m), farads per meter (F/m), ohms
per meter (W/m).
To describe the behavior of a line in these terms, it is helpful to consider two special
cases that are idealizations, that is, are never fully attainable in reality although they may
be closely approximated. The first of these is the line of infinite length (l → ∞ in Fig. 2–2)
and is discussed in sec. 2.1.2. The second is the lossless line discussed in sec. 2.1.3,
corresponding to R1 = 0 and R2 → ∞ in Fig. 2–3.
50 Transmission Lines
Vi ( t ) = V0 sin (ω t + φ ) (2–2)
in which V0 is the amplitude and f is the phase angle at t = 0. Since the line is lossless,
this voltage will travel down the line at velocity v with undiminished amplitude. If
distance along the line is denoted by x, the voltage at the point x will be given by
V ( x, t ) = V0 sin (ω t − β x + φ ) (2–3)
provided that enough time has elapsed for the voltage wave to reach the point x, that is,
at times t greater than t = x/v.
The quantity b, known as the phase constant, is defined as 2p/l. Equation (2–3) is
identical to the equation of a plane electromagnetic wave of equation (1–2), ch. 1, with
the substitution of a voltage V for the electric field intensity E.
The analysis of the transmission line of Fig. 2–3 leading to this result and to the further
results that will be presented is made by writing and then solving second-order differen-
tial equations, based on the assumptions that the circuit sections of Fig. 2–3 represent
infinitesimal lengths of the line. From the analysis, which will not be given here, it is
also found that the velocity of the voltage waves is
1
v= (2–4)
LC
for a lossless line, where L and C are the inductance and capacity per unit length of line.
In this case if L and C are in henrys and farads per meter, v is obtained in meters per
second.
Note the similarity of (2–4) to the equation for velocity of an electromagnetic wave
in a lossless medium, equation (1–18), ch. 1. Moreover, it turns out that if the transmis-
sion line is primarily immersed in air or a vacuum (except perhaps for solid insulating
supports of low loss at well-spaced intervals), the product of L and C will be a constant
such that v is equal to 3 × 108 meters per second—exactly the velocity of an electromag-
netic wave in space. Therefore the wavelength l, as given by (2–1), will be the same on
the line as for a wave in space.
These statements assume that the conductors are straight, smooth-surfaced, and are
made of nonferrous metal having magnetic permeability of the same value as that of air
or vacuum, that is, a relative permeability of unity. It may be noted in passing that copper-
plated steel wire does not violate this assumption. This is because, at radio frequencies,
the phenomenon called skin effect confines the current and fields to the surface of the
Basic Transmission-Line Concepts 51
wire and the external region of space; therefore the steel core does not affect the
inductance.
But if a material having relative permittivity or relative permeability greater than one
occupies the region between and surrounding the conductors, L and C will have a larger
product. Then, v will be less than 3 × 108 meters per second, and is given by
c
v= (2–5)
µr ε r
where c = 3 × 108 meters per second, mr is the relative permeability, and er is the relative
permittivity. For most dielectric materials mr = 1. A lossless line is being discussed here,
where er is a real number. For a lossy dielectric, er is complex (see sec. 1.2.1), and then
er in (2–5) must be replaced by its real part (the dielectric constant).
When the voltage Vi(t) is applied to the input terminals of the line, a current flows
into the line. It is related to the input voltage by a quantity called the characteristic
impedance, Z0. That is,
Vi( t )
Ii(t ) = (2–6)
Z0
L
Z0 = ohms (2–7)
C
Equation 2–6 is of course Ohm’s law, and (2–7) is analogous to the equation for the wave
impedance of a propagation medium, equation (1–8), ch. 1. Therefore the current as a
function of distance x along the line is given by
V
I ( x, t ) = 0 sin (ω t − β x + φ ) (2–8)
Z0
V0
I0 = (2–9)
Z0
where the subscript rms stands for “root mean square” or effective value; it is related to
the amplitudes by the well-known relations
V0
Vrms = = 0.707V0 (2–11)
2
and
I0
Irms = = 0.707I 0 (2–12)
2
2
Vrms
P= = I rms
2
Z0 (2–13)
Z0
It is to be emphasized that this result applies to the power flow in terms of in-phase
voltage and current in an infinite line, or in a finite line before the voltage wave reaches
the load.
The load impedance ZL may be purely resistive, purely reactive, or some combination
of resistance and reactance. Suppose that it is purely resistive and of value RL. Then the
power delivered to the load will be
VL2
PL = = I L2 RL (2–14)
RL
where VL and IL are the voltage and current at the load. Comparison of (2–13) and (2–14)
shows that if RL = Z0, all the power flowing down the line will be delivered to the load.
But if RL does not have this value, the two powers cannot be equal if the voltage and
current of (2–14) are assumed to be the same as the voltage and current of (2–13).
What actually happens is that some of the voltage and current are reflected. The
reflected voltage and current then travel back toward the input end of the line, at the
Basic Transmission-Line Concepts 53
velocity v. Consequently, at every point x along the line, the voltage is that resulting from
linear super-position of the original (incident) voltage wave, of amplitude V0(i), and a
reflected wave, of amplitude V0(r). A similar statement applies to the current.
This situation is entirely analogous to the problem of interference between two
electromagnetic waves at a point in space. The voltage and currents at various points
on the line may either add or subtract from each other, depending on the difference
in their phase angles, which in turn varies from point to point. The equation of the
reflected voltage wave corresponding to (2–3) for the incident wave is obtained by sub-
stituting V0(r) in place of V0, and (2l − x) in place of x. The phase angle f also will in
general be changed in the process of reflection, depending on the phase angle of the
impedance ZL.
The result is that the voltage amplitude will vary along the line. There will be maxima
at half-wavelength intervals, with minima at positions halfway between the maxima. The
positions of these maxima and minima do not change with time, that is, they do not move
along the line. The resulting pattern (actually, an interference pattern) is therefore a
standing wave. A standing wave of current is also set up in the same way. The current
maxima coincide with the voltage minima, and vice versa (see Fig. 2.5 of sec. 2.2.6).
The arrival of the reflected voltage and current waves at the line input terminals results
in a changed voltage-current relationship at this point. Consequently, the effective input
impedance of the line is no longer equal to the characteristic impedance Z0, in general.
It is by definition the resulting ratio of voltage to current.
Because the effective input impedance undergoes a sudden change when the reflected
wave first reaches the input terminals, the applied voltage Vi may also undergo a change
at that instant (since the output voltage of the usual source, or generator, will vary when
the impedance presented to it varies). The voltage wave traveling down the line therefore
takes on a new value at this instant, and so the reflected wave coming back 2l/v seconds
later will also have a suddenly different value at that instant, resulting in a further read-
justment of the input voltage, and still further adjustments at future intervals of 2l/v
seconds. However, these readjustments become progressively smaller and gradually “die
out.” They are called transients. When they have died out, a steady state is said to exist.
(Parenthetically, if the source impedance were Z0, there would not be reflections at each
end of the transmission line).
For some purposes it may be important to analyze the transient behavior of the line,
but ordinarily only the steady-state behavior is important. This may be seen by consider-
ing a typical value of the time 2l/v required for a wave to travel down the line and be
reflected back to the input. For a line 100 meters long (328 feet) this time is
2l 2 × 100
t= = = 6.7 × 10 −7 second
v 3 × 10 8
end. Transient effects cannot be ignored if the time 2l/v is comparable to the reciprocal
of the highest modulation frequency fm of the rf applied voltage. For example, in televi-
sion transmission the video modulation may contain frequencies up to several megahertz,
so that 1/fm may be less than a microsecond, and transient effects in a 100-meter line
length would be of consequence. (The picture quality would be degraded. The remedy
would be to eliminate the reflections at the load by methods that will be described in
sec. 2–3.)
Z cos β ( l − x ) + jZ 0 sin β (l − x )
V ( x ) = Vi L (2–15)
Z L cos βl + jZ 0 sin βl
Vi Z 0 cos β ( l − x ) + jZ L sin β ( l − x )
I(x) = (2–16)
Z 0 Z L cos βl + jZ 0 sin βl
The input impedance, Zi, may be obtained dividing V(0) by I(0), as given by (2–15) and
(2–16) for x = 0. The resulting equation is
Z cos βl + jZ 0 sin βl
Zi = Z 0 L (2–17)
Z 0 cos βl + jZ L sin βl
This equation is often written in a modified form, obtained by dividing through both the
numerator and denominator of (2–17) by Z0 cos bl, which gives
( Z Z ) + j tan βl
Zi = Z 0 L 0 (2–18)
1 + j ( Z L Z 0 ) tan βl
The definitions of the quantities in these equations are listed below for ready
reference:
Transmission-Line Equations 55
Z L = RL + jX L (2–19)
1 + j tan βl
Zi = Z 0 = Z0 (2–20)
1 + j tan βl
That is, when the load impedance is equal to the characteristic impedance, the input
impedance is also equal to this value. The line then acts as a one-to-one transformer.
This equality of the load and characteristic impedances is called the matched load condi-
tion. The load impedance is said to be matched to the characteristic impedance of the line.
The one-to-one transformation of impedance results regardless of the length of the line, l.
* The reader is assumed to have a basic familiarity with the concept of complex numbers and
complex-number algebra. For reference and review, Appendix C covers the elements of this subject.
†
The subscript L here stands for “load.” The symbol XL often stands for “inductive reactance,” as opposed
to capacitive reactance XC. But here XL denotes the load reactance, which may be either inductive or
capacitive.
56 Transmission Lines
Moreover, if the condition ZL = Z0 is also substituted into (2–15) and (2–16), it is found
that the amplitudes of V(x) and I(x) remain constant for all values of x, that is, everywhere
on the line. Only the phases change, as would be expected from (2–3) and (2–8); in fact,
these equations of the infinite line now apply to the finite line (if and only if ZL = Z0). This
property is of great importance, for it provides a method of eliminating reflected waves
and transient effects. Since there is no standing wave, with maxima and minima, such a
line is called a flat line. It is important to note that since Z0 for a lossless line is an entirely
real (resistive) impedance, the matched-load condition requires that the load impedance
be nonreactive (purely resistive) as well as have the magnitude of Z0.
(2) l = l/2. (Line length equal to one-half wavelength.) For this case, the angle
bl that occurs in (2–18) becomes equal to p radians or 180 degrees. Since
the tangent of this angle is zero, the equation becomes
ZL
Zi = Z 0 = ZL (2–21)
Z0
But there are two respects in which the half-wavelength one-to-one transformer is infe-
rior to the matched line. First, there is a standing wave on the half-wavelength (or integral-
half-wavelength) line, which means there are reflections and transient effects. Second,
since the one-to-one transformation depends on the fact that l/l has a particular value, the
line will have this property only at one frequency and its integral multiples. If it is desired
to operate over a wide range of frequency without readjustment of the line length, the
matched-load line should be used. Nevertheless, the half-wavelength line is useful as a
one-to-one impedance and voltage transformer when its limitations are not of concern.
(3) l = l /4. (Line length equal to one-quarter wavelength.) In this case the angle
bl in (2–17) becomes equal to p/2, or 90 degrees. Note that (2–17) rather
than (2–18) is used because the tangent of 90 degrees is infinite, and conse-
quently (2–18) leads to difficulties that require careful handling.) Since
cos 90° = 0 and sin 90° = 1, (2–17) becomes
jZ Z 2
Zi = Z 0 0 = 0 (2–22)
jZ L Z L
The quarter-wavelength line transforms a load impedance ZL that is smaller than Z0 into
a value Zi that is larger than Z0; and vice versa. It is sometimes called an impedance
inverter for this reason. If ZL is resistive, Zi will also be resistive (real). Thus the quarter-
wavelength transformer is useful when it is desired to transform a resistive impedance
into a different resistive value, either larger or smaller. The desired transformation is
accomplished by choosing the appropriate value of Z0 in accordance with (2–22). The
procedure for obtaining a desired value of Z0 is described in sec. 2–4; it depends on the
ratio of the spacing of the line conductors, s in Fig. 2–1, to their diameter d. Since these
physical dimensions cannot practically have an infinite range of values, the transforma-
tion ratio of the quarter-wave transformer is subject to practical limitation. Nevertheless,
Transmission-Line Equations 57
it is a very useful device because of its simplicity and the ease with which its behavior
is calculated. Lengths l equal to odd-number multiples of l/4 will have the same trans-
formation property, but as the length is made longer the sensitivity to a small change of
frequency becomes greater.
The voltage transformation of the quarter-wave transformer is found from (2–15) by
taking x = l, and as before bl = 90°. The result is
ZL
VL = − jVi (2–23)
Z0
which shows that when ZL is real the amplitude is changed by the factor ZL /Z0 and the
phase is changed by 90 degrees. (It should be noted here that (2–23) describes the trans-
formation of input voltage to output voltage, whereas (2–22) describes the transformation
of output impedance to input impedance, since these are the usual directions of interest.)
A transformation of the current also takes place, but in the inverse ratio, so that the
product of voltage and current is unchanged.
In both cases the input impedance Zi is purely reactive (imaginary). This means that the
input terminals of such lines “look like” the terminals of either an inductance or a capaci-
tance, depending on the length of the line. If the length is between zero and a quarter
wavelength, so that bl is an angle between zero and 90 degrees, the open-circuited line
will be capacitive, a negative reactance in accordance with (2–24). In contrast, the short-
circuited line of the same length will be inductive, following (2–25). For lengths between
a quarter and a half wavelength, bl lies between 90 and 180 degrees, and both the tangent
and cotangent are negative; therefore the open-circuited line becomes inductive, and the
shorted line becomes capacitive. For each successive quarter wave increase of length,
their behaviors interchange in this way. Such line sections may therefore be utilized as
elements of inductance-capacitance circuits; and they are so used, especially at the higher
frequencies (shorter wavelengths) where the required line lengths are not great.
Open-circuited and short-circuited lines can also behave like resonant circuits. This
behavior occurs when the length of the line is an integral multiple of a quarter wave-
length; then tan bl → ∞, and cot bl = 0. The quarter-wave short-circuited line presents
an infinite impedance at its input terminals, like a parallel-resonant LC circuit; and the
58 Transmission Lines
quarter-wave open-end line presents a zero impedance at its input terminals, like a series-
resonant LC circuit. These behaviors are interchanged for each successive quarter-wave
increase of the line lengths. Transmission-line resonant circuits have many applications
in high-frequency radio circuitry and in antenna design.
1
Z= (2–26)
Y
That is, they are reciprocals of each other. This equation holds when both Z and Y are
complex, as well as for purely resistive or reactive cases. That is, in general,
Z = R + jX (2–27)
and
Y = G + jB (2–28)
with the understanding that R and X refer to resistance and reactance in series, and G
and B refer to conductance and susceptance in parallel.
The restricted nature of the reciprocity between conductance and resistance, and
between susceptance and reactance, may be understood in terms of the diagrams of Fig.
2–4. In Fig. 2–4a, it is evident that Z = R + j0. Substituting this value for Z in (2–26),
and also applying (2–28), immediately gives Y = 1/R = G + j0. Thus in this simple case,
G = 1/R. In other words, the reciprocal relationship between G and R holds for a purely
resistive (conductive) circuit or branch of a circuit.
In Fig. 2–4b, the impedance is Z = jX (purely reactive). (The reactance in this diagram
is shown as a box, since for the purposes
of this discussion it may be either an
R1
R X G2 B2 inductance or a capacitance, or any com-
(G) (B) (R2) (X2)
X1 bination of the two.) If this value of Z is
substituted in (2–26), the result is Y = 1/jX
(a) (b) (c) (d) = −j(1/X). If this result is compared with
FIGURE 2–4. (2–28), it is evident that B = −1/X. Thus
Various two-terminal arrangements of for a purely reactive circuit, or branch of
resistance, reactance, conductance, and a circuit, the susceptance is the negative
susceptance. reciprocal of the reactance.
Transmission-Line Equations 59
1 1 1 R1 − jX1 R1 X
Y1 = = = = 2 2
− j 2 1 2 (2–29)
Z1 R1 + jX1 R1 + jX1 R1 − jX1 R1 + X1 R1 + X1
Again comparison with (2–28) makes it evident that for this circuit:
R1
G1 = (2–30)
R12 + X12
and
− X1
B1 = (2–31)
R12+ X12
On the other hand, for the parallel circuit of Fig. 2–4d, the corresponding relations are
simply G2 = 1/R2 and B2 = −1/X2. It is therefore apparent that the circuits of Fig. 2–4c
and 2–4d are equivalent if
R12 + X12 1
R2 = = (2–32)
R1 G2
and
R12 + X12 1
X2 = = (2–33)
X1 B2
“Equivalent” means that if the same voltage is applied to the terminals of either circuit,
the current that flows at the terminals will be exactly the same for both circuits. (Of
course, this will in general hold—for fixed values of inductance and capacitance—at only
one frequency.) This equivalence is of great importance and utility in impedance-
matching calculations. It means that the impedance at the input terminals of a transmis-
sion line, for example, can be regarded as either a series combination or a parallel
combination of resistance and reactance, whichever happens to be most convenient.
Equation (2–18) for the input impedance of a line can be converted, by applying (2–26)
to it, into an equation for the input admittance, Yi, of the line, where Yi = 1/Zi. The
equation is
1 + j (Y0 YL ) tan βl
Yi = Y0 (2–34)
(Y0 YL ) + j tan βl
60 Transmission Lines
The symbol Y0 denotes the characteristic admittance of the line, which is simply the
reciprocal of the characteristic impedance; that is, Y0 = 1/Z0. The symbol YL denotes the
load admittance, which is equal to 1/ZL.
The equations for open-circuited and short-circuited lines, (2–24) and (2–25), can also
be expressed in terms of the input susceptances. For the open-circuit termination, the
equation corresponding to (2–24) is
Bi = Y0 tan βl (2–35)
Equations (2–35) and (2–36) can be obtained either by making the substitution Xi = −1/Bi
in (2–24) and (2–25), or by letting YL be zero for the open-circuit case and infinite for
the short-circuit case in (2–34.)
where |r| is the magnitude (modulus) of the complex quantity and f is the phase angle.
The magnitude of the reflection coefficient is the ratio of the amplitudes (or rms values)
of the reflected and incident voltage (or current):
V0(r )
r = (2–38)
V0(i )
The phase angle is the phase difference between the phases of the incident and reflected
waves relative to an arbitrary reference phase:
φ = φi − φr (2–39)
(Since the phase of the incident wave is the usual reference, ordinarily fi = 0, and
f = −fr.)
The reflection coefficient is determined entirely by the relationship of the load
impedance, ZL, to the characteristic impedance of the line, Z0. Analysis shows that the
relationship is
Transmission-Line Equations 61
Z L − Z0 ( Z L Z0 ) − 1
r= = (2–40)
Z L + Z0 ( Z L Z0 ) + 1
At other points the two voltage waves will be exactly out of phase and will therefore
subtract, resulting in a voltage minimum Vmin given by
the ratio of the maximum to the minimum voltage is called the voltage standing wave
ratio, VSWR, that is
Obviously the VSWR is a number equal to or greater than one. (It is equal to one when
there is no reflected wave, or V0(r) = 0. From (2–38) and (2–40) it is apparent that
VSWR = 1 when |r| = 0 and therefore when ZL = Z0.)
62 Transmission Lines
Z
V ( d ) = VL cos β d + j 0 sin β d (2–44)
ZL
The magnitude (modulus) of V(d ), in terms of ZL /Z0 = RL /Z0 + jXL /Z0, with RL /Z0
abbreviated to R and XL /Z0 abbreviated to X, is
2 2
X R
V ( d ) = VL cos β d + 2 sin β d + 2 sin β d (2–45)
R +X
2
R + X
2
Typical plots of |V(d)| are shown in Fig. 2–5 for a number of values of RL /Z0 and
XL /Z0.
Noteworthy features of these plots are
Voltage Current
the following. (1) The maxima are sepa-
RL = Z0 /2 rated by l/2 and the minima are separated
XL = 0
by l /2; the distance from a minimum to
Vmin Vmax
a maximum is l /4. (2) The minima are
more sharply defined than the maxima;
RL = 2Z0
XL = 0
consequently, in making standing wave
position measurements it is better to
dm measure the position of a minimum than
RL ≠ 0, ∞
the position of a maximum (greater accu-
XL ≠ 0, ∞ racy is possible). (3) When a resistive
(nonreactive) load has a value less than
dm
Z0, a voltage minimum is located at the
|V (d )|
ZL = 0 load terminals. (4) When a purely resis-
tive load has a value greater than Z0, a
dm voltage maximum is located at the load
terminals. (5) When the load is complex
ZL → ∞
(partially or wholly reactive), the voltage
9l/4 2l 7l/4 3l/2 5l/4 l 3l/4 l/2 l/4 0
at the load is neither a minimum nor a
d maximum. (6) When the load is either
FIGURE 2–5. zero or infinite, the standing-wave ratio is
Standing-wave voltage patterns for several infinite (since Vmin = 0) and the minima
ZL/Z0 conditions. have the form of cusps.
Transmission-Line Equations 63
1+ r
Z L = Z 0
1 − r
(2–46)
The relationship of the magnitude of the reflection coefficient, |r|, to the standing-wave
ratio VSWR has been shown indirectly through (2–38) and (2–43). From these equations
it may be deduced that
VSWR − 1
r = (2–47)
VSWR + 1
Thus a measurement of the VSWR provides one ingredient necessary for determination
of ZL, in accordance with (2–46). The additional ingredient required is of course the phase
angle f of (2–37).
It can be shown that f is related to the distance dm, which as shown in Fig. 2–5 is the
distance from the load terminals to the first voltage minimum. The relation is
φ = π − 2β dm = π 1 − m radians
4d
λ (2–48)
= 180 1 − m degrees
4d
λ
Therefore, if the two quantities VSWR and dm are measured, ZL can be calculated,
through (2–46), (2–37), (2–47), and (2–48). Moreover, Zi can also be calculated through
(2–18), either by first calculating ZL or directly by substituting (2–46) into (2–18), which
gives
1 + r + j tan βl
1− r
Zi = Z 0 (2–49)
1 + j 1 + r tan βl
1− r
In practical work (2–46) and (2–49) are solved, including the intermediate steps of (2–47)
and (2–48), by using a Smith Chart, as described in sec. 9.12.
64 Transmission Lines
2.2.7. Attenuation
The results presented thus far have all been based on the assumption that the line is loss-
less, which means that R1 in Fig. 2–3 is zero and R2 is infinite. In real transmission lines
these assumptions are of course not perfectly justified; there are some losses. When the
loss of a line is considerable, even the basic form of the transmission line equations is
considerably modified. However, when the losses are fairly small, as is usually the case,
the equations that have been given can be used for all purposes except computing the
power ultimately delivered to the load. This power will be slightly less than the power
delivered to the line at its input terminals. The power loss in decibels is given by multi-
plying the length of the line by a decibel attenuation constant. For most manufactured
lines, the value of this constant is published in the manufacturer’s sales literature.
Additionally, a useful collection of data on transmission lines and waveguides is
contained in Lowman and Simons (2007).
The attenuation constant (sec. 1.1.9) can also be expressed in terms of the conductor
resistance R per unit length and the leakage conductance G per unit length. Ordinarily,
transmission lines used with antennas are of the low-loss type. For these lines, the
attenuation constant, when expressed in terms of decibels per meter (dB/m), is
R G
α dB = 8.686 +
2 Z 0 2Y0
dB m (2–50)
Here Z0 is, as usual, the characteristic impedance of the line and Y0 is the characteristic
admittance (= 1/Z0). In terms of Fig. 2–3, R = R1 and G = 1/R2. When the line attenuation
is appreciable, a more complicated formula must be used. “Appreciable attenuation,” for
practical purposes, may be taken to mean more than about one decibel per wavelength
of the line, insofar as the validity of (2–50) is concerned.
The equations for transmission-line input impedance, voltage, and current, (2–15)
through (2–18), and subsequent equations derived from them, are accurate for most
antenna applications. However, this is not so if the total attenuation* of the line (adBl)
is greater than about a decibel. Attenuation reduces the amplitude of the reflected
wave as it travels back toward the line input, and the standing-wave ratio consequently
diminishes toward the input. Thus, with large attenuation, the standing-wave ratio as
calculated from equation 2–47 is valid only in the vicinity of the load end of the line.
Additionally, if the attenuation is great enough the input impedance of the line will be
essentially equal to the characteristic impedance regardless of the reflection coefficient
of the load. More detailed treatments of transmission lines give equations for high attenu-
ation, analogous to (2–15) through (2–18) (see, e.g., Lowman and Simons 2007,
p. 51–5).
* Total attenuation from ohmic loss is adBl without a standing wave. However, e.g., with VSWR = 2, the
attenuation is 1.25 times adBl, because of increased mean-squared voltage across the line (Ragan 1948,
pp. 30–31).
Impedance Matching and Power Division 65
If now it can be arranged that Bs = −By, these two susceptances in (2–51) will cancel
each other, leaving
Yp = Gy = Y0 (2–52)
This admittance presents a matched load to the main line at point P, and so there will be
no standing wave to the left of this point, that is, between point P and the line input. This
is the desired result.
Impedance Matching and Power Division 67
The susceptance of the stub, Bs, is computed from (2–36). By making s a suitable
length, it is demonstrable that any needed value of Bs can be achieved. Therefore, if the
two lengths y and s can be varied at will, the desired transformation of impedance
can be achieved. In practice the position and length of the stub are roughly determined
by Smith Chart calculations based on standing-wave measurements. Then the lengths
are experimentally adjusted to minimize the standing wave ratio (VSWR) on the
main line.
When it is not practical or advisable to provide a movable stub (one for which the
distance y can be varied), the same result can be accomplished by means of two fixed-
position stubs separated a suitable distance, usually between one-fourth and three-eighths
of a wavelength. The analysis of the impedance transformation proceeds on the same
principles, though it is of course somewhat more involved.
Y0( M ) = YA + YB (2–53)
If each branch line is terminated in its own characteristic admittance, Y0(A) and Y0(B), as
will usually be true, then of course YA and YB will be equal to the characteristic values.
Since the branch lines are in parallel, the voltage applied to each of them at the junc-
tion point is the same and is equal to the voltage output of the main line, VM. Therefore,
the power delivered to each branch is
PA = VM2 G A (2–54)
hA
nc ad
PB = VM2 GB (2–55) Bra Lo
Main line YA
Line YM
and the total power is input Br
YB an
Branch point ch
Ptotal = PA + PB = VM2 ( G A + GB ) (2–56) Lo
ad
B
division property of the junction. That is, the fraction of the total power delivered to each
branch will be
PA GA
FA = = (2–57)
Ptotal G A + GB
PB GB
FB = = (2–58)
Ptotal G A + GB
Y0( A)
FA = (2–60)
Y0( M )
Y0( B )
FB = (2–61)
Y0( M )
It would be possible to divide the power at a junction among more than two branching
lines. The analysis of the power division would be a simple extension of the foregoing
analysis; the fraction of power delivered to each branch would be equal to the conduc-
tance component of the input admittance of that branch, divided by the sum of the con-
ductances of all the branches. The conductances in the usual cases would be equal to the
characteristic admittances of the branches. It is not customary to divide power in this
way, however, partly because the required characteristic impedances of the branch
lines become impractically large and partly because there are mechanical construction
difficulties, especially with lines of the coaxial type.
When multiple power division is required, it is more common to employ one of the
successively branching line structures shown schematically in Fig. 2–8. In these diagrams
the two conductors of the lines are shown as single lines, for the purpose is solely to
show the branching arrangements.
In each of these arrangements the ultimate power division is accomplished by a
number of successive two-branch divisions. The required division of the total power to
the ultimate loads results in a system of simple simultaneous equations for the division
ratio required at each branch; this ratio is then achieved by the method that has been
described. Although eight loads are shown in Fig. 2–8, the methods are not restricted to
this number.
Forms of Transmission Lines 69
Eight loads
the voltages and currents have the same 3
4
phase at each load. This is readily achieved Input
5
in the arrangement of Fig. 2–8a, by
6
keeping the lengths of the parallel branches Branch point
7
equal. In Fig. 2–8b it would be necessary
8
to make the lengths denoted by d equal to (a)
a wavelength or an integral number of
Branch point
wavelengths; or d may be a half wave- d
length if the line is of the two-wire bal- Input
anced type and if the line is “turned over”
in each half-wavelength interval to reverse 1 2 3 4 5 6 7 8
Eight loads
the polarity of the voltage. (This cannot (b)
be done, of course, with a coaxial or other FIGURE 2–8.
unbalanced line type).
Alternative arrangements for division of power
It may also be necessary to have among multiple loads.
stepdown impedance transformers in the
branch lines to keep the characteristic
impedances required in the successive branches from building up to too high a value; in
short, there are numerous practical problems. Here the purpose has been primarily to
present basic principles.
It is possible also to have two lines branching with a series connection instead of par-
allel, although such an arrangement is less common. The analysis would be made in a
similar way, except that it would be more convenient to work with the impedances than
with the admittances.
Insulating
spacers
Two-wire line
Insulating
“wafers”
Coaxial line
a b
Strip line
FIGURE 2–9.
Perspective sketches and transverse sections of two-wire, coaxial, and strip transmission lines.
Coaxial lines confine the fields entirely to the space between the inner and outer con-
centric conductors; therefore there is no radiation problem. Being unbalanced, these lines
require special balanced-to-unbalanced transformers (called “baluns”) to connect them
to any balanced source or load. However, sources and loads are in some cases unbalanced,
so that the unbalanced character of the line is as often an advantage as a disadvantage.
Coaxial lines have the advantage of permitting the interior to be completely sealed against
weather and contamination. The internal space may be pressurized to increase voltage
breakdown and to resist intrusion of moisture. Some coaxial lines are made with a flexible
braid outer conductor and flexible solid-dielectric material between the conductors. Such
lines are flexible and compact and find many uses at low and intermediate power levels.
For very high-power transmitting applications, coaxial lines employ solid copper pipe,
with outer-conductor diameters up to 9 or more inches.
Forms of Transmission Lines 71
The strip line is a modification of the two-wire line in which the flat faces of the
adjacent conductor surfaces are wider than the spacing between the conductors. Accord-
ingly the field is virtually confined to the space between the conductors, and there is very
little radiation. The two strips may be alike (balanced), or one of them may be a large
flat (grounded) plate. Often a solid dielectric material is used as a spacing material. It is
especially convenient for some purposes, usually at low power levels.
log ohms
276 2s
Z0 = (2–62)
εr d
where s and d are the center-to-center conductor spacing and the conductor diameter
respectively, as shown in Fig. 2–9. The relative permittivity er is of the medium between
and around the conductors (er = 1 for air or vacuum). Typical values of Z0 for two-wire
lines range from about 200 to 800 ohms.
For a coaxial line the characteristic impedance is given by
log ohms
138 b
Z0 = (2–63)
εr a
with b the inner diameter of the outer conductor and a the outer diameter of the inner
conductor. The relative permittivity er is for the material between the conductors. Typical
values of Z0 range from about 20 to 100 ohms.
For the strip line the following formula is good as long as the dimension b (Fig. 2–9)
is much smaller than the dimension a:
377 b
Z0 = ohms (2–64)
εr a
The typical range of characteristic impedances is about the same as for coaxial lines.
Numerous variations of these basic forms are sometimes used; however, there is a
practical upper-frequency limit for any given type of line. Usually the limit is set either
by the losses in dielectric material or by the spacing of the conductors in relation to the
wavelength. When this spacing is an appreciable fraction of a wavelength, the open-wire
lines will radiate excessively. Coaxial and strip lines will act like waveguides, as will be
discussed in sec. 2–5, if the conductor spacing becomes comparable to or greater than a
half wavelength.
A basic property assumed for all the lines discussed is uniformity, that is, constant
spacings and conductor diameters. At joints, elbow turns, couplings, and so forth, there
72 Transmission Lines
may be discontinuities that cause reflections that can be troublesome if care is not taken
to minimize them. For broad-band impedance-transforming applications deliberately
nonuniform tapered lines are used. These lines have a characteristic impedance that
changes as the conductor diameters and spacings change, so that a gradual impedance
transformation takes place.
2.5. Waveguides
At frequencies too high for successful operation of conventional transmission lines, rf
power may be transmitted by means of waveguides. A waveguide in its commonest form
is a hollow pipe, sometimes circular but more often rectangular in cross section, and
occasionally of some other form. The dimensions of the cross section are such that an
electromagnetic wave can propagate in the interior of the guide. Being confined by the
walls of the guide, the field does not spread spherically as it would in free space, so there
is no inverse-square-law decrease of the power density. There is some attenuation due
to currents in the walls, but generally this loss is less than in a coaxial line of comparable
size. Moreover, since no insulating supports are required, there is virtually no “leakage”
loss and less tendency to flashover at high voltages.
Obviously the analysis of waveguide behavior must be made in terms of electromag-
netic field concepts rather than in terms of an equivalent-circuit representation. The
starting point is Maxwell’s equations, and their solution requires advanced mathematical
methods. Yet the results are that waveguides behave essentially like the conventional
transmission lines that have been described, although there are some important
differences.
v ph = f λ (2–65)
might be transmitted through a long guide of length l, then reflected back to the sending
end, and the elapsed time t might be measured. The velocity of the pulse (group velocity)
is given by the formula
2l
vgr = (2–66)
t
(The factor 2 occurs because the total distance traveled is twice the length of the
guide.)
If these two measurements are made at the same frequency in a waveguide, it will be
found that in general the two velocities are not the same. At some frequencies they will
be nearly the same; at others they will be considerably different. The distinction between
these two velocities was not necessary for transmission lines because in conventional
lines the group and the phase velocities are equal.
To explain why these velocities are different in waveguides requires a detailed analy-
sis, which cannot be given here. It must suffice to define them and to state, at appropriate
points, which velocity is meant. Unless otherwise stated, the phase velocity will always
be meant; it is the one of importance whenever considerations of wavelength are involved.
Yet obviously the group velocity is also of great importance in other applications—radar
signal transmission, for example. In general it is the velocity at which signals of any
kind—that is, intelligence—are propagated; it is also the velocity at which energy is
propagated.
The phase velocity is always either equal to or greater than the group velocity. The
following relationship holds between them in evacuated or gas-filled waveguides:
That is, the product of the two velocities is equal to the square of the free-space propaga-
tion velocity, 3 × 108 meters per second. This means that Vph is always greater than the
free-space velocity and Vgr is always less, unless both are equal to c.
It is a principle of modern physics that no form of matter or energy can travel with a
velocity greater than the free-space velocity of light (electromagnetic waves). However,
this principle is not violated by phase velocities, because it is the group velocity that
represents the velocity of propagation of energy.
Since the phase velocity in waveguide is greater than the velocity in free space, the
wavelength at a given frequency will be greater in the guide than in free space. Using
the conventional symbol c for the velocity in free space, the relationship of the
wavelength in the guide, lg, to the wavelength in free space, l0, is:
v ph
λ g = λ0
c
(2–68)
The distinction between the guide wavelength and the free-space wavelength is an
important principle of waveguide theory.
74 Transmission Lines
c
λg = (2–69)
f − fc2
2
where c is the free-space propagation velocity. This can also be written in another
form:
λ0
λg = (2–70)
1 − ( fc f )
2
Comparison of this equation with (2–68) shows that the phase velocity is given by
λg c
v ph = c = (2–71)
λ0 1 − ( fc f )
2
It is evident from this expression that if f becomes less than fc, the phase velocity becomes
imaginary (in the complex-number sense); physically this means that the wave is not
propagated. Equations (2–69) and (2–71) also show that as the frequency f approaches
the cutoff frequency fc, the phase velocity and the guide wavelength become infinite.
Consequently, the group velocity tends to zero, according to (2–67).
The foregoing results have all presupposed that the waveguide is filled with air (or a
vacuum) or any gas having virtual unity relative permittivity. Waveguides are seldom
filled with solid material, but, if such a low-loss material were used, the foregoing
equations still hold if c is replaced by c ε r , where er is the relative permittivity of the
material in the guide.
c
fc = (2–72)
2a
Waveguides 75
or
λc = 2 a (2–73)
377 λg
Zc = = 377 ohms (2–74)
1 − ( fc f)
2 λ0
In general Zc is greater than 377 ohms. At the cutoff frequency fc it becomes infinite, and
at f = 2fc it has the value 435 ohms, as found from (2–74).
The quantity Zc has the same general significance for waveguide as does the charac-
teristic impedance Z0 for ordinary transmission lines, but there are certain exceptions to
this statement. As shown (see Fig. 2–10) by (2–72), fc is determined by only the dimen-
sion a. Therefore two waveguides with the same a dimension and different b dimensions
will have the same value of fc and also the same value of Zc. But if these two waveguides
are joined together end to end, so that a wave propagating in one of them passes through
the junction into the other, a reflection will be set up at this point due to the disparity of
dimensions, in spite of the impedances being “matched.”
But for constant guide dimensions and a single mode of propagation, the characteristic
wave impedance of the guide, Zc, has the same significance as the characteristic impedance
of a two-conductor transmission line, with respect to reflection and standing waves. The
same principles of impedance matching and transformation apply, and the transmission-
line equations, (2–15) through (2–18) and beyond, can be used provided that the guide
wavelength, lg, is used in calculating b when using equation (1–32), ch. 1.
Inducitive Capacitive
iris iris
b b
a a
FIGURE 2–13.
Inductive and capacitive irises (as they would appear if the waveguide walls were transparent).
may be employed in the manner described for transmission-line stubs. Reactance may
also be introduced into waveguides by means of irises, posts, and tuning screws. Their
roles in impedance transformation and matching are the same as were described for the
transmission-line tuning stub of Fig. 2–6; that is, the correct position of the reactance
with respect to the load and the amount of reactance (susceptance) needed are determined
in the same way.
Figure 2–13 illustrates capacitive and inductive irises in a rectangular waveguide. They
behave like susceptances, that is, shunt reactances. As shown, they consist of thin metallic
plates perpendicular to the guide walls and joined to them at the edges, with an opening
between them. When the opening is parallel to the narrow walls of the guide, the suscep-
tance is inductive; when it is parallel to the wide walls, it is capacitive. The amount of
the susceptance is determined by the size of the opening. Many variations of these
configurations may be used, for example, an unsymmetrical iris or a unilateral one.
A post placed across the narrow dimension of the guide acts as an inductive shunt
susceptance, of a value depending on its diameter and its position in the transverse
plane. A tuning screw projecting part way across the narrow guide dimension acts like
a capacitive susceptance and may be made adjustable. These devices are illustrated in
Fig. 2–14.
Tuning
screw
Post
b b
a a
FIGURE 2–14.
Post and tuning screw in waveguide.
other. The various dimensions labeled l/4 are approximate; in practice, they are experi-
mentally adjusted for best operation. Note, however, that sometimes the approximate
dimension is l0/4 and in others lg/4. The couplers illustrated are representative of the
methods that may be employed; there are many variations of them in use.
Quarter-wave Straight-through
probe coupler coupler
Waveguide Waveguide
l0 /4
b b
l0 /4
lg /4
lg /4
Cross-bar coupler
Waveguide
Waveguide
a lg /4
End view Side view
FIGURE 2–15.
Coaxial-line-to-waveguide couplers.
80 Transmission Lines
B B
A A
C C
B
B
a A C b
A C
a
b
Channel 2–4
Coaxial line
Port 2 Port 4
Coupling loop
Port 1 Port 3
no output will occur when it flows the other way. The ratio of these outputs for equal
power flow in the two directions of the main channel is called the directivity factor. A
typical value is 40 dB.
Figure 2–19 shows one form of directional coupler using coaxial line sections. The
coupling between the two channels is by means by a short loop of the inner conductor
of the 2–4 channel projecting into the interconductor space of the 1–3 channel. The plane
of the loop is parallel to the inner conductor of the line into which it projects. It couples
to both the electric and the magnetic field components in such a way that the currents
that each type of field induces reinforce in the coupler branch leading to port 2 and cancel
in the branch leading to port 4 (when the flow in the 1–3 channel is from port 1 to port
3). Many other coupling mechanisms exist for achieving a similar result.
In practice, directional couplers are often constructed with a matched-load termination
on one port, such as port 4 in the above example, and only ports 1, 2, and 3 are made
accessible. Then when there is a reflected wave in the 1–3 channel, which results in an
output to port 4, the power thus received is absorbed in the load. (Otherwise it could be
reflected and reach port 2, which would be harmful to the intended operation of the
coupler.) Thus it might be considered that the coupler is not virtually lossless, as assumed.
However, this load, though “built in,” is not truly a part of the coupler itself, so that it
does not invalidate the assumption of losslessness.
The principal uses of directional couplers in antenna applications are for measurement
and monitoring of the power flow and the VSWR, as discussed in secs. 9.8.3 and
9.11.1.
References
Lowman, R. V., and R. N. Simons, “Transmission Lines and Waveguides,” ch. 51 in
Antenna Engineering Handbook, 4th ed., J. L. Volakis (ed.), McGraw-Hill, 2007.
Ragan, G. L. (ed.), Microwave Transmission Circuits, McGraw-Hill, 1948.
84 Transmission Lines
7. If in Problem 6 the load is not purely resistive, suppose that it is found (by sliding
the voltmeter along the line) that a voltage minimum exists at a distance
dm = (5/24)l from the load terminals. (The VSWR is still 4.) (a) What is the
(complex) reflection coefficient, r? (b) From equation (2–46), what is the value
of the load impedance, ZL?
8. The conductors of a transmission line have a series resistance of 0.01 ohm per
meter of line and a leakage conductance of 10−5 mho per meter. (a) What is the
attenuation, in decibels, of a 1,000-meter length of this line, assuming that
Z0 = 50 ohms and that the wavelength is less than 10 meters? (b) If a transmitter
delivers 1,000 watts of rf power at the input terminals of this line, and if an
antenna connected at the load terminals has a nonreactive impedance equal to
the characteristic impedance of the line, how much power is delivered to the
antenna?
9. A low-loss transmission line system connects a single source of power to eight
loads in the manner of Fig. 2–8a. The loads are resistive (nonreactive) and are
each 400 ohms in value. If the power divides equally at each branch point of the
system, and if all line sections are completely free of standing waves, what must
the characteristic impedance of the input line be? (Assume that no impedance
transformers are used, so that the impedance [admittance] presented at each
branch point depends solely on the characteristic impedances [admittances] of
the branching lines.)
10. A rectangular air-filled waveguide has lateral dimensions a = 7.2 cm, b = 3.4 cm.
(a) What is the cutoff frequency, fc? (b) At an operating frequency of
f = 3 × 109 Hz, for which the free-space wavelength is l0 = 10 cm, what is the
guide wavelength, lg?
CHAPTER 3
Antenna Parameters
87
88 Antenna Parameters
some of which will be mentioned throughout the book in connection with related discus-
sions of electromagnetic behavior. It is desirable, however, for the reader to have some
general “feeling” for the structural factors in the design of antennas, as an aid in relating
the somewhat abstruse electromagnetic concepts to real physical structures. A thumbnail
sketch of some of these aspects of antenna design will be given here, before the defini-
tion of electromagnetic parameters is taken up.
3.1.2. Supports
As mentioned in sec. 1–3, an antenna must be located “in the clear” for good results—
away from large conducting or absorbing objects. Accordingly there must often be some
supporting structure to place the radiating element or elements in a clear location (which
often is synonymous with a high location). Antennas are supported by such devices as
towers, masts, and pedestals. Towers are used when great height is required. Masts may
be quite high, but they are often as short as a few feet. Pedestals are the base structures
of antennas such as reflectors and lenses, for which height is not so important as strength.
Sometimes an antenna may be mounted directly on a vehicle, such as an automobile,
ship, aircraft, or spacecraft; no intermediate “support” is required. Moreover, towers and
masts are sometimes themselves used as antennas rather than as supports. In the standard
Antenna Structures 89
broadcast band (535–1,705 kHz), for example, vertical towers of heights up to several
hundred feet are used as transmitting antennas.
3.1.4. Conductors
Metals are the usual conducting materials of antennas. Metals of high conductivity, such
as copper and aluminum (and its alloys), are naturally preferred. Brass may be used for
machined parts. Magnesium is sometimes used where ultralight weight is important,
usually in an alloy and with a protective coating or treatment. Where strength is
of primary importance, steel may be used, either with or without a coating or plating of
copper. The conductivity of unplated steel is adequate when it is used in the form of
90 Antenna Parameters
sheets or other large-surface-area forms (as for the surface of a paraboloidal reflector).
Antenna wire is sometimes made with a steel core for strength and to minimize stretch-
ing, and with a copper coating to increase the conductivity. Such wire is virtually as good
a conductor as solid copper, since rf currents are concentrated near the surfaces of con-
ductors (skin effect). For this reason brass and other metals are sometimes silver plated
when exceptionally high conductivity is required. For the same reason large-diameter
conductors may be hollow tubes without loss of conductivity. At low radio frequencies
the conductivity of large-diameter conductors may be increased, compared to a solid
conductor, by interweaving strands of small-diameter insulated wires; the resulting con-
ductor is called Litz wire. This technique is most effective below about 500 kHz.
At higher frequencies it is not effective because the currents tend to flow only in the
outer strands.
Conductor size in antenna design is determined by many factors, principally the per-
missible ohmic losses and resultant heating effects in some cases, mechanical strength
requirements, permissible weight, electrical inductance and capacitance effects, and
corona considerations in high-voltage portions of transmitting antennas. Corona is mini-
mized by large-diameter conductors, by avoidance of sharp or highly curved edges, and
by using insulators with metal end caps bonded to the insulating material, so that small
air gaps between wires and insulators do not exist. Corona can occur on metal supports
of the antenna as well as on the antenna conductor itself, as a result of induced
voltages.
3.1.5. Insulators
The conducting portions of an antenna not only carry rf currents but also have rf voltages
between their different parts and between the conductors and ground. To avoid “short
circuiting” these voltage, insulators must sometimes be used between the antenna and its
supports, or between different parts of the antenna. Insulators are also used as spacer
supports for two-wire and coaxial lines and to “break up” guy wires used with masts and
towers so that resonant or near-resonant lengths will not occur. The maximum permissible
uninterrupted length of guy wire sections is about 1/8 wavelength. (A half-wavelength
section would be resonant.) Insulators used to support long heavy spans of wire must be
of high strength. Typical insulating materials for such insulators are glass and ceramics.
Other “low-loss” materials such as polystyrene and other plastics are used where less
strength is required. Very large and heavy insulators are necessary in high-power trans-
mitting applications to prevent flashover. Coaxial lines and waveguides in high power
applications may be filled with an inert gas, or dry air, at a pressure of several atmo-
spheres, to increase the voltage-breakdown level. Printed circuit boards rely on low loss
dielectric insulators.
ice loads is primarily a matter of mechanical strength and bracing. Guy wires are used
with tall structures or towers to prevent their overturning in high winds. At very low rf
frequencies, where long spans of wire may be used, ice is sometimes melted from the
wires by passing heavy 60-Hertz currents through them to produce heating.
Sometimes an antenna (such as a rotating paraboloidal reflector or lens) is totally
enclosed in a protective housing of low-loss insulating material that is practically trans-
parent to the electromagnetic radiation. Such a housing is called a radome. Radomes are
commonly used on some types of aircraft antennas for aerodynamic reasons. They are
also sometimes used on large microwave ground based antennas, permitting the antenna
itself to be a lighter structure, at the cost of the expensive radome and some loss of
rf power in the radome material. Electrical properties of radomes are addressed in
Appendix E.
Protection against lightning-induced currents and static-charge buildup is necessary
for some types of antennas, such as AM, FM and television broadcasting towers, or any
structure that stands high above its surroundings if the conducting path to ground is not
heavy and direct. Insulators may be protected by horn or ball gaps, and static may be
drained by connecting high-ohmage resistors across insulators.
y
tan φ = (3–3)
FIGURE 3–1. x
Three-dimensional cartesian coordinate system.
Coordinates of an arbitrary point P are shown x = r sin θ cos φ (3–4)
as x1, y1, z1. Corresponding spherical
coordinates are also shown as r1, q1, f1. Also, y = r sin θ sin φ (3–5)
polar coordinates of projection of P (designated
P′) onto xy-plane are shown as r1, f1.
z = r cosθ (3–6)
ρ = x 2 + y 2 = r sin θ (3–7)
x = ρ cos φ (3–8)
y = ρ sin φ (3–9)
The variables x, y, and z range from minus infinity to plus infinity, whereas r and r
range from zero to infinity (always positive). The angle q covers the range of 0 to 180
degrees (0 to p radians), and f goes from 0 to 360 degrees (0 to 2p radians). The angle
q is called the angle of colatitude, or sometimes the polar angle, and f is the angle of
longitude. The system of angular coordinates used to describe positions on the surface
of the earth (latitude and longitude) is essentially a spherical coordinate system with the
radial coordinate fixed, except that the geographic latitude has the value zero at the
equator where the colatitude is 90 degrees, and is +90 degrees at the north pole and −90
degrees at the south pole, where the corresponding values of colatitude are 0 and 180
degrees respectively.
It is in fact sometimes convenient to use the geographic system of coordinates in dis-
cussing antenna patterns that consist of a single main lobe of radiation. The axis of the
lobe is assumed to lie in the xy-plane, so that its longitude angle is f = 0°. Then the angle
q is measured in the xz-plane and corresponds to the latitude angle of geographic coor-
dinates. However, in this context q is usually referred to as the elevation angle, and f is
Radiation Pattern 93
called the azimuth angle. The value of q corresponding to the beam axis is called the
elevation angle of the beam. A positive elevation angle is equal to the complement of
the colatitude angle in the corresponding system of true spherical coordinates.
density in watts per square meter, it is called an absolute pattern. An absolute pattern
actually describes not only the characteristics of an antenna but also those of the associ-
ated transmitter, since the absolute field strength at a given point in space depends on
the total amount of power radiated as well as on the directional properties of the antenna.
An absolute pattern in terms of constant-field-strength contours plotted on a map is a
most useful representation for a radio broadcast station, since it defines the geographic
regions within which various received-signal levels are available to listeners.
Often, however, the pattern is plotted in relative terms, that is, the field strength or
power density is represented in terms of its ratio to some reference value. The reference
usually chosen is the field level in the maximum-field-strength direction. The field
strength or power density is given the value unity in this direction and fractional values
in other directions. This type of pattern provides as much information about the antenna
as does an absolute pattern, and therefore relative patterns are usually plotted when it is
desired to describe only the properties of the antenna, without reference to an associated
transmitter (or receiver).
It is also fairly common to express the relative field strength E or power density p in
decibels. This coordinate of the pattern is given as 20 log(E/Emax) or 10 log(p/pmax). The
value at the maximum of the pattern is therefore zero decibels, and at other angles the
decibel values are negative (since the logarithm of a fractional number is negative).
D2
rmin = 2 (3–10)
λ
where D is the diameter of the smallest stationary sphere enclosing the radiating portions
of the antenna as it is rotated for measurement and l is the wavelength. The factor of 2
in (3–10) is sufficient for sidelobe levels higher than −20 dB, but a factor of 12 may be
required for the measurement of −50 dB sidelobe levels (see sec. 7.10).
Sometimes, however, additional criteria are necessary for assurance that a measured
pattern replicates a far-field measurement. While reviewing the manuscript, E. B. Joy
added two additional criteria: (1) rmin = 20l, that pertains to electrically small antennas
and assures that the reactive near field is negligible at rmin and (2) rmin = 20D, that limits
the relative down range far field decrease across the antenna under test to an acceptable
96 Antenna Parameters
value. As with the factor 2 in (3–10), choice of the factors 20 depends on the accuracy
of measurement required. According to Hollis, Lyon, and Clayton (1970, sec. 14.2), the
relevant effects are usually considered negligible if rmin ≥ 10l and rmin ≥ 10D.
Fields and patterns of an antenna are generally dependent on the distance from the
antenna, broadly separated into near-field and far-field regions. The near-field region is
further subdivided between the reactive near-field and radiating near-field regions.
However, the total field present is actually the vector sum of the reactive (i.e., induction
by time-varying electric and magnetic fields) and radiating fields. At distances from an
antenna that are small compared to a wavelength (or small compared with the antenna
dimensions if the antenna is large), field strengths within reactive near-field regions may
greatly exceed the radiated fields. This is because the nonradiating electric and magnetic
fields, caused by charges and currents on and within the antenna, decrease with range r
as 1/r2 and 1/r3. As a consequence, these reactive fields decrease more rapidly with
increased range than the radiation fields, which vary as 1/r. At some range the reactive
and radiating power densities are equal. This range may serve as a demarcation between
the reactive and radiating near-fields. For a dipole that is electrically very short, or an
equivalent radiator, the outer reactive field boundary is commonly taken to be at a dis-
tance l/2p from the antenna surface (IEEE 1992, p. 1080). Thus, for the VLF band and
lower frequencies, having wavelengths of 104 meters and longer, antennas are electrically
very short. Consequently, the reactive field extends to distances of 1.6 kilometer (about
one mile) and farther. Thus, some of the original “wireless” communications were within
the reactive near zone. Today, short-range communications between equipments is a
developing application of reactive fields. Radio frequency identification (RFID) used for
product inventory is another relatively new application of reactive fields. From near field
measurements with antennas large compared to l, it is known that the strengths of reac-
tive fields are negligible compared to the radiating near fields at distances of 4 or 5l
(sec. 9.1.2). Like all varying fields, reactive fields can, at intense levels, be hazardous to
people. For example, applications of reactive fields include RF heating of metals through
induction and of insulating materials through dielectric loss. Unlike radiating fields, the
magnitudes of electric and magnetic reactive fields are not generally predictable, one
from the other. Thus, an unfamiliar reactive field environment may require monitoring,
separately, the electric and magnetic fields.
The radiating near-field region is that portion of the near-field region between the
reactive near-field and the far-field regions. As already noted, within the far-field the
radiation pattern is essentially independent of the distance from the antenna. However,
within radiating near-field regions, radiation patterns are dependent upon distance from
the antenna. This is because the path length differences between various parts of the
antenna and an observation point, in terms of phase differences, depend on distance from
the antenna center. Consequently, for an antenna small compared to a wavelength, the
radiating near-field region may not exist. Following optical terminology, the radiating
near-field and far-field regions are sometimes called the Fresnel and Fraunhofer regions,
respectively.
Pattern, gain, directivity, and polarization of an antenna are basically far field terms.
Their definitions rely on the simplicity of the far field and do not include the complexities
Radiation Pattern 97
E-field (dB)
–10
antenna patterns at different distances,
–15
only if there is assurance that the strength
of the reactive near field is negligible. –20
The general methods for calculating –25
radiation patterns versus distance are –30
described in Appendix G, and radiating –6 0 6
Observation angle (degrees)
near-field Mathcad calculations for array R=64m
R=320m
and reflector antennas are included in the R=1600m
accompanying website. Figure 3–3 shows FIGURE 3–3.
results of calculating radiation within the Calculated patterns at different distances from
near- and far-field regions for a linear an antenna, for which 2D2/l is 320 m.
antenna, with constant amplitude and
phase over its aperture. The antenna width
D and wavelength l are 4 m and 0.1 m, respectively, giving the calculated 2D2/l
distance of 320 m to the boundary between the radiating near-field and far-field
regions.
Figure 3–3 includes three patterns, corresponding to distances of 10D2/l (1,600 m),
2D2/l (320 m), and 0.4D2/l (60 m). Notice that the 2D2/l and 10D2/l patterns are nearly
identical, except that the first nulls are much deeper for the larger distance. For the
0.4D2/l pattern, the first nulls are not apparent, the maximum field strength is reduced,
and sidelobe levels are increased. Thus, it is seen that the 2D2/l rule-of-thumb provides
useful approximate patterns if very accurate low-sidelobe measurements are not
required.
measurement was made, or the conditions assumed in calculating it, if any possible ques-
tion could exist.
An antenna mounted so that earth-reflection interference effects occur will have its
vertical-plane pattern drastically affected, as Fig. 1–15 indicates. Horizontal-plane pat-
terns may be affected as to absolute values, but relative values will not be affected if the
earth in the vicinity is “smooth.” That is, the shape of the horizontal pattern will not be
affected, unless the earth is irregular.
Pt
pisotrope = watts per square meter (3–11)
4π R 2
Equation (3–11) is true because Pt is distributed uniformly over the surface area of the
sphere, which is 4pR2 square meters.
Actually an isotropic radiator is not physically realizable; all actual antennas have some
degree of nonuniformity in their radiation patterns. A nonisotropic antenna will radiate
more power in some directions than in others and therefore has a directional pattern. A
directional antenna will radiate more power in its direction of maximum radiation than an
isotrope would, with both radiating the same total power. Thus, since the directional
antenna sends less power in some directions than an isotrope does, it follows that it must
send more power in other directions, if the total powers radiated are the same.
Directivity D is a quantitative measure of an antenna’s ability to concentrate radiated
power per unit solid angle in a certain direction, and thus D is highly dependent on the
three-dimensional pattern of an antenna. D will be explicitly defined in materials that
follow. On the other hand, gain G is the ratio of the power radiated per unit solid angle
to the power per unit solid angle radiated by a lossless isotrope, each having the same
input power. Unless otherwise specified, the direction applicable to D and G is that for
which maximum radiation occurs.
Over the years there have been several gain-related terms used. In fact, gain once was
not defined in terms of signal strength relative to any specific antenna type (Terman
1943). Gain is sometimes specified relative to the gain of a one-half wavelength dipole,
whose gain exceeds the isotrope by the factor 1.64 (2.15 dB). Then, gain is specified in
Directivity and Gain 99
dBd (gain above the gain of a lossless one-half wavelength dipole). However, almost
universally, gain is expressed relative to the lossless isotrope and when specificity is
desired, it is referred to as absolute gain or expressed in terms of dBi.
A 2π π 2π
Ω= = ∫ ∫
r2 0 0
sin θ dθ d φ = ∫ 2dφ = 4π (3–12)
0
100 Antenna Parameters
2π π
P= ∫ p (θ , φ, r ) dA = ∫ ∫ p (θ , φ, r ) r 2 sin θ dθ dφ (3–13)
surface 0 0
The power density varies as 1/r2, and thus the product p(q, f, r)r2 is independent of dis-
tance r from antenna. This range-independent product is defined as the radiation intensity
U. Therefore
U (θ , φ ) = p (θ , φ, r ) r 2 (3–14)
dP = p (θ , φ, r ) dA = p (θ , φ, r ) r 2 sin θ dθ dφ = U (θ , φ ) d W
where dW = sin q dq df. Thus, the radiation intensity U(q, f) is dP/dW, that is, the incre-
mental power per unit solid angle in direction (q, f). U(q, f) is usually expressed in units
of watts per steradian (i.e., watts per square radian). Then, from (3–13) and (3–14), total
radiated power P is
2π π 2π π
P= ∫ ∫ p (θ , φ, r ) r 2 sin θ dθ dφ = ∫ ∫ U (θ , φ ) dΩ (3–15)
0 0 0 0
Directivity and Gain 101
In other words, total radiated power equals the sum of all radiation intensity that encloses
the antenna.
3.3.4. Directivity
Directivity D is a quantitative measure of an antenna’s ability to concentrate energy
in a certain direction. Specifically, D is the ratio of the maximum radiation intensity
Umax to the average radiation intensity Uav. Then
If the radiation is isotropic, the radiation intensity in every direction is Uav. Thus, from
(3–12) and (3–15), total radiated power P is a sphere’s solid angle 4p times Uav and is
given by (3–17)
P = 4π U av (3–17)
Now, by shifting the constant Umax to the denominator, it is seen from (3–19) that D is a
function the relative value on U(q, f) (the bracketed term) that has a maximum of unity:
4π
D= 2π π
(3–19)
U (θ , φ )
∫ ∫ U max sin θ dθ dφ
0 0
Therefore, with r fixed, D can be determined with relative values of either U(q, f),
p(q, f, r), or |E(q, f, r)|2.
It is also to be noted that directivity D of (3–19) can be expressed as
4π
D= (3–21)
ΩA
102 Antenna Parameters
where WA is known as the beam solid angle. For the special case of an isotrope, since an
isotrope radiates equally in all directions, WA = 4p and D = 1. In general, however,
2π π
U (θ , φ )
ΩA = ∫ ∫ U max
sin θ dθ dφ (3–22)
0 0
and with r constant, U(q, f)/Umax can be replaced with either p(q, f, r)/p(q, f, r)max or
|E(q, f, r)/E(q, f, r)max|2. Use of the term WA in (3–21) helps to emphasize that directivity
is a measure of an antenna’s ability to concentrate energy in a certain direction. Finally,
it is important to again underscore that directivity is calculated by integrating relative
values of an antenna’s radiation pattern, and this does not require knowledge of an abso-
lute value.
3.3.5. Gain
Gain, previously defined in sec. 3.3.1, is the ratio of the radiation intensity in a given
direction to the radiation intensity that would be obtained if the power were radiated
isotropically. If the direction is not specified, the direction of maximum radiation intensity
Umax is implied. In addition, for isotropically radiated power, the radiation intensity is
Uav. Therefore, for a lossless antenna, gain G equals directivity as defined previously by
(3–16).
Gain is determined, in principle, by comparing the radiation intensity if both the actual
test antenna and an isotrope have the same input power. The isotrope is assumed to radiate
all of its input power, but some of the power delivered to the actual antenna may be dis-
sipated in ohmic resistance (i.e., converted to heat). Thus, gain takes into account the
antenna efficiency as well as its directional properties. The efficiency factor k is the ratio
of the power radiated by the antenna to the total input power. Thus, the relationship
between gain G and directivity D is
G = kD (3–23)
where the word log denotes the common logarithm (base 10). Directivity in decibels is
calculated using the same formula, with D substituted for G. Gain is expressed in decibels
more commonly than as a power ratio.
Pr = pr Ae (3–24)
The area Ae, which bears this relationship to Pr and pr, is the effective area of the antenna.
(The conception of the capture process thus conveyed is a much oversimplified one, but
it has validity for the present purposes.) Since received power depends on the antenna
polarization, effective area is a function of the polarization of the incident field.
As might be supposed, there is a connection between the effective area of an antenna
and its physical area as viewed from the direction of the incoming signal. The two areas
are not equal, however, although for certain types of high-gain antennas they may be
nearly equal. But some antennas of physically small cross section may have considerably
larger effective areas. It is as though such an antenna has the ability to “reach out” and
capture power from an area larger than its physical size, as in the case of a dipole
antenna.
As is discussed in sec. 5.7.6, there is also a relationship between the gain of a lossless
antenna and its physical size. This relationship suggests that there may also be a connec-
tion between the gain and the effective area, and this indeed turns out to be true. The
equation relating the two quantities is
104 Antenna Parameters
Gλ 2
Ae = (3–25)
4π
where l is the wavelength corresponding to the frequency of the signal. (This relation-
ship may be proved theoretically and verified experimentally.)
Because of this connection between the effective area and the gain, (3–24) may be
rewritten, with the right-hand side of (3–25) substituted for Ae. Thus:
pr Gλ 2
Pr = (3–26)
4π
Therefore the concept of the effective area of an antenna is not a necessary one. As (3–26)
shows, it is possible to calculate the received-signal power without knowing Ae. As seen,
the effective area definition has a conceptual value, however, and is a convenient quantity
to employ in some types of problems. We now consider the relationship between transmit
and receive powers with two antennas separated by far-field distance R. This relationship
is highly useful, and it is known as the Friis Transmission Equation (Friis 1946). The
following abbreviations are used:
Pr = received power
Pt = input power of transmit antenna
Aet = effective aperture of transmit antenna
Aer = effective aperture of receive antenna
Gt = gain of transmit antenna
Gr = gain of receive antenna
Now, from equation (3–11) for a lossless isotropic radiator and the definition of gain, the
power density pr at distance R from a transmit antenna may be expressed as (Pt/4pR2)Gt.
Recall that another antenna in the presence of pr will receive power Pr that equals pr Aer.
Thus Pr becomes (Pt/4pR2)AerGt. Finally, we use relationships, based on (3–25), between
Gr and Aer and between Gt and Aet to express Pr as functions of Aet Aer and of GrGt. The
result is the Friis Transmission Equation which is given in (3–27) that follows.
G G λ2
Pr = t 2 Aer Gt = Pt et2 er2 = Pt t 2r 2
P A A
(3–27)
4π R R λ (4π ) R
When reciprocity applies, as is usually so, the effective transmit aperture area Aet of
an antenna equals its effective receive aperture area Aer. Similarly, transmit gain Gt and
receive gain Gr of an antenna are equal. Then, because of reciprocity, the transmitter and
receiver locations can be interchanged. Thus, either antenna can be designated as the
transmit antenna and the other as the receive antenna.
Equation (3–27) underscores the practical significance of gain. For example, a transmit
power of 1,000 watts and a transmit antenna gain of 10 (10 dB) will provide the same
received power as will a transmit power of 500 watts and a transmit antenna gain of 20
Beamwidth 105
(13 dB). Obviously, this relationship has great economic significance. Sometimes it may
be much less expensive to double antenna gain (add 3 dB) than it would be to double
transmit power (though in other cases the converse may be true). Generally, it is desirable
to use as much antenna gain as feasible, because it increases received signals.
3.5. Beamwidth
When the radiated power of an antenna is concentrated into a single major “lobe,” as
exemplified by the pattern of Fig. 3–2, the angular width of this lobe is the beamwidth.
The term is applicable only to antennas whose patterns are of this general type. Some
antennas have a pattern consisting of many lobes, for example, all of them more or less
comparable in their maximum power density, or gain, and not necessarily all of the same
angular width. One would not speak of the beamwidth of such an antenna, but a large
class of antennas do have patterns to which the beamwidth parameter may be appropri-
ately applied.
On the other hand, there are situations that call for a wide beam. For example, a
broadcasting station must radiate a signal simultaneously to listeners in many different
directions—typically, over a 360 degree azimuth sector. Any narrowing of the beam to
obtain gain must therefore be done in the vertical plane. At the low frequencies of the
AM broadcast band (535–1,705 kHz), and at lower frequencies, such vertical-plane beam
narrowing is not feasible because it would require an impractically large (high) antenna.
In television and FM broadcasting in the VHF and UHF bands, however, this means of
obtaining gain while preserving 360 degree azimuthal coverage is much used. A situation
in which an antenna beam must be moderately broad in the vertical plane is that of a
search radar antenna on a ship that rolls and pitches. It is usually required that the beam
of such antenna be directed at the horizon. But if the beam has, say, a 2-degree vertical
beamwidth, and the ship rolls 20 degrees (which is not unusual), it is evident that at times
no part of the beam will remain directed at the horizon. The vertical beamwidth in this
case must be of size comparable to the maximum roll angle of the ship, unless some
method of stabilizing the beam is employed (as is sometimes done).
4 Beam pattern
5 points correspond to the minus 3 dB
6 points. For this reason the half-power
10° beamwidth is often referred to as the 3 dB
beamwidth. Figure 3–5 illustrates the pro-
–10° –5° 0° +5° +10° cedure of determining the 3 dB beam-
Angle width on a rectangular pattern plot. Only
FIGURE 3–5. the nose region of the beam pattern is
Determination of half-power (3 dB down) shown. As indicated, this beamwidth is
beamwidth. approximately 10 degrees.
Minor Lobes 107
P
Rr = (3–28)
I2
The concept of radiation resistance is applicable only to antennas in which the radia-
tion is associated with a definite current in a single linear conductor. Even then, the defi-
nition of radiation resistance is ambiguous. This is because of standing waves, the current
is not the same everywhere along a linear conductor. It is therefore necessary to specify
the point along the conductor at which the current will be measured. Two points some-
times specified are where the current has its maximum value and the feed point (input
terminals). These two points are sometimes at the same place, as in a center-fed dipole,
Radiation Resistance and Efficiency 109
but they are not always the same. And in principle, any point may be specified. The value
then obtained for the radiation resistance of the antenna depends on what point is speci-
fied; this value is the radiation resistance referred to that point.
The above words “maximum current” refer to the rms current in that part of the
antenna where the current is maximum. It does not mean the peak value of the current
located where the current is greatest. The reader will recall that, in accordance with equa-
tion (1–3), by definition the amplitude I0 of a current sine wave is actually the peak value
of current during its cycle. Furthermore, the reader is cautioned that, in some texts, for-
mulas for radiation resistance are written as a function of current “amplitude” I0, instead
of rms current. Thus, use of the wrong combination of formula and current level will
cause the radiation resistance to be incorrectly calculated.
The radiation resistance of some types of antennas can be calculated, and yet not for
others. Sometimes radiation resistance can be obtained by measurement (see section 9.10,
chapter 9). Typical values of the input radiation resistance of actual antennas range from
a fraction of an ohm to several hundred ohms. The very low values are undesirable,
because they imply large antenna currents (i.e., P = I 2R). Therefore, there exists the
possibility of considerable ohmic loss of power, that is, dissipation of power as heat
rather than as radiation. An excessively high value of radiation resistance is also undesir-
able, because this requires that a very high voltage (i.e., P = E 2/R) be applied to the
antenna.
Antennas always have ohmic resistance, although sometimes it may be so small as
to be negligible. The ohmic resistance is usually distributed over the antenna; and since
the antenna current varies, the resulting loss may be quite complicated to calculate. In
general, however, the actual loss can be considered to be equivalent to the loss in a ficti-
tious lumped resistance placed in series with the radiation resistance. If this equivalent
ohmic loss resistance is denoted by R0, the full power (dissipated plus radiated) is
I 2(R0 + Rr), whereas the radiated power is I 2Rr. Hence the antenna radiation efficiency
k of (3–23) is given by
Rr
kr = (3–29)
R0 + Rr
It must be acknowledged, however, that this definition of the efficiency is not really very
useful even though it may occasionally be convenient. The fact is that both R0 and Rr
are fictitious quantities, derived from measurements of current and power; Rr is given in
these terms by (3–28), and R0 is correspondingly equal to P0/I 2. Making these substitu-
tions into (3–29) gives the more basic definition of the efficiency:
Pr
kr = (3–30)
P0 + Pr
Ei
Zi = (3–31)
Ii
Z i = Ri = Rr + R0 (3–32)
The input impedance will be nonreactive for this feed point if the antenna is resonant (as
in the case of a wire approximately an integral number of half wavelengths long, dis-
cussed in secs. 4.2–4.4) or if the antenna is terminated by a resistance of proper value
at its far end so that there is no standing wave of current on it (as described in sec. 4.4).
In other cases, however, the input impedance may consist effectively of the radiation and
loss resistances in series with a very high reactance.
If this reactance has a large value, the antenna input voltage must be very large to
produce an appreciable input current. If in addition the radiation resistance is very small,
the input current must be very large to produce appreciable radiated power. Obviously
this combination of circumstances, which occurs with the “short dipole” antennas that
must be used at very low frequencies, results in a very difficult “feed” impedance-match-
ing problem. Methods of solving it are described in sec. 4.1 and 4.11.
The concepts of radiation resistance and feed point or input impedance have been
discussed here in terms of antennas that have currents flowing in linear conductors as
the basis of the radiation. As will be discussed in secs. 4.7, 4.8, and 4.10, other types of
antennas exist with radiating properties difficult to analyze in these terms. They are, for
example, fed by waveguides rather than by transmission lines. The equivalent of an input
Bandwidth 111
impedance can be defined at the point of connection of the waveguide to the antenna,
just as waveguides have a characteristic wave impedance analogous to the characteristic
impedance of a transmission line. It is difficult to define a “radiation resistance” for
such antennas.
Even for some types of antennas consisting of current-carrying conductors this is dif-
ficult, and it may even be difficult to define an input impedance. This is true, for example,
of an array of dipoles, when each dipole is fed separately; sometimes each dipole, or
groups of dipoles, will be connected to separate transmitting amplifiers and receiving
amplifiers. The input impedance of each dipole or group may then be defined, but the
concept becomes meaningless for the antenna as a whole, as does also the radiation
resistance. Both these terms have meaning primarily for simple linear-current radiating
elements; but they comprise a very large class of antennas.
3.9. Bandwidth
All antennas are limited in the range of frequency over which they will operate satisfac-
torily. This frequency range, whatever it may be, is called the bandwidth of the antenna.
If an antenna were capable of operating satisfactorily from a minimum frequency of
195 MHz to a maximum frequency of 205 MHz, its bandwidth would be 10 MHz. It
would also be said to have a 5 percent bandwidth (the actual bandwidth divided by the
center frequency of the band, times 100). This would be considered a “moderate” band-
width. Some antennas are required to operate only at a fixed frequency with a signal that
is narrow in its bandwidth; consequently there is no bandwidth problem in designing
such an antenna. But in other applications much greater bandwidths may be required;
in such cases special techniques are needed. Such techniques are known and will be
described in secs. 7–1 and 7–2. In fact, some recent developments in broad-band anten-
nas permit bandwidths so great that they are described by giving the numerical ratio of
the highest to the lowest operating frequency, rather than as a percentage of the center
frequency. In these terms, bandwidths of 20 to 1 are readily achieved with these antennas,
and ratios as great as 100 to 1 are possible.
The bandwidth of an antenna is not as precise a concept as some of the other parame-
ters that have been described, because many factors are involved in what is meant by
“operating satisfactorily.” The principal ones are input impedance, radiation efficiency,
gain, beamwidth, beam direction, polarization, and sidelobe level.
These may not all be involved in every case, because one or another of them may be
so much more critical than the others, in a particular case, that it alone determines the
bandwidth. The two basic factors involved, between which a distinct separation may be
made, are the antenna pattern and the input impedance. Accordingly, the terms pattern
bandwidth and impedance bandwidth are sometimes used to emphasize this distinction.
The beamwidth, gain, sidelobe level, beam direction, and polarization are parameters
associated with the pattern bandwidth, whereas input impedance, radiation resistance,
and efficiency are associated with the impedance bandwidth.
The definition of the bandwidth of an antenna is less precise than the definitions of
other parameters for still another reason: there is no established criterion of “satisfactory”
112 Antenna Parameters
3.10. Polarization
The radiation of an antenna may be linearly, elliptically, or circularly polarized, as defined
in sec. 1.1.4 (also see Appendix B). Polarization in one part of the total pattern may be dif-
ferent from polarization in another. For example, the polarization may be different in the
minor lobes and in the main lobe, or may even vary in different parts of the main lobe.
The simplest antennas radiate (and receive) linearly polarized waves. They are usually
oriented so that the polarization (direction of the electric vector) is either horizontal or
vertical. Sometimes the choice is dictated by necessity, at other times by preference based
on technical advantages, and sometimes one polarization is as good and as easily achieved
as the other.
For example, at the very low frequencies it is practically impossible to radiate a hori-
zontally polarized wave successfully (because it will be virtually canceled by the radia-
tion from the image of the antenna in the earth); also, vertically polarized waves propagate
much more successfully at these frequencies (e.g., below 1,000 kHz). Therefore vertical
polarization is practically required at these frequencies.
At the frequencies of television broadcasting (between 54 and 698 MHz), horizontal
polarization has been adopted as standard. This choice was made to maximize signal-to-
Interdependencies of Gain, Beamwidths, and Aperture Dimensions 113
noise ratios; it was found that the majority of man-made noise signals are predominantly
vertically polarized. (Interestingly, in Great Britain the opposite choice was made, because
from the pure wave-propagation point of view, vertical polarization provides maximum
signal strength.)
At the microwave frequencies (above 1 GHz) there is little basis for a choice of hori-
zontal or vertical polarization, although in specific applications there may be some pos-
sible advantage in one or the other. Of course in communication circuits it is essential
that the transmitting and receiving antennas be polarization matched.
Circular polarization has advantages in some VHF, UHF, and microwave applications.
For example, in transmission of VHF and low-UHF signals through the ionosphere, rota-
tion of the polarization vector occurs, the amount of rotation being generally unpredict-
able. Therefore if a linear polarization is transmitted it is advantageous to have a circularly
polarized receiving antenna (which can receive either polarization), or vice versa.
Maximum power transfer is realized when both antennas are either left- or right-circularly
polarized. (This applies, for example, to transmission and reception between the earth
and a space vehicle.) Circular polarization has also been found to be of advantage in
some microwave radar applications to minimize the “clutter” echoes received from rain-
drops, in relation to the echoes from larger targets such as aircraft.
It is apparent that the polarization properties of an antenna are an important part of
its technical description—a parameter of its performance. Sometimes it may be desirable
to provide a “polarization pattern” of the antenna, that is, a description of the polarization
radiated as a function of angles within a spherical coordinate system, although such a
complete picture of the polarization is not ordinarily required.
Since gain G = kD from (3–23), we can express G in terms of an effective area Ae and
l as
where k = Ae/Aem.
114 Antenna Parameters
The definitions of Aem and Ae are defined by (3–33) and (3–34). Physical aperture Ap
is another frequently used aperture term, but it is less clearly defined. Obviously, Ap is a
measure of the physical size of an antenna and it is an area over which radiation is trans-
mitted or received. Physical area for a horn or reflector antenna can be readily visualized.
Presumably Ap for a stub antenna is its cross-sectional area, but what about one attached
to an automobile which affects its radiation pattern? Another definition sometimes used
is aperture efficiency eap, where
ε ap = Ae /Ap (3–35)
Depending on how Ap is defined, eap can exceed unity. For a uniformly continuously
illuminated aperture, that is, one for which the amplitude and phase are both constant
over the physical area, Aem = Ap. This can be shown to be true from results of integrating
the total radiation pattern for a uniformly illuminated aperture and calculating D from
(3–33), as done by Silver (1949, p. 183). Thus, for a lossless, uniformly continuously-
illuminated aperture eap = Ae/Ap = 1 and
The actual gain of an antenna is of course less than D of (3–36), because of losses that
may occur from large and extraneous sidelobes, impedance mismatches, and dissipative
(I 2R) losses.
Theoretically, the half-power beamwidth qHP for a uniformly illuminated line
source is
or (3–37)
For the equations of (3–37), l and d are in the same units. The equations of (3–37) can
be derived from the pattern for a uniformly illuminated aperture, which is discussed in
sec. 6.10.
Now consider a rectangular aperture. Its gain can be determined with (3–36) by
expressing physical aperture Ap in terms of principal plane beamwidths, qHP and fHP,
using (3–37). Accordingly, let dq and df denote the linear dimensions that control the
beamwidths in q and f directions, respectively. Then, theoretically, the gain G for a loss-
less, uniformly continuously-illuminated aperture is
The approximate, “typical” gain of (3–39) is nearly 1.0 dB less than that of (3–38), which
is for a lossless, uniformly illuminated rectangular aperture. Furthermore, the use of
(3–39) is recommended only if the known antenna efficiency (“k” in G = kD) is nearly
1. The point being that directivity and thus gain are reduced by spurious radiation not
readily detected and thus not ordinarily accounted for when determining directivity.
For example, if only the usual principal-plane measurements are made, there may be
unsuspected out-of-plane spurious radiation.
References
Friis, H. T., “A Note on a Simple Transmission Formula,” Proceedings of the IRE, vol.
34, May, 1946, pp. 254–56.
Hollis, J. S., T. J. Lyon, and L. Clayton, Microwave Antenna Measurements,
Scientific-Atlanta, Inc., 1970. This publication may be available through MI
Technologies, Inc. (www. mi-technologies.com).
Jurgenson, R., and R. G. Brown, Geometry, Houghton Mifflin Co., 2000.
IEEE Standard 100-1992, The New IEEE Standard Dictionary of Electrical and Electronic
Terms, 1993.
Silver, S., Microwave Antenna Theory and Design, McGraw-Hill, 1949.
Stutzman, W. L., “Estimating Directivity and Gain of Antennas”, IEEE Antennas and
Propagation Magazine, August, 1998, pp. 7–11.
Terman, F. E., Radio Engineers’ Handbook, McGraw-Hill, 1943, p. 785.
(b) A steel tower can be used to support an antenna, or sometimes it can itself
be the antenna.
(c) The place at which the feed line connects to an antenna is called its input
terminals or input port.
(d) Guy wires of antenna towers should be interrupted by insulators spaced a
half-wavelength apart.
(e) Solid copper is always the preferred form of conductor for an antenna.
(f ) A radome is a dome-shaped radar antenna.
2. A directional antenna has a maximum electrical dimension of 50 meters, and its
operating frequency is 100 MHz (108 Hz). A field strength measurement is made
at a distance of 1 km from this antenna, in the main beam. (a) Is this a near-field
or a far-field measurement? (b) What is the approximate distance at which the
near field ends and the far field begins?
3. The solid angle subtended by the sun as viewed from the earth is W = 6 × 10−5
steradian. A microwave antenna, designed to be used for studying the microwave
radiation from the sun, has a very narrow beam whose “equivalent” solid angle,in
the sense of the denominators of equations (3–19) and (3–21), is approximately
equal to that subtended by the sun. Assume that this antenna has no minor lobes.
What is its approximate directivity D? Express this gain as both a power ratio
and a decibel value.
4. An antenna radiates a total power of 100 watts. In the direction of maximum
radiation, the field strength at a distance of 10 km (104 meters) was measured
and found to be E = 12 millivolts (0.012 volt) per meter. (a) What is the directiv-
ity D of this antenna, assuming free-space propagation to the measuring point?
(b) If its efficiency factor is kR = 0.92, what is its gain G? (c) If the wavelength
is l = 3 meters, what is the effective area Ae in square meters?
5. If the operating wavelength of an antenna, whose gain G = 30,000, is l = 0.1
meter, (a) what is its effective area, Ae, and (b) If the actual cross-sectional area
of the antenna is the same as its effective area, and if the outline of this area is
circular, what is the antenna diameter? (Recall from elementary geometry that
the area of a circle of diameter d is p(d/2)2.)
6. Derive the two forms of the Friis Transmission Equation given in equation
(3–27). In doing so, start with equation (3–11) and the definition of gain. (These
two forms of the Friis equation are in fact of greater practical value than
equations that include power density.)
7. A dipole slightly shorter than a half wavelength, fed at the center, has an input
impedance that is purely resistive and of value 75 ohms. The conductor has some
ohmic resistance, and the resulting ohmic loss of power is equivalent to what
would result if the dipole were a perfect conductor but had a “lumped” resistance
of value R0 = 8 ohms connected in series with it at the feed point. (a) What is
the radiation resistance, Rr, of the dipole? (b) What is the radiation efficiency
factor, kr? (c) The directivity of a dipole approximately a half-wavelength long
is D = 1.64. What is the gain, G, of this particular dipole?
Problems and Exercises 117
8. The two VHF television bands are 54 to 88 MHz (low band) and 174 to 216 MHz
(high band). (a) What are the percentage bandwidths of each of these two bands?
(b) If a single receiving antenna is designed to perform satisfactorily at all fre-
quencies between the bottom end of the low band and the top end of the high
band, what is its bandwidth, expressed in the conventional way?
9. Assume N isotropes are placed along a line, and are energized equally in ampli-
tude and phase by the same power source. By using equations (1–9) and (3–11),
what is the theoretical gain perpendicular to the line of isotropes?
10. List three quantities pertaining to the electric field that should be plotted, as
functions of the direction angles q, f, to present all possible information about
the antenna pattern in the far field. (Note: Do not include the magnetic intensity
or the power density, as these are both derivable from a knowledge of the electric
field.)
CHAPTER 4
The preceding chapters have been largely groundwork, covering principles, concepts,
and terminology, as well as some basic antenna theory. It is now possible to discuss
specific antennas.
Initially the basic forms of radiating structures will be discussed. They are sometimes
used by themselves as simple antennas; they may be very effective in some applications.
They are also used in combination with each other and with other components to form
more complicated antennas having special properties such as high directivity, large band-
width, omnidirectionality, steerable beams, low noise, and so on. These more complicated
antennas are described in chapters 5 through 8. The properties of the basic radiators will
be discussed in terms of their behavior in free space, except for specific types that require
the presence of the earth for their operation, as will be indicated in the discussion. It will
also be assumed, except where specifically otherwise stated, that current-carrying por-
tions of the antennas are perfect conductors. The conductors of actual antennas are
usually very good so that this assumption is reasonably well approximated in practice.
When some conductor losses do occur, the principal effect is to reduce the efficiency, as
defined by equation (3–30), chapter 3. Ordinarily the pattern (and hence the directivity
and beamwidth) are not seriously affected, although appreciable conductor resistance
would affect these parameters also.
119
120 Basic Radiators and Feed Methods
A short dipole that does have a uniform current will be called an elemental dipole.
Such a dipole will usually be considerably shorter than the tenth-wavelength maximum
specified for a short dipole. The term infinitesimal dipole may be used to imply extreme
shortness, as required in certain mathematical analyses. Other terms sometimes used for
an elemental dipole are elementary dipole, elementary doublet, and Hertzian dipole. Part
of the importance of this type of dipole is that many more complicated antennas can be
analyzed by considering them to be assemblages of many elemental dipoles. For example,
a long-wire antenna may be regarded as composed of many elemental dipoles connected
end-to-end. Although the current is considered constant (at any instant of time) along the
length of each elemental dipole, the currents in different dipoles may be different, both
in magnitude and phase; therefore, a nonuniform current in the long wire can be approxi-
mated by this representation.
If a center-fed short dipole is initially in a neutral condition and then a current starts
to flow in one direction, one half of the dipole will acquire an excess of charge and the
other a deficit (since a current is a flow of electrical charge). There will then be a voltage
between the two halves of the dipole. If the current then reverses its direction, this charge
unbalance will be first neutralized and then reversed. Therefore an oscillating current will
result in an oscillating voltage as well (or vice versa). If the current oscillation is sinu-
soidal, the voltage oscillation will also be sinusoidal and approximately 90 degrees
lagging the current in phase angle; that is, the short dipole is capacitive in nature, from
the viewpoint of its current-voltage relationship.
Since such a dipole can be regarded as one in which electric charge oscillates, it is an
oscillating electric dipole. It is distinguished from an oscillating magnetic dipole, which
is equivalent to a bar magnet whose magnetic strength and polarity oscillate. An example
of a magnetic dipole is discussed in sec. 4.5.1.
The current, though uniform throughout the length of the elemental dipole at any
instant, is assumed to vary sinusoidally in time, according to (4–1) that follows
where I(t) is the current at any time t, I0 is the amplitude of the current (the rf peak value),
f is the frequency in Hertz, t is the time in seconds, and a is the phase angle (which
simply means that when t = 0, I(t) = I0 sin a).
In the radiation field there is a continual exchange of energy between the electric and
magnetic field components. As part of his theory, Maxwell postulated that an electric
field acting in empty space causes a “displacement current” just as it causes an electron
current in a conductor. He further hypothesized that this displacement current has the
same ability to set up a magnetic field as does an electron current. Many experiments
confirm this theory, which is now unquestioned. A varying electric field therefore causes
a varying displacement current, which in turn results in a varying magnetic field. The
magnetic field, in its turn, creates a varying electric field in accordance with Faraday’s
law. The cycle is thus complete; each field component sustains the other, and in regions
remote from conductors or charges (empty space), one cannot exist without the other, in
the fixed ratio given by equation (1–10), chapter 1.
That electric and magnetic fields represent energy may be demonstrated by placing a
charged object in an electromagnetic field. The field will exert a force upon such an object
and cause it to move if there are no restraining forces. This effect occurs, for example,
when radio waves encounter the earth’s ionosphere, in which there are free electrons.
These electrons are accelerated by the field, which means that kinetic energy is
imparted to them. The energy is supplied by the electromagnetic field. Similarly, an
electromagnetic field impinging on a receiving antenna transfers energy (power) to it,
which may be amplified by the receiver and eventually converted into sound variations
in a loudspeaker, light variations on a television screen, or some other form of
intelligence.
l (x) f = 0°
where l is the length of the elemental
dipole and I is the dipole current, in
amperes. If l, l, and r are in meters, and
I is in rms amperes, E is obtained from
this formula in rms volts per meter. FIGURE 4–3.
The fact that the angle f does not Relationship of elemental dipole radiator to
appear in this expression means that for spherical coordinate system.
124 Basic Radiators and Feed Methods
π 2π r π ⎛ 4r ⎞
γ = − = ⎜1 − ⎟ radians (4–3)
2 λ 2⎝ λ⎠
Note that −2pr/l is the delay of phase, in radians, created by the separation distance r.
Equation (4–3), as well as (4–2), applies only in the far field, which means at values of
r that are of the order of l or greater.
Short Dipoles 125
The relative-power-density pattern may be obtained from (4–2) by making use of the
relation p = E2/377 from (1–9). (It is useful here to note that the number 377 actually
represents the quantity 120p.) The result is
30π I 2 l 2 sin 2 θ
p (r , θ , φ ) = (4–4)
λ 2r 2
This formula gives p in watts per square meter when I is in rms amperes and all lengths
are in meters.
2π π 30π I 2 l 2 2π π sin 2 θ 2
Ptotal = ∫
0 ∫0 [ p (r, θ , φ )] r 2 sin θ dθ dφ = λ2
∫0 ∫0 2
r sin θ dθ dφ
r (4–5)
60π 2 I 2 l 2 π 790 I 2 l 2
=
λ2
∫0 sin θ dθ =
3
λ2
watts
The radiation resistance may now be calculated, from (3–28), by dividing Ptotal by I2.
The result is
l2
Rr = 790 ohms (4–6)
λ2
For example, if l/l = 0.1, Rr = 7.9 ohms. (It will be recalled that this value of l/l was
stipulated at the beginning of this section to be approximately the maximum permissible
length for an antenna to qualify as an elemental dipole.) The dipole current required for
one kilowatt of power with this value of radiation resistance is 11.3 amp (a rather large
current). This result indicates a major disadvantage of short dipoles, namely, the very
large currents required for radiation of appreciable power. Therefore, radiators with larger
values of radiation resistance are preferred when they can be used; but this is not always
possible, and short dipoles are very useful radiators under these circumstances.
4.1.4. Directivity
The directivity of an elemental dipole can be computed by using foregoing results and
the fundamental definition of directivity, equation (3–16), that follows:
providing p(r, q, f)max and p(r, q, f)av are evaluated at the same distance r. The numerator
of (3–16), the equation above, is available from (4–4) with q = 90°, so that sin2 q = 1
(the maximum value). The denominator is Ptotal/4pr2, where 4pr2 is the area of a sphere
of radius r and Ptotal is given by (4–5). Then, complete expression for the directivity (with
790 written as 80p2) is
30π I 2 l 2 4π r 2 λ 2
D= = 1.5 (4–7)
λ 2 r 2 80π 2 I 2 l 2
Thus the directive is 1.5 regardless of the exact length l in relation to the wavelength l,
as long as the radiator qualifies as an elemental dipole. In fact, this result applies to any
short dipole.
4.1.5. Beamwidth
The pattern of Fig. 4–4 (the two circles formed by heavy lines) is created by a “slice”
through the three-dimensional pattern. However, the total pattern in space is shaped like a
doughnut (with a pin-sized hole). The two lobes of Fig. 4–4 are really cross sections of the
same lobe, and the angular width of this doughnut-shaped lobe will now be discussed.
The half-power beamwidth may be determined either by measuring the angular width
of the pattern, as plotted in Fig. 4–4, between the points of electric intensity equal to
0.707Emax, or from analysis of (4–4). The only angle-dependent term in this expression
is sin2q, and it determines the pattern and the beamwidth. The quantity sin2q has its
maximum value of unity at q = 90°. Therefore, the half-power beamwidth is determined
by finding the values of q where sin2q equals 12 . The values are q = 45° and q = 135°
and thus the half-power beamwidth is 135° − 45° = 90°.
Transmission Transmission
line C R0 line
l
Rr
Input terminals
Dipole input
terminals
(a) Dipole
(b)
FIGURE 4–5.
Equivalent and actual circuits of center-fed short dipole. (a) Equivalent circuit. (b) Center-fed
dipole and feed line.
This high current, flowing through the inductive and capacitive reactances, produces
very high voltages across the feed-circuit inductor and across the antenna input terminals,
even though the transmitter itself does not have to supply a high rf voltage (because the
inductive and capacitive reactances cancel, leaving only Rr and R0 as the effective trans-
mitter load). If the input resistance Ri (= R0 + Rr) is very small, a more complicated
arrangement than that shown in Fig. 4–6 may be required to provide an impedance
transformation in addition to reactance cancellation. Therefore feeding a short dipole
antenna is somewhat difficult. In particular, the inductance in the feed circuit, when high
power is being radiated, must be very large both to withstand the high voltage and to
carry the heavy current without excessive loss. These are expensive requirements.
Transmission Transmission
line L C R0 line
l
Rr
Tuning coils
Dipole input dipole
terminals
(a)
(b)
FIGURE 4–6.
Method of tuning out dipole capacitive input reactance with series inductance (L).
(a) Equivalent circuit. (b) Dipole with input tuning coils. Reactance cancellation results
when 2p fL = 1/2pfC.
cannot exist where it has “no place to go.”) At the same time, current can exist elsewhere
in the dipole. This statement would seem to violate the physical principle of continuity
for electric current, which implies that a current must have the same value everywhere
along a continuous conductor. It will be recalled that this is not true in a transmission
line on which a standing wave exists (sec. 2.1). The current in the dipole is a standing
wave. It may be thought of as two equal and opposite currents (the reverse current is due
to reflection from the open end) that, at the ends of the dipole, have exactly opposite
phases and thus cancel, resulting in zero current. The cancellation is incomplete (because
of the changing phase relationship) at points a short distance from the end and becomes
progressively less complete going from the ends to the center. The principle of continuity
is satisfied for the currents traveling in each direction, separately.
When the dipole is very short compared with a wavelength, as assumed, the current
will vary approximately linearly along the dipole from end to center. This means that a
graph representing the current as ordinate I, and distance from the end to the center of
the dipole as abscissa x, is a pair of sloping straight lines forming a triangle, as shown
in Fig. 4–7, with zero value at the left end (end of the dipole, x = −l/2), maximum value,
Imax, at the center, x = 0, and zero value again at the right end (x = +l/2, l being the total
length of the dipole). It can be shown that the short dipole with linear current distribution
is like an elemental dipole with an “effective length” equal to half its actual length, and
a uniform current equal to the current at the center of the actual dipole. Equations (4–2)
to (4–6) may be used to describe the behavior of the actual short dipole if l in
these equations is replaced by (l/2). With this modification, the entire discussion of the
Short Dipoles 129
(2) substitute p/2 for p in the q-integration of (4–5), because the earth’s block-
age; this causes the factor 790 to be replaced by 395.
Then, by using the Ptotal for the monopole, its radiation resistance becomes
Ptotal 395h 2
Rr = = (4–8)
I2 λ2
It is to be recalled that h is the actual height of the half-dipole of the monopole, which
is the same as the effective height of the half-dipole plus its image. Alternatively, Rr may
be expressed in terms of the effective height (i.e., length) of the half-dipole he, which
equals h/2. Then, Rr = 1580h2e/l2.
Recall that, for a monopole, the total power radiated is effectively concentrated into
half the solid angle of the free-space case. Then, by the reasoning given in chapter 3,
e.g., equation (3–21), the directivity will be twice that of a free-space short dipole. Thus
for the monopole, D = 3, as compared to D = 1.5 given in (4–7) for the elemental dipole.
The pattern in the horizontal plane is uniform (circular), that is, equal signal strength is
radiated in all horizontal directions. The radiation at the peak of the beam is vertically
polarized. The half-power vertical beamwidth is half that of the free-space short dipole
(i.e., 45 degrees) since the earth eliminates half of the pattern. However, calculations of
field strength at a distance in the actual earth environment cannot be made on the basis
of these “semi-free-space” results because the propagation of the vertically polarized
waves depends on the semi-guiding effect of the earth’s surface, and at very great dis-
tances the ionosphere plays a part, as discussed in sec. 1.4.
A vertical radiator of this type may take the form of a steel tower with its base insu-
lated from earth; then it is fed by connecting the source (transmitter) between the tower
base and the ground, with an inductance in the feed line to compensate for the capacitive
reactance of the antenna. The ground is a source of appreciable loss unless care is taken
to minimize the resistance of the ground connection. In high-power transmitting instal-
lations an elaborate network of buried wires is used to make a good connection.
Flat top
Insulator Insulator
Insulator
Feed wire
Ground
FIGURE 4–8.
A top-loaded short-monopole antenna, semischematic.
* The equivalent procedure for a dipole would be called end loading. The term top loading is used simply
because the end of the monopole is at the top. On a dipole, both ends would be “loaded,” symmetrically.
132 Basic Radiators and Feed Methods
apart, and the vertical radiating portion is close to 300 meters in height. Another United
States Navy installation, in Cutler, Maine, has towers nearly 300 meters high; but even
these heights are short compared to the wavelength at these frequencies (l = 10,000
meters at f = 30 kHz).
To illustrate the loss problem with electrically short vertical antennas using ground
images, typical radiation efficiency factors (kr, equation 3–30) range from 0.05 to 0.5
(often expressed as 5 to 50 percent). The majority of the loss occurs in the ground resis-
tance, though some occurs in the feed-line tuning coil and, in very high-power installa-
tions, in insulator leakage losses and corona. (Efficiencies of high-frequency antennas
are often close to 100 percent.)
2π x
I 0 ( x ) = jI 0(m ) sin
λ
(4–9)
2π x
V0 ( x ) = V0(m ) cos
λ
(4–10)
where I0(m) and V0(m) are the maximum amplitudes. These equations do not show the time
variation, or instantaneous values; they are obtained by multiplying the amplitudes by
the factor sin (2p ft + a), as in (4–1).
The space–time relationships of current and voltage on a half-wave dipole are some-
what difficult to visualize at first. As an aid in this effort, Fig. 4–11 shows the instanta-
neous patterns of current and voltage on the dipole at several instants during a single
rf cycle. In these diagrams, the rf period T is the time for completion of a single cycle;
* In these expressions, the j factor indicates that the current is 90 degrees out of phase with the voltage,
and the positive and negative sign changes resulting from the sine and cosine functions of x indicate the
180-degree phase changes that occur at the standing-wave nulls. I0(m) is the value of I0(x) that occurs at
x = l /4, and V0(m) is the value of V0(x) that occurs at x = 0 and x = l /2.
134 Basic Radiators and Feed Methods
Half-wave dipole
l /4
l /2 V0
Two-wire line
l /4
I0
(a) (b) – 90° 0° + 90°
(c)
FIGURE 4–10.
Current-voltage distribution on half-wave dipole. Comparison with Figure 4–9 indicates
correspondence with distribution on open-ended transmission line. (a) End-quarter-wave section
of two-wire line bent back to form half-wave dipole. (b) Amplitude distribution of voltage and
current on half-wave dipole. (c) Voltage phase distribution (solid) and current phase (dashed).
l /2
FIGURE 4–11.
Instantaneous distributions of half-wave-dipole current (dashed lines) and voltage (solid lines)
at various times (t) during one rf cycle (period T = 1/f ). (At t = 0, T/2, and T the voltage is
zero everywhere, and at t = T/4 and 3T/4 the current is zero everywhere.) The t = T diagram is
identical to the t = 0 diagram, indicating that the cycle is complete and starting over.
it is equal to 1/f, where f is the frequency in Hertz. The patterns are drawn assuming that
at zero time (t = 0) the current is at its instantaneous maximum value, corresponding to
a = p/2 in (4–1).
Linear antennas longer than a half wavelength are also used. They may or may not be
fed at their centers. On each side of the feed point, the current and voltage distributions
are determined by (4–9) and (4–10) above. For correct operation of such an antenna,
Current and Voltage in Longer Antennas 135
radiation resistance of the antenna and by the fact that the antenna wires are not equiva-
lent to a uniform transmission line. The radiation resistance, as well as any actual resis-
tance in the antenna wire, results in a small component of current that is in phase with
the voltage, rather than 90 degrees out of phase. But the sinusoidal approximation is quite
good for linear antennas whose conductors are very thin compared to their length, and
of high conductivity. (It is also assumed that the antenna wire is not close to any large
irregular conducting bodies or dielectric material that would disturb the uniformity of
the electrical environment. In fact, a free-space environment is assumed, but the assumed
distributions apply reasonably well in practical situations.)
Antennas may also be designed to have uniform current and voltage amplitudes along
their lengths, that is, no standing waves. This result is achieved by terminating the end
of the antenna wire in a resistive load so that no reflection occurs. In one form of such
an antenna (Beverage or wave antenna), the wire runs approximately horizontally above
the earth, and the input terminals consist of one end of the wire and the ground. The ter-
minating resistor is connected between the other end of the wire and the ground. In
another form (rhombic antenna) long wires form an array in the shape of a diamond
(rhombus) in a horizontal plane. The two sides of the diamond are fed at one vertex, and
the terminating resistor is connected between them at the other vertex. The current and
voltage are approximately constant along the wires, but there is a gradual decrease of
both with increasing distance from the feed point, owing to the radiation losses and the
ohmic loss in the wire. The current and voltage are in phase with each other everywhere,
rather than approximately 90 degrees out of phase as with standing-wave distributions,
but their phases change linearly with distance along the wire in the amount of 2p radians
or 360 degrees for every wavelength. This description is characteristic of traveling waves,
as described by (1–2) and (1–3) of chapter 1 for waves in space, and by (2–3) of chapter
2 for waves on wires. Antennas having traveling-wave current and voltage distribution
are called nonresonant antennas or traveling wave antennas.
cos π cos θ
60 I 2
E (r , θ , φ ) = (4–11)
r sin θ
where E is in rms volts per meter if r is in meters and I is the rms current in amperes at
the center of the dipole. This pattern is seen to be a slightly more complicated mathemati-
cal expression than that of the short or elemental dipole of (4–2), but the patterns are
only slightly different. They are compared in Fig. 4–14. The half-wave dipole has a
slightly narrower beamwidth—78 degrees compared to 90 degrees for the short dipole.
Consequently its directivity is slightly greater—1.64 compared to 1.5 for the short dipole.
The power-density ratio is 1.64/1.5 = 1.093, and the field-strength ratio 1.047.
The slightly greater directivity of the half-wave dipole is thus almost insignificant. Its
advantage lies primarily in its increased radiation resistance and reduced or nonexistent
feed-point reactance. The radiation resistance for an exactly half-wavelength dipole is
found, by the method illustrated in the case of the elemental dipole, to be 73.1 ohms,
referred to the maximum current point (dipole center). Therefore this is also the resistive
component of the input impedance when the dipole is fed at the center. There is also a
small reactive component of 42.5 ohms, inductive. This small inductive reactance may
138 Basic Radiators and Feed Methods
length. Such antennas are not properly called dipoles. On the other hand, the half-wave
antenna is commonly called a half-wave dipole, even though a true electric dipole is
equivalent to two equal and opposite polarity point charges separated by a definite dis-
tance. The elemental dipole and the short dipole are, in essence, equivalent to oscillating
electric dipoles, but longer antennas are not. However, usage sanctions the term for the
half-wave dipole.
The radiation patterns of long-wire antennas may be determined by the method
described for the half-wave dipole by considering them to be composed of end-to-end
infinitesimal dipoles. The current amplitude and phase in each infinitesimal dipole are
taken to be the values indicated by the current distributions calculated from (4–9), and
as shown for a one-wavelength wire in Fig. 4–12 for a resonant antenna. For a nonreso-
nant antenna a constant current amplitude along the wire is assumed, with a linear phase
change corresponding to a traveling wave of current (2p radians or 360 degrees per
wavelength). These current-distribution assumptions are valid for a thin wire of perfect
conductivity, ignoring the effect on current distribution of the radiation losses. Therefore,
the results are approximate, but useful in that they indicate the general nature of the
radiation patterns.
cos nπ cos θ
60 I 2
E (r , θ , φ ) = (4–12a)
r sin θ
where n is the number of half wavelengths in the wire length, assumed to be an odd
number, and as usual E is in rms volts per meter if r is in meters, and I is the rms current
140 Basic Radiators and Feed Methods
q = 0°
q = 0° Wire
axis
Wire
axis
q = 90° q = 90°
q = 180°
n=3
(a) q = 180°
n=4
(b)
FIGURE 4–15.
Patterns of resonant long-wire antennas.
sin nπ cos θ
60 I 2
E (r , θ , φ ) = (4–12b)
r sin θ
The nature of these patterns is shown in Fig. 4–15 for an odd and an even number of
half-wavelengths.
In these formulas, it is assumed as usual that the antenna is located at the origin of a
spherical coordinate system with the axis of the wire along the q = 0° axis (z-axis, Fig.
3–1) and that I is the rms current at a current-maximum point of the sinusoidal standing
wave. The patterns shown are for long wires of modest length (n = 3 and n = 4). As the
number of half wavelengths is made larger, the number of lobes increases proportionately
and the lobes of maximum radiation lie closer to the wire. Because of the factor “sin q ”
in the denominator of (4–11) and (4–12), an envelope of the lobe pattern is, in three
dimensions, a circular cylinder parallel to the axis of the wire. In a plane containing the
wire axis, the edges of the envelope are straight lines parallel to the wire. The effect is
shown in Fig. 4–16 for a many-lobed pattern.
These patterns, and equations (4–11) and (4–12), are for antennas an integral number
(n) of half wavelengths long. The total length, however, should be shortened by about 5
Long-Wire Antennas 141
The maximum directivity may be attained through knowledge of the maximum field
strength Emax, the radiated (transmitted) power Pt, and other previously developed rela-
tionships. By substituting cosqmax into (4–12a) or (4–12b), according to whether n is odd
or even, the value 60I/(rsinqmax) for the maximum field strength Emax is obtained.* The
power Pt is I2Rr from equation (3–28). Furthermore, the maximum and average power
densities pmax and pav are (Emax)2/377 and Pt/4pr2, respectively. Then, the maximum
directivity Dmax becomes
It is interesting to note that this formula gives the correct result for a half-wave dipole,
Dmax = 1.64, when Rr is taken as 73 ohms and qmax = 90°; also, (4–13) and (4–14) are
correct for n = 1.
* When (4–14) is substituted into (4–12a) and (4–12b), their numerators become, respectively,
60I|cos {(n − 1)p/2}| and 60I|sin{(n − 1)p/2}|. For n odd, |cos{(n − 1)p/2}| = 1, and for n even,
|sin{(n − 1)p/2}| = 1. The denominators become r sinqmax.
142 Basic Radiators and Feed Methods
60 I sin θ π L
E (r , θ , φ ) = sin λ (1 − cos θ ) (4–16)
r (1 − cos θ )
where L is the length of the wire. This pattern has the same number of lobes as a resonant
wire of the same length, and the maxima and minima occur at approximately the same
positions. Their magnitudes, however, are quite different, as shown by the pattern of a
3/2-wavelength nonresonant wire in Fig. 4–17. As seen there, the lobes directed toward
one end of the wire are much larger than those at the other end of the pattern. The lobe
nearest the axis of the wire and pointed in the direction of the traveling wave of current
on the wire is the largest. The smallest lobe is the one at the other end of the pattern, the
magnitudes increasing progressively toward the large-lobe end.
This type of pattern has an advantage when it is desired to radiate or receive in pre-
dominantly one direction, rather than two. Suppression of the pattern in one direction is
accomplished by eliminating the reflected current at the end of the wire by means of a
resistive termination. This usually takes the form of a resistor connected from the end of
the wire to ground. Such termination can be successful, however, only if the height of
the antenna above ground is a very small fraction of a wavelength (otherwise the con-
nection would have reactance as well as resistance and would not be a reflection-free
termination). The correct value of the resistor, being connected between the end of the
wire and ground, is half the value that matches the impedance of a transmission line
consisting of the antenna wire and its earth image. If the antenna height is h and the wire
diameter is d, the resistance is from equation (2–62), chapter 2:
q = 180°
4.4.4. Polarization
The radiation from a long-wire antenna is linearly polar-
FIGURE 4–17.
ized, but the polarization (electric field) direction is not the
Pattern of three-half- same in all parts of the pattern. (This is true even of short
wavelength nonresonant and half-wave dipoles, but for them the variation is not as
long-wire antenna (L = 1.5l).
(Compare with resonant wire great because there is only one lobe perpendicular to the
of same length, n = 3 pattern wire; in the perpendicular plane the polarization is simply
of Figure 4–15.) parallel to the wire.) The polarization in one of the oblique
Long-Wire Antennas 143
lobes of a long-wire antenna, or in fact in any particular part of the pattern, may be
determined by the following procedure (assuming free-space propagation):
(i) Draw a line from the center of the antenna in the direction of interest.
(ii) Form the plane that contains both the antenna wire and this direction line.
(iii) At any point on this direction line the polarization (electric field) vector is
perpendicular to the direction line and lies in the plane thus formed.
(This procedure is in fact applicable to the radiation from any straight-wire radiator,
including short dipoles and half-wave dipoles.)
fields, from the ground, induce voltages along the antenna that add in phase at the receiv-
ing end. Waves propagating along the wire from the opposite direction are, ideally,
absorbed by the terminating resistor. Therefore, the Beverage antenna provides a highly
directive pattern in the horizontal plane for vertical polarization. Beverage antennas
are not ordinarily used for transmitting, because the power absorbed in the terminating
resistor results in poor radiation efficiency.
horizontal. The small loop is an oscillating magnetic dipole, which is equivalent to a bar
magnet whose magnetic strength and polarity oscillate. Because the pattern has the same
shape as the elemental electric dipole, the directivity is the same (D = 1.5) and so is the
half-power beamwidth (BW3dB = 90°).
Now consider Fig. 4–18 and assume an observer is on the z-axis, which extends
outward from the page, and where angle q = 0°. Recall that the current I is assumed
constant and of the same phase everywhere on the loop. Then, the radiation reaching the
observer from any two diametrically opposite located elemental dipoles are equal in
magnitude and opposite in phase, and thus cancel. In other words, at q = 0° the pattern
has zero amplitude, as indicated in Fig. 4–4 for the dipole. Now notice that as q is
increased by being removed from the z-axis, the path length difference (and thus the
phase difference) increases between the radiation from diametrically opposite elemental
dipoles. Therefore, the pattern’s amplitude increases with increases in q. Furthermore,
because of the symmetry of the loop about the z-axis, it is clear that the amplitude is
constant and independent of q. Therefore, analogous to the pattern of a short dipole, the
E-field pattern of a loop antenna is donut shaped.
The polarization can also be determined from considering Fig. 4–18. Let an observer
be anywhere on the x-y plane and far removed from the loop. Then, recalling that the
polarization from each elemental dipole within the loop is aligned with the direction of
its current, one can discern that the vectorial sum of the radiation at each angle q is per-
pendicular to the z-axis. But, because of loop symmetry, the electric field lines are neces-
sarily concentric about the z-axis. Thus, if the z-axis is vertical, the polarization is
horizontal.
The relationship between the loop current and the radiated power is quite different for
a loop, since the radiation resistance of the elemental dipole depends on the ratio of its
length to the wavelength, and the geometry of a loop is not comparable. The formula for
the radiation resistance of a small loop is
31, 200 A2
Rr ≅ ohms (4–18a)
λ4
where A is the area of the loop and l is the wavelength. Since A can be expressed in
terms of circumference C as C2/4p, (4–18a) can also be expressed as
4
Rr ≅ 197
C
(4–18b)
λ
It turns out that the results found for a circular loop apply equally well for a loop of
any shape as long as its dimensions are sufficiently small compared to a wavelength; the
radiation resistance depends only upon its area, following (4–18a). The loop may be
square, triangular, or even irregular in shape.
At very low frequencies it is common to make a loop with more than one turn of wire.
If the number of turns is N, the resulting radiation resistance is found by multiplying the
one-turn value of (4–18a) or (4–18b) by N2. It is necessary that the total length of wire
be small compared to the wavelength if the small-loop behavior is to apply, but if the
frequency is very low (wavelength very long) this requirement is not difficult to meet.
Because of the large wavelength, however, the radiation resistance may still be a very
small value in spite of the number of turns, and the loop may have a significant ohmic
resistance also. Therefore the radiation efficiency will be poor.
When such loops are used for receiving, the terminals of the loop may be connected
to a very high-impedance receiver input circuit, so that the quantity of primary interest
is the voltage induced in the loop rather than the power delivered. When the loop has its
plane in the direction of a properly polarized incoming signal, if the incident-wave field
intensity is E volts per meter, the induced voltage will be
2π NAE
V= volts (4–19)
λ
In both cases, the outer conductor of the coaxial cable is attached to the base plate. Also
in both cases, the center conductor passes through a hole in the conducting base plate,
from which an image is formed. In the monopole arrangement, an extension of the center
conductor serves as the half-dipole above the base plate. However, for the loop, the
extension is bent into a half circle, whereby its image in the base plate forms the other
half of the loop.
For light-weight receiver applications, a low-loss magnetic ceramic (ferrite) is com-
monly used as the core within a multiturn loop to improve efficiency. A reflecting back-
plane (parallel with the plane of the loop) can be used for providing a unidirectional
pattern and increased directivity. Resonant loops (discussed below) are also used as the
elements of phased arrays to increase directivity: an example being coaxially positioned
loops to function as the reflector, driven element, and directors of a Yagi-Uda array
(Fig. 5–11).
Small-loop analysis is applicable to loop wire-lengths of roughly 0.1l or less, so
that the current distribution around the loop is approximately uniform. Larger loops are
also used, especially at the higher frequencies. Ordinarily the patterns of larger loops
have multiple lobes, and the current distributions on these loops affect the patterns
considerably.
The Alford loop, shown in Fig. 4–19, is an example of a larger loop. It is more efficient
than, and has a pattern similar to, that of a small loop. It consists of a square one-turn
loop with quarter-wavelength sides, and it is fed with opposite phases at opposite corners.
The other two corners are capacitively connected. The out-of-phase feed is achieved by
transposing one branch of the feed line as shown. The capacitors are commonly open-end
sections of transmission line; with the reactance given in chapter 2 by (2–24). The radia-
tion resistance of the Alford loop is about 80 ohms. Its radiation pattern is similar (though
not identical) to that of a small loop, but it is much more efficient.
As the circumference C of a loop increases and approaches a wavelength, the peak of
the beam moves toward the loop axis. Near this resonant length, the far field pattern is
nearly the same as two parallel dipoles separated by approximately the loop diameter.
The input impedance varies significantly
with C, having large peaks in both the l/4 Capacitive
stub
resistance and reactance for circumfer-
ences near odd multiples of l /2. Near the Transposed
resonant loop length of l, the input resis- feed line
loop in the range of 0.05l to 0.2l, measurements have given directivities of approxi-
mately 10, and input impedances that can be readily matched (resistance R ≤ 135 W).
The reflector used was square reflector and had different side dimensions between 0.6l
and 1.9l.
As already noted, when the loop circumference is increased to value near one wave-
length, the maximum of the far field pattern is along the loop axis. Resonant loops of
this type are used as the elements of a Yagi-Uda array to form a unidirectional beam
along the axes of the loops and of the array. Details for choosing the dimensions of the
loops and their spacings for a Yagi-Uda array are given by Balanis (2005, p. 599).
Feed wire a
d z (q = 0°)
Feed line y
(coaxial)
f
Ground plane
S
d–diameter of wire
D–diameter of helix
S–turn spacing
f–spacing of helix
from ground plane
a–pitch angle
FIGURE 4–20.
Helical antenna.
Helical Antennas 149
NS, and the circumference C = pD. The length of the wire per turn of the helix is
L = S 2 + C 2 = S 2 + π D 2 . The pitch angle a, an important parameter of the helix, is
the angle that a line tangent to the helix wire makes with the plane perpendicular to the
axis of the helix; and it can be found from this relation: sin a = S/L or tan a = S/pD = S/C.
The properties of helical antennas are described in terms of these geometric parameters.
Many different radiation characteristics may be obtained by varying their magnitudes in
relation to the wavelength l.
The feed wire (Fig. 4–20), which connects the terminus of the co-axial center conduc-
tor to the beginning of the actual helix, lies in a plane through the helix axis and is
inclined with respect to the ground plane at approximately the pitch angle of the helix.
Variation of its geometry affects the input impedance of the antenna.
When the dimensions of the helix are very small compared with the wavelength, the
maximum radiation is in the plane perpendicular to the helix axis, and the radiation
pattern is a combination of the equivalent radiation from a short dipole positioned on the
helix axis and a small loop also coaxial with the helix. This type of helix is known as a
normal mode helical antenna (NMHA), and it is widely used in wireless handsets. To
reduce cost of manufacture, NMHAs have been fabricated by printing a conducting path
into a groove in a small, cylindrical plastic post or mast (see Fig. 36–2, Vardaxoglou and
James 2007).
For a NMHA, the patterns of the two equivalent radiators are of course the same, but
the linearly polarized components are at right angles, and the phase angles at a given
point in space are 90 degrees apart. Therefore, as explained in sec. 1.1.3, the resultant
field is elliptically polarized or circularly polarized, depending on the field-strength ratio
of the two components. This ratio depends on the pitch angle a. When a is very small,
the loop type of radiation predominates; when it becomes very large, the helix becomes
essentially a short dipole. In these two limiting cases the radiated polarization is linear,
in one having loop polarization, and in the other, dipole polarization. For intermediate
values of a the polarization is elliptical, and at a particular value of a it will be circular.
Wheeler (1947) showed that this result is obtained when S = p 2D2/2l, which corresponds
to a value of a given by
πD
tan α = (4–20)
2λ
The analysis of the helix leading to these conclusions may be made by considering it to
be equivalent to a number of small loops having the same diameter as the turns of the
helix, with their planes parallel and their axes in line with the helix axis and spaced the
same as the helix turn spacing. Then these loops are considered to be connected by short
dipoles parallel to the helix axis and of length equal to the helix turn spacing. The radia-
tion field of the helix is equivalent to that obtained by superposition of the fields of these
elemental radiators.
When the diameter and spacing (D and S ) are appreciable fractions of a wavelength,
an entirely different radiation pattern is obtained. The maximum intensity is radiated in
the direction of the helix axis, in the form of a directional beam with minor lobes at
150 Basic Radiators and Feed Methods
oblique angles. The radiation in the main lobe is circularly polarized. It is this feature of
the helix, in this mode of radiation (the axial mode), that probably accounts for most of
the practical applications of this type of antenna.
A typical helical antenna operating in this mode has a circumference C of approxi-
mately one wavelength and spacing S approximately a quarter wavelength. The antenna
will operate quite well over a range of frequency so that these dimensions are noncritical.
The ground plane (which may be either a solid sheet or a wire grid or mesh) should be
at least 34 of a wavelength in diameter. The pitch angle may range from about 12 to 18
degrees; approximately 14 degrees is optimum. The gain and beamwidth depend on the
helix length (equal to NS, Fig. 4–20, where N is the number of turns). The feed-point
impedance is resistive and of the order of 150 ohms at the frequency for which C = l.
At higher and lower frequencies the resistive component varies, and a reactive component
appears. Detailed design information are available (Kraus and Marhefka 2002, ch. 8).
In terms of a three-dimensional spherical coordinate system (Fig. 3–1) with the q =
0° axis coincident with the helix axis, the beam (pattern) has axial symmetry, that is, it
is the same in any plane containing the axis (does not depend on the longitude angle f).
The 3-dB beamwidth, obtained empirically, is given approximately by the formula
52 λ3
θ 3 db = degrees (4–21)
C NS
The formula assumes that the pitch angle is between 12 and 15 degrees, that N is equal
to or greater than 3, that NS (the helix length) is not greater than 10 wavelengths, and
that C is between 0.75l and 1.33l.
The directivity, subject to the same assumptions, is given by
12 NSC 2
Dmax = (4–22)
λ3
pipes, and bars). The most basic of them is the elemental dipole, since in principle all
other current-carrying conductors can be regarded (mathematically and conceptually) as
an assemblage of elemental dipoles, and the radiation field is then deduced by applying
the principle of linear superposition to the fields of the individual dipoles.
Another class of radiators is based on the existence over a surface of a specific elec-
tromagnetic field configuration. The intensity, phase, and polarization of the field over
this surface, or aperture, are analogous to the current amplitude, phase, and direction in
antennas represented by an assemblage of dipoles. The description of the variation of
these field quantities over the aperture is called the aperture distribution. When this dis-
tribution is known, it is possible in principle to calculate the radiation pattern, just as it
is possible for a wire or arrangement of wires in which the current distribution is known.
As for radiation due to currents, the analysis of radiation due to the field distribution of
an aperture is based on Maxwell’s equations.
An example of an aperture over which the field distribution is known is the cross
section of a waveguide, in which a particular known mode is propagating, as described
in sec. 2.5. It is a well-known experimental fact that if such a guide is “sawed off ” in a
plane perpendicular to the axis of the guide, leaving an open end, radiation will occur
from this open mouth. The resulting field pattern can be calculated from the known
configuration of the field for the particular waveguide mode. This is the simplest case of
a waveguide horn radiator.
It should not be surprising that radiation occurs under these circumstances because
the fields inside a waveguide are propagating in essentially the same way that fields
propagate in free space; the only difference is that they are constrained from spreading
spherically by the walls of the guide. When this propagating field reaches the guide mouth
it continues to propagate in the same general direction except that, in accordance with
Huygen’s principle, it also spreads laterally, and the wavefront eventually becomes
spherical, although there is a “near field” region in the vicinity of the mouth of the guide
in which the wavefront is more complicated. It can be thought of as a transition region
in which the changeover from guided propagation to free-space propagation takes place.
This changeover involves a change of phase velocity and a change in the characteristic
wave impedance, from those of the guide to the free-space values, c = 3 × 108 meters
per second and Zc = 377 ohms.
Because the waveguide impedance is ordinarily different from this free-space value,
the radiating open end does not usually present a matched-impedance load to the guide,
resulting in an undesirable standing wave. This can be eliminated by some form of trans-
former matching device, such as those described in sec. 2.5. A better method, however,
is to flare the walls of the guide. If this is done properly, it results not only in a matched
impedance but also in a more concentrated radiation pattern, that is, narrower beamwidth
and higher directivity. This flared structure is what is ordinarily meant by the term horn
radiator.
Various possible flaring arrangements, resulting in different types of horns, are shown
in Fig. 4–21. As shown, a rectangular guide may be flared on the narrow walls, the wide
walls, or both. A sectoral horn is flared in only one dimension. If the flare is in the direc-
tion of the electric vector, as when the broad walls are flared with the TE10 mode in
152 Basic Radiators and Feed Methods
fH dE
dE
f
dH
fE
Sectoral H-plane horn Pyramidal horn dH
L
L
dE
f d
f
Conical horn
dH
rectangular guide, the result is an E-plane sectoral horn; when the narrow walls are flared,
the radiator is an H-plane sectoral horn. Flaring both walls results in a pyramidal horn.
A conical horn is formed by uniform flaring of the walls of a circular waveguide.
If the flare angle f is too great, the wavefront at the mouth of the horn will be curved
rather than plane. This means that the phase distribution over the aperture will be non-
uniform, resulting in decreased directivity and increased beamwidth. On the other hand,
too small a flare angle results in a small aperture area for a given length L of the horn.
The directivity is proportional to the aperture size for a given aperture distribution. Thus,
there is an optimum flare angle that provides the maximum gain for a given horn length,
and therefore the optimum flare angle is intermediate between slight and abrupt. In other
words, an optimum horn or optimum-gain horn is one that compromises in aperture phase
error to maximize its gain for a given length L, and therefore its flare angle is intermedi-
ate between slight and abrupt.
Graphs of radiation patterns are available for a variety of horn types and dimensions
(Balanis, ch. 13, 2005; Love 1993). These graphs, known as universal patterns, were
generated by aperture theory and have been validated, generally, by measurements. The
“universal” patterns provide details of the major lobe shapes for large as well as small
horns, with aperture widths as small as 1.5 or 2l. Even so, designs for center-fed reflec-
tors and lenses often require feed-horn patterns with dimensions even smaller than 1.5l.
(Section 6.3.3 provides guidance on design for horn aperture widths less than 2l).
Horn Radiators 153
It is apparent that the flare angle must be made smaller as the length is increased to
maintain a given maximum phase variation across the aperture. Therefore there is a
practical limit to the gain that can be obtained with a horn radiator; very high gain
requires an excessive horn length. For moderate gains, however, horn radiators are very
useful. They are of course especially appropriate when the feed line is a waveguide. Their
bandwidth is then essentially the bandwidth of the guide—2 : 1 typically, for the TE10
mode in rectangular guide and a sectoral or pyramidal horn.
The published literature on optimum-gain horns includes somewhat different results
for beamwidths versus horn dimensions, depending on how the horn dimensions are
specified and the details of analysis. Kraus and Marhefka (2002, p. 339) provide simple
equations for the beamwidths of optimum horns, that were obtained from analyses of
measured beamwidths versus horn flare angle, for various horn lengths. Those equations
for the 3 dB beamwidths of “optimally” tapered horns follow:
56
θE = degrees (4–23a)
dE
and
67
θH = degrees (4–23b)
dH
where E and H refer to the horn’s E and H plane patterns. The symbols dE and dH are
the aperture dimensions (widths), expressed in wavelengths along the E and H planes.
Use of (4–23a) and (4–23b) does not require detailed horn dimensions, yet they
provide good “first” beamwidth estimates for an “optimumly” tapered horn. As already
mentioned, the universal patterns are considered accurate when detailed horn dimensions
are available. It is to be noted that optimum gain horns are often called standard gain
horns, because of their wide laboratory usage. When detailed horn dimensions are avail-
able and more accurate calculated beamwidths are desired, the reader is referred to Bird
and Love (2007, pp. 14–17). These authors also include an equation for the gain of an
optimum-gain antenna, which follows
6.5dE d H 6.5 A
G= = 2 (4–24)
λ2 λ
where A = dEdH is the area of the horn-mouth opening (aperture). From (3–39) of chapter
3, a recommended formula for estimating gain is
26, 000
G= (4–25)
θ Eθ H
By using (4–23a) and (4–23b) with (4–24), one finds a difference of only 0.3 dB in gain
from that provided by (4–25).
154 Basic Radiators and Feed Methods
φ
sin = λ 2 s (4–26)
2
Horizontal polarization may be excited in this horn by means of a small loop antenna
with its plane perpendicular to the cone axes and lying between the vertexes, and the
loop axis collinear with the cone axes. The optimum flare angle is then given by
φ
sin = 3λ 4 s (4–27)
2
The radiation pattern is omnidirectional in the horizontal plane with the cone axes
vertical, the vertical beamwidth and directivity depending on the dimension h. For
optimum flare angles the directivity is given by
D = m
2h
(4–28)
λ
where for vertical polarization (assuming vertical cone axes) the factor m ⯝ 0.8, and
for horizontal polarization m ⯝ 0.6. The vertical beamwidths may be estimated using
(4–23a) for vertical polarization and (4–23b) for horizontal polarization, with dE and dH
replaced by h.
Slot Radiators 155
Since Zd = 73.1 + j42.5 ohms for a thin half-wave dipole, a thin (narrow) half-wave slot
is found, from the above formula, to have impedance Zs = 363 − j211 ohms. A thin slot
of length 0.475l, the complement of a dipole having nonreactive input impedance of 67
ohms, will have Zs = 530 ohms, nonreactive.
There is no such thing in the real world as an infinite plane conducting sheet, but if a
slot is cut into a sheet that is very large compared to the slot, then the behavior predicted
above will be realized to a high degree of approximation. Slots cut into sheets of even
moderate size will radiate effectively, but their exact behavior is not as readily
predictable.
The slot as described will radiate on both sides of the sheet. If radiation on one side
only is desired, the “back” side of the slot may be enclosed by a box, or cavity. The
field distribution along the slot is then affected by the dimensions of the cavity. The
problem of theoretical design is quite complicated, and design is often determined by
experiment.
A unidirectionally radiating slot may also be obtained by cutting it in a proper position
and orientation in the wall of a waveguide. Figure 4–24 shows the appearance of several
slots in a TE10-mode rectangular waveguide that will radiate, and two that will not radiate.
A waveguide slot, if it is to radiate, must be positioned so that it interrupts currents that
would otherwise flow across its length in the inner walls of the guide. This is equivalent
to saying that there must be a component of magnetic field (H ) parallel to the slot at the
C C
Radiating Radiating
slot slot
C
C
C
Nonradiating
Radiating slots
slot
FIGURE 4–24.
Radiating and nonradiating waveguide slots in rectangular guide, TE10 mode.
Patch or Microstrip Antennas 157
inner surface of the guide. The field configurations in waveguide for various modes are
given in advanced engineering textbooks and handbooks. The examples of radiating slots
shown in Fig. 4–24 by no means exhaust the possibilities. To be most effective the
waveguide slot should be resonant, which it will be if the length is approximately half a
wavelength. However, the exact length for resonance depends on the position of the slot
in the guide.
A waveguide slot does not radiate the total power flowing in the guide; it “extracts”
or “couples out” some fraction of it. The remaining power continues on down the guide
to whatever additional load there may be further on. Or if it encounters a “short circuit”
(end cap), it is reflected. By impedance matching devices it may then be possible to
couple all the power to the slot. Waveguide slots, however, are seldom used singly;
usually an “array” of them is cut along the length of the guide, as will be described in
sec. 5.9.
Slot radiators need not be complements of half-wave dipoles. Many other radiating
shapes are possible, but all may be analyzed in terms of their complementary radiators.
For example, an annular-ring slot in an infinite plane sheet will have a radiation pattern
exactly like that of a loop of the same size, with E-fields and H-fields interchanged. Slot
radiators are practically restricted to the higher frequencies by the requirement of a con-
ducting sheet considerably larger than the slot. They are very useful when a radiator must
be devised that will not project from a surface, for example, in an airplane wing or fuse-
lage. (Here the opening may be covered by a protective cover of low-loss dielectric
material.)
1998; Waterhouse 1999) and electrical sizes small enough for hand-held mobile com-
munications at frequencies less than 2 GHz (Waterhouse, Targonski, and Kokotoff
1998).
FIGURE 4–26.
Block diagram of basic antenna feeding arrangement.
160 Basic Radiators and Feed Methods
transmitter (or, in reception, optimum transfer of received signal power from the antenna
to the receiver).
Sometimes the transmission line can be eliminated altogether by making a direct con-
nection from the antenna to the transmitter output terminals or the receiver input termi-
nals. Here also only one transformer is required. This situation can exist at very low
frequencies, where a “line” of even a few hundred feet in length may be in effect a direct
connection. Any conductor of length less than about a hundredth of a wavelength may
be so regarded, since no appreciable standing wave pattern can exist on a conductor so
electrically short. The direct connection can also exist at higher frequencies when the
antenna is “built into” a receiver or transceiver.
The transformers at low frequencies are inductor-and-capacitor devices. In addition to
impedance step-up or step-down, they also can provide reactance cancellation; both
effects may occur in the same circuit elements or they may be separated. At the higher
frequencies the transformers may be composed of transmission-line or waveguide ele-
ments, as described in secs. 2–3 and 2–5.
At frequencies up to about 30 MHz, two-wire balanced transmission lines can be used,
since at these frequencies radiation due to the line spacing being a significant fraction of
a wavelength is not a serious problem. Line impedances range from slightly less than
100 ohms to perhaps 800 ohms. The lower impedances are achieved with close-spaced
wires embedded in low-loss polyethylene plastic. The higher impedances result with
air-insulated wires or tubing. The wire lines have spacing bars of porcelain or other
insulating material at intervals of a few feet or more, depending on the spacing and the
wire stiffness.
When the higher-impedance lines are used to feed an ordinary half-wave dipole at its
center, or a long-wire antenna at a current maximum, a transformer must be used if the
impedances are to be matched. One possible type of transformer is shown in Fig. 4–27.
Shorting bar
Transmitter-receiver
house
FIGURE 4–27.
An antenna installation for the HF range (3–30 MHz), showing method of center feeding a
half-wave dipole with two-wire balanced line and matching stub. Line section above stub has
standing wave, but with proper adjustment there is no standing wave on the feed line to left of
the stub.
Basic Feed Methods 161
Half-wave or
resonant-long-wire
antenna
Insulators Insulator
Zepp feeder
FIGURE 4–29.
Method of feeding a half-wave or resonant-long-wire antenna at one end by means of a
two-wire line with standing waves (Zepp feeder).
The pure Zepp feeder arrangement, shown in Fig. 4–29, omits transformer A in Fig.
4–26. The line is operated with a standing wave. When the antenna is operated at more
than one of the possible frequencies (values of n) transformer B must in general be
adjusted differently for the different frequencies. In the HF band, where such antennas
are most commonly used, this transformer typically consists of a variable inductance-
capacitance circuit (ARRL Antenna Book 2007). This reference and its earlier editions
contain much practical information on antennas and feed systems of the type described
in this chapter and on some of those described in chapter 5.
A matching circuit (i.e., transformer A) can also be incorporated at the antenna end
of this type of feeder, to eliminate standing waves on the line. The stub arrangement of
Fig. 4–27 (and Fig. 2–6, chapter 2) is used for this purpose. When a half-wave vertical
dipole is fed in this way, the resulting arrangement is called a “J” antenna. This matching
will in general be effective at only one frequency.
As the Zepp feeder indicates, “balanced” two-wire lines may be used to feed an unbal-
anced antenna, though at the sacrifice of perfect balance of the line currents. (Therefore
such lines will radiate somewhat more than would a perfectly balanced line.) In general,
however, balanced two-wire lines are preferred when the antenna is fed at a point of
symmetry—where the structure is electrically balanced with respect to the ground on
both sides of a center feed point. But when one side of the feed point is the ground, or
a metallic ground plane, the favored transmission line for feeding the antenna is coaxial
line. Examples of such antennas are monopoles, antennas whose axes are perpendicular
to a ground plane in which they are imaged. All the feed methods described for two-wire
lines are applicable in these cases, except that the center conductor of the coaxial line
connects in the manner indicated for one of the conductors of a two-wire line, and the
coaxial outer conductor connects to the base of the antenna (ground).
Transmission-line feed is appropriate to such radiators as dipoles, long wires, loops,
and helixes, whose radiation is based on currents flowing in wires. Waveguide feed is
more appropriate for horns and waveguide-slot antennas. However, these “rules” have
Basic Feed Methods 163
current flows vertically along the outer surface of the outer coax conductor. This vertical
current creates vertically polarized radiation, yet the radiation of the dipole, per se, is hori-
zontal. Therefore, without the l /4-length “bazooka” sleeve, the effective antenna gain is
reduced because the antenna will simultaneously radiate horizontally and vertically polar-
ized waves. The essence of this balun is the l /4 length conducting coaxial section, formed
by the larger diameter conducting sleeve and the outer conductor of the coax line, with the
l /4 section open at its top and connected its bottom to the outer conductor of the coaxial
line. In other words, looking from the top and downwards, the impedance of the l/4 coax
(formed by the sleeve and the outer conductor of the coax line) appears as an open circuit.
Thus, the vertically moving outer coaxial-line surface current, and hence the vertically
polarized radiation, is appreciably reduced.
Figure 4–31c shows a half-wave-line balun for coupling between the bottom coaxial
(coax) line and the balanced two-conductor line. To recognize how this balun functions,
it may be helpful to recall that the voltages between points on a transmission line are
equal if separated by a distance l /2, but opposite in phase. Then, one will recognize that,
at the ends of the l/2-length coaxial line, the voltages are of equal amplitude but of
opposite phase. Therefore, the voltage across the balanced line is twice that at the input
coaxial line. This balun may provide a 4-to-1 impedance transformation. First, note that
the two coax lines formed by the l /2 length line are in parallel across the input coax.
Thus, an impedance-matched condition may exist if the l/2 length coax characteristic
impedance is 2Z0, where Z0 is the characteristic impedance of the input coax line. Note
that the double voltage across the two output conductors is consistent with the ends of
the l /2 length line being in series with one another. Thus, an impedance matched condi-
tion exists if the two-conductor output line is terminated with an impedance of 4Z0, where
Z0 is the characteristic impedance of the input coaxial line.
Figures 4–31(b) and (c) show only simple baluns that provide connections between
coaxial and two-conductor balanced lines. Munk (2002) discusses numerous other balun
configurations; including those that use multiple transmission lines for increased band-
width, as well as baluns where connections are made between coax and balanced printed
lines, and between unbalanced and balanced printed lines.
References
ARRL Antenna Book, 21st ed., American Radio Relay League, 2007.
Balanis, C. A., Antenna Theory: Analysis and Design, 3rd Edition, John Wiley & Sons,
2005.
Beverage, H. H., C. W. Rice, and E. W. Kellog, “The Wave Antenna: A New Type of
Highly Directive Antenna,” Transactions of the AIEE, vol. 42, 1923, p. 215.
Bird, T. S., and A. W. Love, “Horn Antennas,” ch. 14 in Antenna Engineering Handbook,
4th ed., J. L. Volakis (ed.), McGraw-Hill, 2007.
Brainerd, J. G., ed., Ultra-High-Frequency Techniques, Van Nostrand, New York, 1942,
p. 415. (The formula is there given as Rr = 72.5 + 30logen.).
Carver, K. R., and J. W. Mink, “Microstrip Antenna Technology,” IEEE Transactions on
Antennas and Propagation, January 1981, pp. 2–24.
Problems and Exercises 165
which is at a distance from the center equally 90 percent of the distance from
the center to the end? (c) What is the current at the end of the dipole?
4. A half-wave dipole in free space is center-fed with a current of 10 amp (rms).
At a distance of 1,000 meters from the dipole, which is in the far field, and in a
direction that is 60 degrees from the dipole axis, what is the field strength in
volts per meter?
5. A medium-frequency (MF) radio station (for which a long-wire horizontal
antenna is practical) is required to communicate with only four other stations.
These stations are in four different directions; one is due north, one due south,
one east, and one west. It is desired to utilize a single horizontal long-wire reso-
nant antenna that will have four major lobes, one directed at each of the four
other stations. (The major lobes are those nearest the wire axis.) (a) How many
half wavelengths long should this antenna be? (b) What are the two possible
directions of the antenna wire? Suggestions: Assume that equation (4–14) is valid
for the value of n involved. Draw a diagram showing the relative positions and
directions of the stations, and consider possible orientations of the antenna wire.
Determine from this the required value of qmax in equation (4–14). Then, using
equation (4–14), solve for n. The value obtained for n will not be an exact integer;
the correct value is the integer nearest to the value found.)
6. A helical antenna with a ground plane (as in Fig. 4–20) has a turn diameter and
spacing that are appreciable fractions of a wavelength, so that it radiates in the
axial mode. The circumference of a turn, C, is equal to the wavelength, l, and
the turn spacing S is equal to 0.25l. The number of turns is N = 16. (a) What is
the 3-dB beamwidth of this antenna? (b) What is its directivity?
7. A rectangular waveguide pyramidal horn has aperture dimensions, dE and dH of
Fig. 4–21, of one wavelength. (a) Determine approximate E- and H-plane half
power beamwidths. (b) What is the approximate directivity?
8. A slot in an infinite plane sheet of metal is in the form of an annular ring so that
it is the complement of a small loop antenna. (The central portion of the sheet,
inside the annulus, is supported by low-loss insulating material.) (a) Is the
maximum radiation of this annular slot in the direction perpendicular to the plane
of the sheet, or is it parallel to it? (b) What is the polarization of the field near
the metal sheet at a point distant from the slot?
9. A thin vertical monopole antenna is 95 percent of a quarter wavelength in height
above a ground plane of infinite extent and perfect conductivity. It is fed at a
gap at its base by a coaxial line. No transformer is used at the feed point, yet
there are no apparent standing waves on this line. What is the approximate
characteristic impedance of the line?
10. Several means of delivering power to an antenna from a transmitter are listed
below, and identified by capital letters:
(A) Coaxial line.
(B) Balanced two-wire line followed by an impedance transformer to provide
nonresonant operation (no standing waves).
Problems and Exercises 167
(C) Waveguide.
(D) Direct connection.
(E) Resonant two-wire line.
After each of the antenna types below, write in one of the above capital letters
to indicate which form of line or connection you consider most appropriate. (Use
each letter once and only once.)
(i) Automobile radio antenna for broadcast-band reception ⵧ (535–1,605 kHz)
(ii) Helical antenna with ground plane ⵧ
(iii) Long-wire antenna fed at one end ⵧ
(iv) Horn radiator ⵧ
(v) Half-wave dipole having 20-meter length, fed at gap in center ⵧ
CHAPTER 5
Arrays
This chapter discusses array antennas that are designed so that the major lobe (the main
beam) is pointed in one fixed direction or more. Commonly, array antennas are also
designed to electronically, and thus almost instantly, change beam pointing direction.
Such a beam pointing antenna is called a phased array, electronic scanning array, or
electronically steered array (ESA). The general principles of arrays are discussed in the
present chapter, and chapter 8, titled Electronically Steered Arrays, focuses on concepts
and techniques more directly applicable to electronic beam steering.
The term array, as applied to antennas, means an assembly of radiating elements in
an electrical and geometric arrangement of such a nature that the radiation from the ele-
ments “adds up” to give a maximum field intensity in a particular direction or directions
and cancels or very nearly cancels in others. Obviously this principle can be used for the
design of a directional antenna with increased gain.
The long-wire antennas discussed in sec. 4.4 may in a sense be regarded as arrays,
since they are analyzable as an assembly of elemental dipoles in a geometric configura-
tion that provides directionality and gain. (The resonant long-wire antennas also might
be regarded as an array of half-wave dipoles.) The term array, however, is usually
reserved for arrangements in which the individual radiators are separate rather than part
of a continuous radiator.
169
170 Arrays
uniform intensity in all directions and has no physical size, and here it is assumed that
it does not block or otherwise affect the radiation of the other elements of the array. An
array radiation pattern can be calculated on the basis of these assumptions, and then a
correction to it can be made to take into account the fact that the individual radiators in
practical cases do affect each other and do not radiate isotropically.
r1
r r2
f1
Radiator 1
f
d x
f2
Radiator 2
FIGURE 5–1.
Array of two isotropic point sources.
to all points and thus defines the radiation pattern of the array. Calculation of the field
due to the array at an arbitrary point P is basic to the understanding of array theory. This
calculation is here made for the two-element isotropic-point-source array, with the two
sources of equal strength (intensity).
As shown, the distance from Radiator 1 at point R1 to P is r1, and from Radiator 2 at
point R2 to P the distance is r2 (R1 and R2 are shown in Fig. 5–2). The two sources are
assumed to lie on the y-axis of a cartesian coordinate system, and P is assumed to be in
the xy-plane. Therefore, in terms of the corresponding spherical coordinate system
(Fig. 3–1), the direction of P from R1 is f1, and from R2 it is f2.
The distances r1 and r2 are assumed to be very large compared to the distance d, the
separation of the two radiators. If the difference of these two distances to the field point
is d = r2 − r1, the maximum possible value that d can have, whatever the location of P,
is equal to d, the radiator separation. (This equality will occur when P lies on the y-axis;
i.e., when f1 = f2 = 90° or 270°.) This means, since both r1 and r2 are very much larger
than d, that d will always be very much smaller than either r1 or r2. The very important
conclusion that may then be drawn is that if the two radiators are of equal strength, the
amplitudes of their separate fields at P will be very nearly the same; that is, both will
be reduced in strength, because of the distance traveled, by virtually the same amount.
172 Arrays
(This reduction in strength due to the spherical spreading of waves radiated from point
sources in free space was discussed in sec. 1.1.6).
On the other hand, the relative phases of the two fields at P, from R1 and R2, will be
very importantly affected by d. The resulting phase difference of the fields due to d is in
fact equal to −2pd/l radians or −360d/l degrees. To this difference must be added the
initial phase difference, that is, the phase difference of the two radiators themselves, a.
The total phase difference y of the two fields at P will be*
2πδ
ψ =α − radians (5–1)
λ
The resultant field at P is determined by the superposition of two fields of equal ampli-
tude, which will be denoted E0, and of phase-difference y. This involves phasor addition,
and it is mathematically the same as the addition of vectors separated in direction by the
angle y. The magnitude of this phasor addition was given in sec. 1.2.4, and for the case
of equal amplitudes the result was found from (1–28) to be
ψ
E = 2 E0 cos (5–2)
2
where, at observation point P, E0 is the amplitude of each field and y is the phase dif-
ference of the two fields.
This is a first step toward finding the radiation pattern of the array, which is an expres-
sion of E as a function of f, the angle of the direction of P from the center of the array.
To obtain such an expression, it is evidently necessary to express y in terms of f. Equa-
tion (5–1), an expression for y in terms of d and a, is already available. Since a is a
(presumably) known quantity (the phase difference of the array elements), the only vari-
able here is d, which can be expressed in terms of f. To see how this can be done, it is
helpful to consider an enlarged diagram, Fig. 5–2, showing the region in the immediate
vicinity of the array. As this diagram shows, because P is so distant in relation to the
element separation d, the lines labeled r1, r2, and r can be considered almost parallel to
each other, so that also f1 = f2 = f to a very close approximation. If a construction line
R1Q is drawn as shown dashed, so that it is perpendicular to the line R2P (also designated
r2), the distances R1P and QP can be considered equal. Therefore, the distance R2Q is
the difference, d, between R1P and R2P (i.e., between r1 and r2). The angle R2R1Q is
equal to f, since R1R2 is perpendicular to the x-axis and R1Q is perpendicular to R2P.
The triangle R1QR2 is a right triangle, with base d, hypotenuse d, and angle f opposite
the base. Therefore,
δ
sin φ = (5–3)
d
* In (5–1) y and a denote phases of Radiator 2 field and current with respect to those of Radiator 1, that is,
the Radiator 1 phase is the reference phase.
Basic Array Theory 173
r1
f1
P
R1
f
f
d x
r2
f2
R2 d
Q
FIGURE 5–2.
Enlarged portion of two-element-array geometry.
or
δ = d sin φ
Substituting this result into (5–1), and then substituting the resultant expression for y
into (5–2), gives the following equation for the magnitude of the field at P:
α π d sin φ
E (φ ) = 2 E0 cos − (5–4)
2 λ
This is the desired expression for the magnitude of the field at P as a function of the
angle f that the direction of P makes with the line perpendicular to the line of the array.
This equation gives the shape of the pattern in the xy-plane (which may also be called
the q = p/2 plane). Recall from (5–1) that y is expressed in radians. Thus, for calcula-
tions using degrees, p in (5–1) and (5–4) is replaced by 180°, because p radians = 180°.
The absolute-value brackets are used to indicate that the field intensity being calculated
is proportional to the amplitude or to the rms value and is therefore a positive number,
174 Arrays
although the quantity inside the brackets may be sometimes positive and sometimes
negative.
In order to obtain the relative pattern, for which the field strength in the maximum-
intensity direction has the value unity, (5–4) must be divided by the maximum value of
E. In this example E = 2E0, because the separate fields of each radiator are assigned the
intensity E0.
It is evident that this pattern depends on the spacing of the elements d and their phasing
a, as stated at the outset. It is worth noting that the result just presented for a two element
array is very similar to that of the field due to an antenna and its image in a conducting
earth, which was considered in sec. 1–4. The differences are that in the antenna-and-
image configuration, the phase difference of the two has the fixed value p radians or 180
degrees, and the distance designated 2h in Fig. 1–14 may be much larger than d is usually
permitted to be in an array (although sometimes 2h may be comparable to d ). Otherwise
the two situations are the same, mathematically. Physically there is still another differ-
ence, in that the interference pattern of the antenna-plus-image occupies only half of
three-dimensional space, the other half being occupied by the reflecting earth.
α π d ⋅ sin θ ⋅ sin φ
E (θ , φ ) = 2 E0 cos − (5–4a)
2 λ
The quantity sin f in (5–4) is merely replaced by the product sin q ⋅ sin f. In the xy-plane,
when q = 90°, sin q = 1; hence (5–4a) becomes (5–4) in this plane, as it should.
axes are parallel to the z-axis. In the xy-plane each dipole considered individually will
have an omnidirectional pattern, that is, uniform radiation in all directions in that plane.
Consequently the dipole-array pattern will be represented by (5–4) in the xy-plane only.
These two-element dipole-array patterns for a number of values of d and a within the
ranges of greatest practical interest are shown in Fig. 5–3, as originally published by
G. H. Brown (1937). They may be thought of as the horizontal-plane patterns of a pair
of vertical dipoles or monopoles separated by a distance d and with currents having phase
difference a.
As shown in Fig. 5–3, the effect of the spacing and phasing of the dipoles is rather
remarkable, and the variations in the patterns are extreme. The light-line circles represent
the E-field pattern that would be obtained from a single dipole located at the midpoint
of the array (with the array dipoles removed) radiating the same total power as the array.
The heavy-line plots are the electric-intensity radiation patterns. A great deal can be
learned about arrays in general by studying these two-element patterns.
It should be noted that when the two elements are in phase (a = 0°), the radiation has
a maximum in the direction perpendicular to the line joining the elements. That is,
because the distances from the elements to the field point (P, Fig. 5–1) are equal in that
direction (f = 0°), the phase difference due to path difference is zero also; hence the total
phase difference of the superimposed fields is y = 0°. Consequently, the fields add
directly, and the maximum possible resultant field is obtained. When a = 0° and a pattern
maximum is in the direction perpendicular to the array line, the antenna is called a
broadside array.
It should also be noted that for certain conditions the resultant field in some directions
is zero—when the sum of the radiator phase difference (a) and the phase difference due
to path difference (2pd/l) is an odd integral multiple of p radians. The fields of the indi-
vidual radiators are in this case of equal amplitude and opposite phase, so they cancel.
This occurs, for example, when a = 0° and d = l/2, in the f = 90° and 270° directions,
since in these directions the field phase-difference due to path difference is 180 degrees,
whereas the phase difference due to radiator phase difference is zero. The same result
occurs when the 180-degree net phase difference is due to a combination of path differ-
ence and radiator phase difference. These directions of zero intensity in a pattern are
called nulls.
Finally, it is to be noted that certain combinations of d and a result in maximum radia-
tion in the direction of the line joining the array elements. The array is then said to be
operating as an endfire array. The radiation of an endfire array may be either bidirectional
(radiation lobes in both directions along the line of the array) or unidirectional (a lobe
in one direction and a null in the opposite direction). Examples of unidirectional endfire
arrays are the case of d = l /8 with a = 135°, and the case of d = l/4 with a = 90°. When
d is greater than l /2, the array will have lobes in more than two directions, and may
even be simultaneously broadside and endfire.
Although the two-element array is the simplest of all arrays, the variety of patterns
obtainable makes it useful in many applications. In radio broadcasting in the MF band,
for example, two vertical towers with appropriate spacing and phasing can be used to
produce a horizontal-plane pattern (possibly one depicted in Fig. 5–3) to favor certain
176 Arrays
7/8
3/4
5/8
1/2
d/l
3/8
1/4
1/8
receiving element (as well as in the terminated elements) will be the same as in actual
array operation, including mutual-coupling effects. The pattern measured under these
conditions is in general different from the pattern of an isolated element. Although much
of this discussion has been in terms of dipole elements, mutual coupling occurs with all
types of array elements.
In the practical design of arrays, the effects of mutual coupling are often evaluated
experimentally, because theoretical calculations are usually very difficult, although the
theory is helpful in understanding the general nature of the effect. The electromagnetic
analysis of element coupling involves complex mathematics and detailed use of
Maxwell’s equations. Such analysis will of course depend on the shape and size of the
elements and on their electrical properties, element spacing, locations of the elements
within the array aperture, wavelength, polarization, and the amplitudes and phases of the
element input currents. The total of the electric and magnetic fields near an element
includes the fields generated by its currents and the coupling from reactive near-fields
from the other elements and possibly from nearby objects. For antennas with dimensions
of many wavelengths, the fields of an element may also include significant radiating
near-field components from other elements. Computer codes are used to compute the
effects of mutual coupling among the elements of an array (Hansen 2007, pp. 20–24).
where
V1, V2, etc. = voltages applied to elements 1, 2, etc.
I1, I2, etc. = currents flowing in elements 1, 2, etc.
Z11, Z22, etc. = self-impedances of elements 1, 2, etc.
Z12, Z2n, etc. = mutual impedances between elements denoted by the subscripts
The driving-point (input) impedance offered to the voltage applied to the kth element
is the phasor ratio (in general complex) Vk/Ik, obtained by dividing the right-hand side
Basic Array Theory 179
of the above equation by Ik. Thus for k = 1, the input (i.e., driving-point) impedance
follows:
V1 I I
Z1input = = Z11 + 2 Z12 + 3 Z13 + . . . (5–6)
I1 I1 I1
From above, it is seen that the input or driving-point impedance of a particular element
is not only a function of its own self-impedance, but also a function of the relative cur-
rents flowing to the other elements and possibly in nearby objects. Thus, it is important
in an array to design the feed system to match the input impedances instead of simply
the self-impedances.
Radiation resistance is the resistance Rr that, inserted at the point where a current I is
flowing, dissipates the same energy that radiates from the antenna. Thus Rr equals radi-
ated power divided by I 2 (sec. 4.1.6). In an array composed of several radiating elements,
the total radiation resistance of an element may be expressed in terms of the radiation
resistance of the individual element and its mutual impedances. Specifically, the effective
radiation resistance encountered by a voltage V1 applied to element 1 of an antenna array
is the resistance component of the driving-point (input) impedance Z1input offered to V1,
as calculated by (5.6). The power radiated by element 1 delivered by the voltage V1 is
the square of the current multiplied by this effective radiation resistance. Likewise, the
total power delivered to an element is the square of the current at the input multiplied
by the resistive component of Z1input. This resistive component consists of the radiation
resistance Rr, plus whatever additional resistance is needed to account for the dissipated
(heat) losses that may arise from the presence of earth, and from the conductors and
dielectrics within the antenna. The radiation resistance to the voltages applied to other
elements in the array and the power delivered by these other voltages is determined in
the same way. The total radiated power is then the sum of the powers supplied by the
various applied voltages. Likewise, total power input to the elements is the sum of the
powers delivered to all elements.
That is, the pattern of the array of actual elements is obtained by multiplying together
the pattern of a single element and the pattern of the array as calculated for isotropic
point-source elements. The function EA(q, f) is sometimes referred to as the antenna
pattern factor, which is the product of the element factor Ee(q, f) and the array factor
Ea(q, f). Squaring both sides of this equation converts it into a relationship of power-
density patterns, in which the squares of the electric-intensity factors may be replaced
by corresponding power-density pattern factors, pA(q, f), pe(q, f), and pa(q, f), in accor-
dance with equation (1–9), chapter 1.
The factor kn is included in this equation as an arbitrary numerical constant to be
adjusted to whatever value is required to make EA(q, f) a true relative pattern. That is,
it is desired that EA = 1 for the particular values of q and f in the direction of maximum
field intensity, in accordance with the definition of a relative pattern. If both Ee and Ea
are relative patterns, it may happen that EA will automatically be a relative pattern also,
that is, with kn = 1. This will happen only when a maximum direction of Ee coincides
with a maximum direction of Ea, which it will not necessarily always do, although it
often does. The factor kn is called a normalizing constant.
As an example of the application of (5–7), the pattern versus q and f will be deter-
mined of a two-element array of dipole radiators, spaced a distance d apart on the y-axis,
with phasing a. For the case of isotropic radiators instead of dipoles, the pattern is given
by (5–4a). The pattern of a half-wave dipole with its axis parallel to the z-axis is given
by equation (4–11), sec. 4–3. (This is the complete dipole pattern, even though it involves
only the angle q rather than both q and f, because the pattern for this particular dipole
orientation is independent of the angle f. That is, the pattern in any plane corresponding
to a particular value of f is exactly like the pattern in a plane corresponding to any other
value of f.) The complete pattern follows in (5–8). It is a function of q and f, of a two-
element dipole array with this orientation of the dipole axes and is attained by multiplying
the patterns of (4–11), sec. 4–3 and (5–4a) together:
⎡ cos ⎛ π cos θ ⎞ ⎤
α
E A(θ , φ ) = kn cos ⎛⎜ −
π d ⋅ sin θ ⋅ sin φ ⎞ ⋅⎢ ⎝2 ⎠⎥
⎝2 ⎟
⎠ ⎢ ⎥ (5–8)
λ ⎢ sin θ ⎥
⎣ ⎦
element excitation currents are critical to the control of the overall array patterns. These
quantities are needed when calculating array factors, yet they do not include, explicitly,
effects of mutual coupling. Furthermore, these are the currents needed, when effects of
mutual impedance are included, to determine the required element input voltages and
radiated power. One must also recognize that element beam shapes change, because of
being within an array environment. However, because of the dominating effect of the
array factor on overall array patterns (sec. 5.1.6), useful pattern shape approximations
are often possible without the use of measured in-array “active” element patterns.
f = 90°
y –z P
Array
elements r1
r2
d
d/2
d x f = 0°
r3
d
r4
Array
elements
FIGURE 5–4.
Linear array of four radiating elements.
Multielement Uniform Linear Arrays 183
π d sin φ α
sin N −
λ 2
Erel = (5–9)
π d sin φ α
N sin −
λ 2
where f, as in (5–4), is the angle between the direction of the field point and a perpen-
dicular to the line of the array. For N = 2 this expression can be shown to be equivalent
to (5–4) (despite the seeming dissimilarity). As in the two-element case, the pattern versus
q and f is obtained simply by replacing sin f by the product sin f ⋅ sin q. The factor N
in the denominator is a normalizing factor, as discussed for (5–7); that is, in this case
kn = 1/N.
The pattern has its maximum values at angles f such that the quantity (pd sin f)/
l − a /2 = 0; that is, when (pd sin f)/l = a /2. The pattern will have additional maxima
at angles for which this quantity is equal to an integral multiple of p radians. At the
maxima, (5–9) is equivalent to
Erel = lim
sin Nx
lim
sin Nx
or
N sin x
(5–10)
x → 0 N sin x x → mπ
where m is any integer. For either limit, this expression is an “indeterminate form” (zero
divided by zero), but application of a method of differential calculus shows that the limit
is unity. The pattern has secondary maxima, or minor lobes, when the numerator of (5–9)
attains values expressed by sin[(2m + 1)(p/2)], where again m is any integer (i.e., m = 1,
2, 3, . . .).
The variety of possible patterns for an array with a given number of elements, obtained
with different values of d and a, is very great, but many of the possibilities are primarily
curiosities, of no practical value. Two basic types of pattern are of special interest.
l /4 l /2
Quarter-wave
stubs
Feed line
FIGURE 5–7.
A six-half-wave Franklin antenna (six-element collinear dipole array).
The in-phase currents in the individual dipoles, required for a broadside pattern, may
be obtained by properly connecting a branched transmission line to the feed point of each
dipole. In other words, if the total line length from the transmitter to each dipole is the
same, as in Fig. 2–8a, chapter 2, the dipoles will be fed in phase. (Care must be taken
to see that the same side of the line is connected to the same side of each dipole; revers-
ing this connection reverses the phase.)
In an alternative feed arrangement, shown in Fig. 5–7, the feed line is connected
between the ends of one pair of dipoles, which causes them to be in-phase. (This is a
high-impedance feed point, since the ends of the dipoles are voltage-maximum points.)
Then quarter-wave stubs are connected between the ends of the additional adjacent pairs
of dipoles, providing simultaneously a feed connection and proper phasing. The l /4 stubs
provide open circuit impedances, and their l/2 round-trip lengths cause a phase reversal
at each of their input terminals. This arrangement is known as a Franklin antenna. (Recall
that if the dipoles are directly connected end to end, without the quarter-wave stub, the
current in alternate half-wave sections will be out of phase; the antenna will be of the
long-wire type, and the pattern will be entirely different.)
The collinear array can also be fed by transmission lines connected between the ends
of pairs of adjacent dipoles. This has an advantage over center-feeding each dipole indi-
vidually, in that only one transmission-line branch is required for each pair of dipoles.
Also, the relatively high feed-point impedance is usually more easily matched to a two-
wire line. The dipole spacing is then restricted to one-half wavelength (as is also true for
the Franklin antenna), but this is a fairly satisfactory spacing.
Broadside arrays may also be formed from other types of elements, such as horns,
slots, microstrip patches, helixes, and polyrods. If the individual elements are themselves
directional, their maximum radiation should be in the broadside direction. If they are
unidirectional radiators, such as sectoral or pyramidal horns, waveguide slots, microstrip
patches, axial-mode helixes, and polyrods, a unidirectional broadside array results. When
186 Arrays
2π d π
α = + (5–11)
λ N
where N is the number of elements in the array. This relationship is referred to in the
antenna literature as the Hansen-Woodyard condition. (This condition does not neces-
sarily result in a unidirectional pattern, however, as does the basic endfire condition.)
The correct phase of current or field to each radiating element of an endfire array
requires a feed system in which the current or field must travel through a longer path in
the transmission line or waveguide to each successive element along the array, the
increased length of line or guide per element being equal to al/2p, where a is the
Multielement Uniform Linear Arrays 187
Dipoles
Direction
of beam
Transmission
line d
Input
terminals
FIGURE 5–9.
One form of feed arrangement suitable for an endfire array.
α = π 1 ± radians
2d
(5–12)
λ
(Choosing the plus or minus sign merely corresponds to changing the particular endfire
direction in which the null occurs.) Satisfying this null condition does not insure a
maximum in the other direction. In fact, if d is an odd multiple of a half wavelength,
(5–12) corresponds to an in-phase condition of the elements (a = 0° or an integral mul-
tiple of 360°). This produces a broadside array, with nulls in both endfire directions.
However, with some spacings, of which examples are given above, a maximum does
occur in the direction opposite the null, producing a unidirectional pattern. In fact, the
two-element array with d = l /8 and a = 135° simultaneously satisfies the condition for
a null in the back direction and the Hansen-Woodyard condition for maximum forward
directivity for that particular spacing.
Incidentally, the null condition of (5–12), applies not only to the two-element array
but also to any uniform linear endfire array with an even number of elements, since the
elements cancel in pairs in the null direction. It may also be noted that the null is never
absolute because of unavoidable imperfections in the spacing and phasing of the ele-
ments, and many-element arrays are more susceptible to such imperfections than those
of few elements.
Section 5.5.6 includes further discussion on endfire arrays and Fig. 5–18 that compares
the beamwidths of an ordinary endfire array and an endfire array that satisfies the Hansen-
Woodyard condition.
Driven element
C
Beam direction
Reflector
Directors
Feed line
FIGURE 5–11.
A Yagi-Uda parasitic endfire array with one reflector and five directors.
lag the induced emf. A dipole shorter electrically than a half wavelength will be capaci-
tive, and the current in it will lead the induced emf. Comparatively close spacing of
elements is used in parasitic arrays to obtain good excitation, and the induction fields of
the elements play a major role, so that an exact analysis is very complicated. It is known,
however, that properly spaced dipole elements that are electrically slightly shorter than
a half wavelength act as directors, reinforcing the field of the driven element in the direc-
tion away from the driven element. Thus a line of directors may be used, and each one
will excite the next one. On the other hand, an element that is electrically one-half-
wavelength long or slightly longer will act as a reflector, if correctly spaced, reinforcing
the field of the driven element in a direction toward the driven element from the reflector.
Therefore if a reflector element is placed adjacent to a driven element, another element
placed beyond the reflector will not be appreciably excited, as very little field exists
beyond the first reflector. For these reasons a Yagi-Uda endfire array usually consists of
one driven element, one reflector on one side of it, and a number of directors on the other
side of it.
A typical Yagi-Uda configuration is shown in Fig. 5–11. Antennas of this type offer
the advantages of a unidirectional beam of moderate directivity with light weight, sim-
plicity of feed-system design, and low cost. The design becomes critical, however, if
high directivity is attempted through the use of many elements. Up to five or six may be
used without difficulty, and arrays of thirty or forty elements are possible. The input
impedance of a Yagi-Uda array tends to be low, and the bandwidth is limited to around
2%, typically. Directivity of around 10 dB is readily achieved with a moderate number
of elements (five or six). Higher gains may be achieved by making a broadside array of
which the elements are Yagi-Uda arrays.
190 Arrays
π d sin φ sin θ ⎞ ⎤
sin ⎡⎢ N h⎛ h
z
⎣ ⎝ λ ⎠ ⎥⎦
Eh(θ , φ ) =
π d sin φ sin θ ⎤
N h sin ⎡⎢ h
q ⎣ λ ⎦⎥
dh (5–13)
Note that the angle f does not enter into (5–14), even though it describes the complete
pattern. This is because of the orientation of the vertical array with respect to the coor-
dinate system.)
192 Arrays
Equations (5–13) and (5–14) express separately the patterns of horizontal and vertical
linear arrays into which the plane array may be decomposed. Thus, the question is, what
is the pattern of the complete array? Note that the array can be viewed as consisting of
vertical line arrays that are equally spaced along the y-axis (the horizontal), where each
vertical array is a triplet, containing three isotropic elements. Thus, the array may be
viewed as a horizontal linear array of the triplets, and (5–14) expresses the pattern of
each triplet. Therefore, the planar array pattern may be attained by application of (5–7),
the principle of pattern multiplication. This principle can be applied in the present case
by considering the planar array to be a horizontal uniform linear array, whose individual
elements are vertical arrays (the triplets). Therefore the complete planar array pattern is
attained by simply multiplying Eh and Ev of (5–13) and (5–14); namely
Equation (5–13) gives the pattern versus q and f of a planar broadside array of equal
amplitude, equal phase, and uniformly spaced isotropic point sources. The subscripts h
and v are used with N and d to denote the number of elements and their spacing in both
the horizontal and vertical dimensions. Here the factor kn of (5–7) is equal to 1/NhNv. It
is to be noted that use of the principle of multiplication for (5–15) is possible because
of the uniformity of the array, that is, the array description in the x direction is indepen-
dent of position along the y-axis; and conversely, the array description in the y direction
is independent of position along the x-axis. This independence between the x and y direc-
tions causes the planar array mathematics to be separable in x and y. However, in general,
complete patterns of planar arrays cannot be attained from the product of principal plane
patterns (see sec. 5.7.3).
There is a further step required, because the usual planar array employs half-wave
dipoles or elements of another type, rather than isotropic elements. Then, the complete
pattern is obtained by multiplying (5–15) by the relative pattern of the element. If dipole
elements are used and are placed with their axes vertical (i.e., lying along the z-axis),
their patterns are given by equation (4–11) (only the part in the square brackets is needed
for the relative pattern). If instead, dipole elements are horizontal (parallel to the y-axis),
the relative pattern (corresponding to (4–11) for a vertical half-wave dipole) is
π
cos sin θ sin φ
2
E (θ , φ ) = (5–16)
1 − sin 2 θ sin 2 φ
(This additional multiplication will not be performed explicitly here because the principle
has already been adequately illustrated.)
Planar and Volume Arrays with Uniform Aperture Distribution 193
Just as a planar array was formed by stacking linear arrays in parallel rows, a volume
array can be formed by combining planar arrays. As noted, a single broadside planar
array has a bidirectional pattern. If two such arrays are placed parallel to each other a
distance d apart and phased so that the corresponding array of the two isotropic elements
would be a unidirectional endfire array in accordance with (5–12), the resulting complete
pattern will be unidirectional. Ordinarily, two planes of elements will suffice. The
complete pattern is then obtained by further application of the principle of pattern
multiplication.
Still another method of obtaining a unidirectional beam is to place a plane reflecting
surface parallel to a single planar array, with a suitable spacing between the array and
the reflector. The resulting pattern may be calculated by considering the actual array in
combination with its image in the reflector, according to the image principle described
in sec. 1.2. The spacing between the planar array and its image will be twice the spacing
of the array and the reflector, and the phases will be opposite, that is, 180 degrees apart.
Thus the array plus its image are in effect a volume array, two elements deep.
The reflector method has advantages over the actual volume-array method, in that
since there are actually only half as many dipoles the feed system is less complicated.
Also the reflector effectively prevents back radiation except for a small amount due to
diffraction around the edges of the reflector, provided that the reflector is either a solid
sheet of metal or a fine-mesh screen. However, a fairly coarse mesh or even a parallel-
wire grating (wires parallel to the antenna polarization) gives satisfactory results, with
much-reduced weight and wind resistance, if the spacing between reflector wires in the
direction perpendicular to the polarization is a small fraction of a wavelength. The reflect-
ing surface should extend some distance beyond the edge elements of the array—usually
a distance comparable to the spacing between elements.
The dipole-to-reflector spacing, within certain limits, is not critical. The useful range
of spacings is from about 1/16 to 1/4 wavelength. Close spacing gives somewhat greater
gain with an array of few elements, but at the expense of low dipole-feed-point imped-
ance and reduced bandwidth. With large arrays the gain does not depend appreciably on
the spacing, within the permitted range. Too large a spacing, however (greater than half
a wavelength), will “split” the pattern and result in more than one beam. A good com-
promise among the competing factors—directivity, ohmic losses, and bandwidth—is a
spacing of about 1/8 wavelength. (Ohmic losses increase with the closer spacings because
of lowered radiation resistance and higher currents.)
The two-element-deep volume array, consisting of two large-area vertical planar
arrays, may be preferable at the lower frequencies because a reflector of sufficient size
would be much more expensive and would have excessive wind resistance. Such anten-
nas, known as “double curtain” arrays, can be used in the frequency range from about 3
to 30 MHz or higher. Above 30 MHz the use of a reflector instead of a true volume array
becomes increasingly advantageous. A compromise method is the use of parasitic dipoles
as the reflecting device. Planar arrays with plane reflectors, sometimes called “mattress”
or “bedspring” antennas, are much used in the VHF and UHF bands. As radar antennas
194 Arrays
R
q
R
y z
f
f y
x
(a) (b)
FIGURE 5–14.
Spherical coordinates with array elements in x-y plane. (5–14a) shows point P at distance R
from origin; (5–14b) shows point P and distance R, and distance r1 from a position xi,yi on the
x-y plane.
they are often mounted on a rotating platform so that the beam “scans” the horizon, typi-
cally at rotation speeds of 1 to 30 rpm, depending on the antenna size. Such reflecting
screens are commonly as large as 30 feet or more with hundreds of elements, and much
larger ones can be used on occasion for special purposes.
N −1
(5–17)
2π
= ∑ Ai exp − j δ i
i=0
λ
Linear Array Pattern Calculations 195
where
Ai = ai exp( jai), where ai is the amplitude and ai is the phase of the radiated E-field
of the ith element
k = 2p/l
d i = ri − R
ri = distance from ith element to an observation point in the far zone
R = distance from the reference point on the array to the observation point in the far
zone
N = total number of isotropic elements
An array factor is a complex function of angle. Therefore, a graph of the amplitude of
an array factor F(q, f) is obtained by plotting |F(q, f)|. To obtain a normalized amplitude
pattern of an array factor, one must of course divide |F(q, f)| by the peak amplitude of
|F(q, f)|.
Note that di is the distance between the ith element and an observation point that
exceeds the distance between the origin (x = y = z = 0), the array reference point, and
the observation point, said point being in the far zone. Thus, when including the effects
of path length differences, the relative far-zone phase contribution of the ith element yi
follows:
2π
ψ i = α i − δ i (5–18)
λ
where
and
Recall that F(q, f), as defined, is the pattern calculated under the assumption that each
array element radiates isotropically. Then, the far-field pattern E(q, f) of an array having
identical elements can be calculated by using the principle of multiplication. In reality,
however, element patterns of physically identical elements are never the same when the
196 Arrays
elements are actually placed in an array. However, in principle, if each element pattern
is described by the function f(q, f), the far-field pattern E(q, f) is
E (θ , φ ) = f (θ , φ ) ⋅ F (θ , φ ) (5–21)
It is to be noted the F(q, f) and E(q, f) are generally complex, and therefore their ampli-
tude patterns are created by plotting magnitudes only, that is, |F(q, f)| and |E(q, f)|.
In summary, (5–17) and (5–19) are equal and are general expressions for the array factor
if each array element is an isotrope; and (5–21) is the far-zone E-field array pattern if the
pattern of each element can be expressed as f(q, f). A note of caution: effects of mutual
coupling between elements may cause the “in-place” element patterns (called scan element
pattern, SEP) to differ from their “free space” patterns. Because of this, even for an array
of physically identical elements, the actual array pattern may not equal F(q, f) times the
“free-space” element pattern f(q, f), as indicated by (5–21). Conventionally, however,
(5–21) is used as a “best” estimate until more detailed information is available.
N −1
2π
F ( 90°, φ ) = ∑ ai exp j α i + λ
yi sin φ
(5–22)
i=0
In (5–22) the element amplitudes, phases, and separation distances are arbitrary. Equation
(5–9), previously introduced in sec. 5.2 is also for arrays along the y-axis, but it is appli-
cable only to arrays for which the element amplitudes, phase differences, and physical
separations are all equal.
Equation (5–22) can be further simplified if, between adjacent elements, the separation
distance d and the phase difference a are constant, as is common in practice. Furthermore,
frequently the input current amplitude and phase of an array are symmetrical about the
array center. Then, F(90°,f) with its phase reference at the array center can be expressed
as follows:
( N −1) 2 2π
F ( 90°, φ ) = ∑ an exp jn α + d sin φ (5–23)
n =−( N −1) 2 λ
where
an = amplitude of the element located on the y-axis at yn = n · d
a = phase difference between adjacent elements
Linear Array Pattern Calculations 197
E ( u ) = sin u u (5–24)
where
u = (π L λ ) sin φ
Linear Array Pattern Calculations 199
and f denotes the angle measured from normal to the aperture of width L. In comparison,
from (5–9), for a broadside linear array of isotropes having equal amplitude and phase
with spacing d, the normalized amplitude pattern is
π d sin φ ⎞ ⎤
sin ⎡⎢ N ⎛
⎣ ⎝ λ ⎠ ⎥⎦ (5–25)
π d sin φ ⎞ ⎤
N sin ⎡⎢⎛
⎣ ⎝ λ ⎠ ⎥⎦
where N is number of elements. With the approximation n times d used as the array
aperture width L, the numerator of (5–25) becomes sin u as in (5–24). Of course, the
array width is actually N − 1, but the difference between N and N − 1 diminishes for
large values of N. Therefore, for a continuous aperture and an array with a large number
of elements, each of the same width, the zeros of E(u) and E(f) occur at practically the
same angle f.
There are also other close similarities between the broadside patterns for closely
spaced arrays and continuous aperture antennas. For example, because sin x ≈ x, if x is
small and in radians, the numerator of (5–25) becomes approximately u at near-broadside
where f is small. Figure 5–17 includes patterns of a uniformly excited broadside array
with spacing d of l /2 and of a continuous aperture, each of the same length. The patterns
are virtually equal at angles near the direction of the main beam. As also apparent from
Fig. 5–17, the peaks in the minor lobes differ increasingly as the angle f becomes closer
to ±90°. For aperture widths of about 5l and larger, the side-lobe peaks at wide angles
(f near ±90°) may differ by several decibels. Even so, at these wide angles the field
strength, and thus the energy, in either pattern is small compared to the energy in the
major lobes. Therefore, since the two major lobe patterns are almost identical, the direc-
tivity of the two antennas of Fig. 5–17 are virtually the same.
For most practical purposes, the approximate equality of broadside antenna
patterns, whether from closely spaced (l /2 or less) elements or continuous apertures, is
retained provided the array amplitudes and phases are matched point-to-point with the
continuous aperture distributions. This
near-equality of patterns exists provided
Relative amplitude in dB
the array axis. On the other hand, there will be only one major lobe (unidirectional
pattern) if the array element spacing d is less than l /2. An endfire array can provide
higher directivity for a given number of elements than a broadside array. However,
endfire arrays require larger dimensions in the along-beam direction. For an ordinary
(conventional) endfire array, the phase difference between elements is opposite to the
phase delay due to element spacing, thereby causing the radiation of all elements to add
constructively at the center of the beam. Thus a = −2pd/l for an ordinary endfire array,
where a is the phase difference between elements.
The directivity is said to be extraordinary if it satisfies the Hansen and Woodyard
(H-W) condition, and then its directivity is larger than that of the ordinary endfire array.
For an H-W array, there are N equal amplitude isotropic elements and the phase a
between elements is
2π d π
α = + (5–26)
λ N
The patterns of ordinary and extraordinary endfire arrays are now compared.
Assume the elements are equally spaced isotropic elements of equal amplitude and
located on the y-axis of Fig. 5–14a. Now consider the pattern of an endfire linear array
in the x-y plane, for which q = 90°. Then, with N = 10, each amplitude an = 1, and the
element spacing d of l /4, (5–23) becomes
4.5
π
F ( 90°, φ ) = ∑ exp jn α + sin φ (5–27)
n =−4.5 2
Now assume that the beams of the ordinary endfire array and of the one satisfying the
Hansen and Woodyard condition are pointed in direction f = 90°. The phase a between
elements for an ordinary endfire array equals the phase delay due to propagation between
elements; thus, a = −p/2 because d = l /4. From (5–26), for the extraordinary (H-W)
condition, with d = l /4 and n = 10 ele-
Normalized array factor
1 ments, a = −0.6p.
0.8 The normalized patterns for the two
0.6 unidirectional endfire arrays having 10
0.4 elements and l /4 element spacing, using
0.2 (5–27), are included in Fig. 5–18. Notice
0 that the Hansen-Woodyard array has a
–90 –45 0 45 90 135 180 225 270
Angle in degrees from array axis substantially narrower beamwidth, and
FIGURE 5–18. larger side lobes. On the other hand, the
shape of the major lobe of the ordinary
Normalized patterns for ten-element endfire
arrays with d = l/4 element spacing: ordinary endfire array is flatter at its peak.
endfire (bold line) and endfire satisfying From (5–27), the values of the array
Hansen-Woodyard condition (fine line). factors along the y-axis are
Array Tapering for Side-Lobe Reduction 201
4.5
π π
F ( 90°, 90°) = ∑ exp jn − + = 10 ordinary enddfire array (5–28)
n =−4.5 2 2
and
4.5
π
F ( 90°, 90°) = ∑ exp jn −0.6π + = 6.392 extraorrdinary endire array (5–29)
n =−4.5 2
Recall that electric field strength is proportional to array factor F. Although (5–29) is
smaller than (5–28), the directivity of the Hansen-Woodyard array is larger. This is
because the Hansen-Woodyard pattern (and thus energy) is more closely confined to the
y-axis. Additionally, F(90°, 90°) is largest in (5–28), because the radiation from all ele-
ments adds in-phase but not so for (5–29). Kraus and Marhefka (2002, pp. 115) report
the directivities obtained by integrating the patterns, including the minor lobes, for the
two ten-isotrope arrays. They give the approximate directivities for the ordinary and
Hansen-Woodyard arrays to be approximately 11 and 19, respectively. For these directivi-
ties, the Hansen-Woodyard directivity is the largest by the factor 1.73.
A focused broadside array of ten isotropes has in theory a directivity of 10. Thus, from
the directivities given above and the same number of elements, lossless endfire arrays
can provide greater gain, as well as larger directivity, than a broadside array.
is approximately 20, or 13 dB. A side-lobe level −13 dB below the major lobe peak is
often unacceptably high. It is common to design high-gain narrow-beam antennas for
side-lobe levels down at least 30 dB, especially for radar use. A −30 dB level is consid-
ered good, and −40 dB excellent. A level of 50 dB below the main lobe is very difficult
to achieve, but even somewhat better side-lobe levels have been obtained. Planar arrays
are typically tapered by different amounts in the horizontal and vertical dimensions,
because side-lobe requirements versus principal plane differ depending on an antenna’s
application.
N Relative amplitudes
3 1 2 1
4 1 3 3 1
5 1 4
1 6 4
6 1 5
10 10 5 1
7 1 6 15 20 15 6 1
8 1 7 21 35 35 21 7 1
Array Tapering for Side-Lobe Reduction 203
The current distribution is symmetrical about the center of the array; hence the ele-
ments can be grouped in pairs as shown. The current values are normalized to the value
unity at the center of the antenna. As the table shows, for this −30 dB side-lobe level,
the end-element currents are only 26 percent of the center-element current. For half-
wavelength element spacing, the 3-dB beam width of this array is 16.4 degrees and
directivity is 6.7 (8.3 dB).
Unfortunately, antennas having a large number of elements and designed for low side
lobes using D-C aperture distributions are impracticable. For example, since the side-lobe
levels are constant (independent of direction), directivity and thus gain decrease with
204 Arrays
1 1 1 1 1 1 4 6 4 1 1 1.9 1
(a) (b) 1.6 1.6
(c)
15°
Edge
1 0 0 0 1
(d)
FIGURE 5–19.
Normalized far-field patterns of five-element broadside arrays. Only the upper half of each
pattern is shown. All elements have same phase, and the relative amplitudes are included below
each figure. From John D. Kraus, Antennas, 2nd ed., ©1988, p. 160, with permission of The
McGraw-Hill Companies.
Array Tapering for Side-Lobe Reduction 205
increase in array size. This is because, as the principal beamwidth is reduced, a larger
fraction of the radiated power is within the side lobes. Also, for increasingly larger arrays,
the required array element excitation becomes nonmonotonic, requiring high peak cur-
rents at the end elements that causes large I 2R losses. This feature reduces available gain.
Another problem with low side-lobe D-C designs is that the element currents must be
controlled to an impracticably high accuracy, which contributes to them being unrealiz-
able. Therefore, aside from the application to arrays with a small number of elements,
D-C patterns are important for identifying the narrowest beamwidth theoretically avail-
able for a given peak side-lobe level. A simplified method is described in the next section
for calculating theoretical, yet unattainable, D-C patterns for arrays having a large
number of elements.
E ( u ) = cos u 2 − (π A)
2
(5–30)
where u = (pL/l)sin q
A = (1/p)arccos h(R)
R = ratio of electric field strength of the peak side lobes to main beam, and
where arcos h is the inverse hyperbolic cosine, l is wavelength, and q is the observation
angle measured from the normal to the aperture.
The 3 dB beamwidth for a D-C pattern expressed in radians is
BWHPCheb = (λ L ) β 0
where (5–31)
(
0.5
β0 = (2 π ) ⎡⎣(arccos h( R ))2 − arccos h(0.70 R )2 ⎤⎦
0.5
β0 = ( 2 π ) ⎡⎣( 5.298)2 − ( 4.952 )2 ⎤⎦ = 1.20
Therefore, the “optimum” 3 dB beamwidth for constant 40 dB peak side lobes, although
unrealizable, expressed in radians, is
The broadside beamwidths for continuous line apertures are discussed in sec. 6.10.
The narrowest broadside continuous line-source beamwidth is one that has an aperture
distribution of constant amplitude and equal phase (uniform distribution). The 3 dB
beamwidth for this basic aperture, expressed in degrees is
Therefore (5–31) and (5–34) can be used to provide a comparison between the most
efficient distribution (which gives −13 dB peak side lobes) and the D-C pattern (which
gives the narrowest major lobe for a specified constant side-lobe level). However, as
already stated, D-C patterns are not always practical or realizable.
Figure 5–20 is included to compare the patterns for a uniform aperture distribution
and a D-C distribution with 26 dB side lobes. Each aperture is 2 m wide and the fre-
quency is 3 GHz. Although the beamwidth for the uniform aperture is slightly narrower
than provided by the D-C distribution, the much larger side lobes created by the uniform
distribution are apparent.
0
–10
5.6.5. The Taylor Distribution
Decibels
–20
The Taylor aperture distribution is often
–30 used where narrow broadside beams and
low side lobes are needed. Taylor (1955)
–40
–45 –30 –15 0 15 30 45 analyzed the deficiencies of Dolph-
Angle from Broadside in Degrees Chebyshev arrays and formulated a com-
bined radiation pattern and aperture
Constant Amplitude and Phase
Dolph-Chebychev Distribution distribution that provides narrow beam-
widths, and yet has good efficiency for
FIGURE 5–20.
large arrays. His analysis began with the
Two patterns, one for a uniform amplitude line-source D-C patterns given by Fig.
distribution and the other for a Dolph-
Chebychev distribution with 26 dB side lobes. 5–20 above. In so doing Taylor developed
Each aperture is 2m wide and the frequency is a method for avoiding the D-C directivity
3 GHz. (and gain) loss problem caused by no
Array Tapering for Side-Lobe Reduction 207
decay in the side lobes. This was accomplished by approximating, arbitrarily closely, the
D-C pattern with a physically realizing pattern. He retained an approximate constant
side-lobe pattern close to the main beam, but let the wide-angle side lobes decay in
amplitude like the side lobes from a uniform continuous line source.
Taylor’s method is basically designed to apply to continuous aperture distributions
rather than to an array of discrete elements, but it can be applied to linear arrays of a
large enough number of elements (about twenty or more). This is the case where the
Dolph-Chebyshev distribution results in poor directivity. Taylor’s method gives improved
directivity for a given peak minor-lobe level, because it allows the far-out side lobes to
diminish in amplitude.
A parameter n, called nbar, defines the number of nondecaying side lobes on each
side of the main lobe. Thus as n becomes large, the Taylor pattern (and aperture distribu-
tion) approaches the Dolph-Chebyshev pattern (and distribution). Then, ordinarily, one
uses the largest n that still provides a monotonically declining amplitude distribution out
to each end of the aperture. In doing so, the beamwidths obtained are almost as narrow
as with the D-C pattern—and the resulting aperture efficiency is excellent.
The Taylor aperture distribution provides a beamwidth wider than that of a D-C pattern
by an approximate factor s. From Mailloux (1994 p. 129), the 3 dB beamwidth for a
Taylor distributed aperture can be expressed as
BWHPTaylor = σ ( BWHPCheb ) = σ (λ L ) β0
where (5–35)
−0.5
σ = n ⎡⎣ A2 + ( n − 1 2 )2 ⎤⎦
with
n = number of equal-amplitude side lobes on one side of the main beam
L = total length of aperture
A = (1/p)arccos h(R)
R = design side lobe voltage ratio
Now to compare the beamwidths of the unrealizable D-C pattern of (5–30) and a
realizable Taylor pattern, consider an array with n = 4 and 40 dB peak side lobes. Then,
from the definitions given above
and
−0.5
σ = 4 ⎡⎣(1.686 )2 + ( 4 − 0.5)2 ⎤⎦ = 1.03 (5–37)
Thus for n = 4 and 40 dB peak side lobes, the 3-dB width of the main beam of the
Taylor pattern is only 3 percent wider than a D-C pattern having 40 dB side lobes.
208 Arrays
Recall from (5–32) that the beamwidth for a D-C pattern with 40 dB side lobes is
1.20 (l /L).
As already mentioned, in addition to the Taylor 3-dB beamwidth being broader than
that of the Chebyshev pattern, in principle the n = 4 Taylor pattern contains four equal-
amplitude side lobes on either side of the major lobe. Also, in principle, the remaining
side lobes have widths and amplitudes like those from a continuous uniform aperture
distribution.
Figure 5–20 is useful for roughly illustrating the formation of a Taylor pattern by a
merger of two patterns: one from a uniform aperture distribution and the other from the
D-C distribution with 26 dB side lobes. Each aperture is 2m wide and operates at 3 GHz.
Ideally, the Taylor side lobes predominate in close to the major lobe and out to angles
where the side lobes of the uniform amplitude pattern are the smallest. Then, the uniform
amplitude pattern dominates a Taylor pattern. From Fig. 5–20, a reasonable guess is that
a resulting Taylor pattern might have four or five side lobes (n = 4 or 5) at the −26 dB
level on either side of the major lobe. However, care is required in a Taylor design when
selecting the nbar and side-lobe levels, because calculated Taylor patterns are seldom as
precise as theory predicts. In addition, the choice of n is not arbitrary, because increasing
it retains more of the side lobes at the design side-lobe level and this makes the calculated
Taylor closer to a D-C pattern. In other words, increasing n eventually leads to aperture
illuminations that do not decrease monotonically down to the aperture edges (Mailloux
1994, p. 133). Thus, the act of increasing n too much can destroy the usefulness of the
aperture distribution.
In summary, the Taylor distribution is widely used, because it is a realistic way of
providing the narrowest beamwidth from a broadside array, given a specific peak side-
lobe level and an aperture width. Neither the theory nor design for Taylor distributions
is simple. However, the necessary equations for calculating aperture distributions and
patterns are available (see, e.g., Mailloux 1994, pp. 128–29; Balanis 2005, pp. 408–10).
Additionally, tables are available that give required array element amplitudes, for a
variety of n values, for Taylor arrays having side-lobe levels for 20 to 40 dB, inclusive,
in 5 dB steps (Reference Data for Radio Engineers 1979, pp. 27–29 through 27–31).
As discussed previously, the patterns of arrays and continuous apertures are similar if
(1) the array element spacing is l/2 or less and (2) the array amplitudes and phases are
matched point-to-point with the continuous aperture distributions. Consequently, from
previously developed knowledge about continuous apertures there are simple amplitude
distributions that are known to provide low side-lobe levels. For example, a cosine-
shaped amplitude distribution with its peak at the aperture center will provide a reason-
ably efficient, low side-lobe pattern. Digital filter algorithms (Harris 1978) also provide
useful aperture amplitude distributions, the Hamming function being an example. Bodnar
(2007; pp. 55–16 through 55–19) provides graphs that show beamwidth versus side-lobe
level and aperture efficiency versus side-lobe level, for the better-known uniform aperture
distributions. These graphs are useful for estimating performance available from array
antennas.
5.7.1. Introduction
A planar array pattern can be calculated by summing the vector contributions of each
array element at every direction and at a distance where far zone approximations are
valid. Then, for an array located in the x-y plane, the array factor is
where
q, f = angles to a far-zone point from the aperture origin
ai = amplitude of element i
2π
yi = α i ( x , y ) − δ i ( x, y ), the phase contribution of element i at the observation point
λ
ai = phase of element i
di = −sin q(xi cos f + yi sin f), which is the difference in distances to an observation
point in the far zone (see Appendix H), from (a) a point xi, yi on the aperture
and from (b) the aperture origin
The reader will recognize that (5–38) is a summation that includes each ai, ai, and di
over all x and y values. Consequently, the calculation of F(q, f) is, in general, complicated
because di (contained in yi) is a function of both q and f. Furthermore, for calculations
210 Arrays
and
Then, by letting Amn = amn exp jamn, where amn is an element amplitude and amn is its
phase, a general expression for the planar array factor is as follows:
with
2π
sx = d x sin θ cos φ (5–42)
λ
and
2π
sy = d y sin θ sin φ (5–43)
λ
where
dx and dy are the element spacings along the x and y axes
m is an integer such that mdx denotes a coordinate of an array element on x-axis
n is an integer such that ndy denotes a coordinate of an array element on y-axis
Planar Arrays: Patterns, Directivity, and Gain 211
Thus, for separable array distributions, we see from (5–44) that a planar array factor is
the product of two terms. On careful examination, it may be seen that the first and second
terms are the array factors for line arrays along the x and y axes, respectively. In other
words,
F (θ , φ ) = Fx (θ , φ ) × Fy(θ , φ ) (5–45)
Thus, Fx(q, f) and Fy (q, f) are array factors for line sources along the x and y axes, as
follow
2π
Fx (θ , φ ) = ∑ am exp jm (α x + sx ) = ∑ am exp ⎡⎢ jm ⎛⎜ α x + d x sin θ cos φ ⎞⎟ ⎥⎤ (5–46)
m m ⎣ ⎝ λ ⎠⎦
2π
Fy(θ , φ ) = ∑ an exp jn (α y + s y ) = ∑ an exp ⎡⎢ jn ⎛⎜ α y + d y sin θ sin φ ⎞⎟ ⎥⎤ (5–47)
n n ⎣ ⎝ λ ⎠⎦
The mathematics of complete patterns are complex, partly because sx and sy are both
functions of q and f. Thus, the complexity of array factors exists, even for arrays having
element excitations that are separable in the x and y dimensions.
For principal plane patterns, the equations for patterns with separable distributions are
considerably less complex, especially if the phase excitations of all elements are equal.
By setting f = 0° and 90° in (5–46) and (5–47), and with ax = ay = 0, (5–45) for the
principal plane patterns becomes
212 Arrays
2π
F (θ , φ = 0°) = ∑ am exp ⎡⎢ jm ⎛⎜ d x sin θ ⎞⎟ ⎤⎥ × ∑ an ( x-z plane ) (5–48)
m ⎣ ⎝ λ ⎠⎦ n
2π
F (θ , φ = 90°) = ∑ an exp ⎡⎢ jn ⎛⎜ d y sin θ ⎞⎟ ⎤⎥ × ∑ am ( y-z plane ) (5–49)
n ⎣ ⎝ λ ⎠⎦ m
2π
(1) ∑ am exp ⎡⎢⎣ jm ⎛⎜⎝ d x sin θ ⎞⎟ ⎤⎥ is a principal plane pattern for a linear array
m λ ⎠⎦
located along the x dimension
2π
(2) ∑ an exp ⎡⎢ jn ⎛⎜ d y sin θ ⎞⎟ ⎤⎥ is a principal plane pattern for a linear array
n ⎣ ⎝ λ ⎠⎦
located along the y dimension
(3) ∑ am and ∑ an are simply constant multiplicative factors.
m n
Items (1) through (3) above relate to the principal plane patterns for an array with
separable amplitude distribution and ones for which each element has the same phase.
From these items it may be seen that the equations for the principal plane patterns are
much simpler than those of the three-dimensionl patterns. In fact, each principal plane
pattern is the same as that of a linear array.
where yx = as + sx.
Equation (5–50) is a geometric progression. Its sum is complex (amplitude and phase)
and can be expressed in closed form. This sum, for a total of M identical elements along
the x axis, with the phase reference taken at the center of the line array, and with the
amplitude normalized for a maximum of unity, is derived by Kraus (1988, pp. 138–40),
and is as follows
sin ( Mψ x 2 )
Fx (θ , φ ) = (5–51)
M sin (ψ x 2 )
Planar Arrays: Patterns, Directivity, and Gain 213
Then, because Fx and Fy are both geometric progressions of the same general form,
and with N being the total number of elements along the y axis,
sin ( Nψ y 2 )
Fy(θ , φ ) = (5–52)
N sin (ψ y 2 )
Then, according to (5–45), and with all element amplitudes am and an being unity,
F(q, f) for a planar array with separable phase distributions can be expressed as follows:
sin ( Mψ x 2 ) sin ( Nψ y 2 )
F (θ , φ ) = Fx × Fy = × (5–53)
M sin (ψ x 2 ) N sin (ψ y 2 )
where
2π
ψ x = α x + s x = α x + ⎛⎜ d x sin α cos φ ⎞⎟ (5–54)
⎝ λ ⎠
and
2π
ψ y = α y + sy = α y + ⎛⎜ d y sin θ sin φ ⎞⎟ (5–55)
⎝ λ ⎠
The simplest equation occurs for the case where the phases as well as the amplitudes
are equal, which results if each am and an in equations (5–54) and (5–55) is zero. This
simplest condition is the one addressed by (5–13), sec. 5.4.2, but it applies to the case
of equal element amplitudes and phases when the array elements are located in the y-z
plane. Instead, the various equations in secs. 5.7.1 through 5.7.4 are applicable when the
array elements are in the x-y plane.
As seen above, the equations for far-field planar array patterns are not mathematically
simple, even for simpler ones like (5–53) above. The creation of patterns is discussed
further in chapter 6, where example patterns for continuous aperture antennas are included.
Papers on calculations of patterns versus q and f that use Matlab software include
Bregains, Ares, and Moreno (2004) and Sevgi and Uluişik (2005). SM 5.0 included in
the accompanying website gives equations for, and examples of, 3-D plots of patterns
versus q and f using Mathcad software.
Equation (5–38) is a general equation for the E-field radiation pattern for a planar
array, where the element amplitudes, phases, and spacing are completely arbitrary. This
equation is written for the radiating far-field; but, as indicated in the text, it can be readily
changed for applicability to the radiating near-field patterns.
Equation (5–41) is a general equation for patterns versus q and f, but it requires there
be equal spacing between elements along the x dimension and along the y dimension.
However, the spacings along the x direction need not be equal to the spacing along the
y direction. The amplitudes and phases of the elements may be arbitrary. Although the
equation is complex, it has wide applicability because, for practical reasons, multielement
arrays are usually constructed with equal spacing between elements.
In addition to having equal spacing between elements, multielement arrays are often
constructed so that the element amplitudes and phases are mathematically separable. This
means that the variations in amplitude and phase along the x dimension are independent
of the variations in amplitude and phase along the y dimension. Equation (5–44) describes
the pattern when the element excitations are separable, but otherwise the amplitudes and
phases of the elements are arbitrary. The requirement of (5–41) for equal spacings along
the x and y dimensions is also applicable to (5–44).
From (5–44) it can be seen that, for separable aperture distributions, the equation for
three-dimensional patterns reduces to the product of two line sources. Equations (5–46)
and (5–47) are the array factors for these line sources, one that has amplitude and phase
variations along the x dimension, and the other has amplitude and phase variations along
the y dimension.
For principal plane patterns, the equations for apertures with separable distributions
are considerably less complex, especially if the phase excitations of all elements are
equal. As may be seen, (5–48) and (5–49) for these principal plane patterns for planar
arrays are of the same form as principal plane patterns for linear arrays.
Equation (5–53) describes the pattern if an array has elements of equal amplitude, and
if the phases along the x- and y-dimensions are separable. The equation is simplified even
further if both the element amplitudes and the element phases are equal. Previously, in
sec. 5.4.2, equation (5–13) was introduced for the simplest case when the amplitudes and
phases of all elements are equal. For (5–13), the array elements are located in the y-z
plane, instead of in the x-y plane for which the equations of secs. 5.7.1 through 5.7.4
apply.
As previously discussed, the array factor F(q, f) is the array pattern calculated under
the assumption that each array element radiates isotropically. In reality, element patterns
are never isotopic, but they may be described by a function f(q, f). If all element patterns
were identical in an actual array antenna, the far-field pattern E(q, f), because of the
principle of multiplication (sec. 5.1.6), would be
E (θ , φ ) = f (θ , φ ) ⋅ F (θ , φ ) (5–56)
Mutual coupling between elements may limit the validity of (5–56) (see secs. 5.1.4
through 5.1.7, and 5.9). Often, however, when (5–56) is normalized to have a maximum
of unity, it provides a practical estimate for actual array patterns. F(q, f) and E(q, f) are
Planar Arrays: Patterns, Directivity, and Gain 215
generally complex, and therefore their E-field amplitude patterns are created by plotting
their absolute magnitudes.
51λ
BW = degrees (5–57)
Nd
(which for d = l /2 is 102/N as given above). The variation of the directivity with element
spacing for a given number of elements is not expressible by a simple formula. Curves
showing this variation have been published by Tai (1964) for both uniform and nonuni-
form linear arrays. As mentioned for broadside arrays in sec. 5.2.2, the gain for a given
number of elements increases as the spacing is increased up to a maximum which occurs
at slightly less than one wavelength spacing (about 0.95 l) for a large array (many ele-
ments). For N = 3 the maximum directivity occurs at about d = 0.8 l.
The beamwidth formula, equation (5–57), can also be applied to the horizontal and
vertical beamwidths of a vertical planar array by considering it to be a linear array in
each direction, of Nh elements with spacing dh horizontally, and Nv elements with spacing
dv vertically. Although this formula is for an array without a reflecting screen it also
applies with a reflector if N is large.
Broadside arrays are often made unidirectional by placing a reflecting screen behind
the array, or by using two curtains of arrays with endfire phasing between the curtains.
216 Arrays
When such arrays are large (many wavelengths) the beamwidths and directivity are given
by formulas of the following type:
λ
BW = k1
Nd
(5–58)
4π A
D = k2 2 (5–59)
λ
where k1 and k2 are of the order of one or slightly less, if the beamwidth is expressed in
radians and if all array elements are similar and radiating in phase with equal intensity.
If the beamwidth is expressed in degrees, k1 is in the vicinity of 60—somewhat less for
a uniform array and somewhat more for tapered arrays. For a tapered array the constant
k2 in (5–59) is less than one. Commonly, k2 is known as the aperture efficiency.
Equation (5–59) with k2 = 1 represents the maximum directivity that can be obtained
with an ordinary unidirectional planar array of area A. As mentioned, k2 = 1 applies to
a broadside uniformly excited planar unidirectional array (all elements in phase and
radiating with equal intensity). Similarly, the directivity D = N for a broadside half-wave-
spaced uniform linear array of N elements (no reflector) is ordinarily the maximum
obtainable value of directivity for an array of that length. Thus, for array length L and
element spacing l/2, the maximum directivity is D = (2L/l) + 1.
However, it is possible, in an array of closely spaced elements, to achieve directivities
greater than the above mentioned uniform-array maximum values by properly phasing
the elements. An array in which this is accomplished is a supergain or superdirective
array. This possibility, however, is not as attractive as it seems at first. The superdirective
array is characterized by relatively large phase changes within a relatively small distance.
The mutual coupling between elements cause much larger currents fields than in a normal
array. The currents increase enormously if any substantial amount of superdirectivity is
attempted, to the extent that ohmic losses quickly overcome any increase in directivity,
so that gain is not increased.
Except for very small arrays, super directive illuminations (Hansen 1964, pp. 82–91)
have proven impractical, because of the high I 2R losses resulting from these very large
currents. Also, the bandwidth of arrays having superdirective illumination becomes
extremely small as the superdirectivity is increased. Therefore, the supergain effect is
not practical except in very limited degree in special situations. Moderate superdirectivity
is achieved, however, in endfire arrays phased according to the Hansen-Woodyard condi-
tion, but even moderate superdirectivity is impractical in broadside arrays.
Even so, as now discussed, the endfire array is being used successfully to obtain a
higher gain for a given frontal area than available from a broadside planar aperture. An
endfire array may provide higher directivity and gain for a given number of elements
than a broadside array. However, endfire arrays require larger dimensions in the along-
beam direction. The UHF array on the United States Navy E-2C aircraft uses the endfire
principle to advantage (Long, 2004). It has a horizontal array of horizontally polarized
Planar Arrays: Patterns, Directivity, and Gain 217
Yagi antennas, which permits a small vertical dimension. The horizontal array provides
a relatively narrow azimuth beam. Of especial importance is that its desired broad eleva-
tion beam is obtained with a much smaller aperture height than would be required of a
broadside planar antenna. The result is an antenna with small vertical dimension having
low aerodynamic drag. Additionally, the antenna gain is about four times the gain that
is available from a continuous aperture of the same frontal area.
The physical aperture Ap is a measure of the physical size of an antenna area over
which radiation is transmitted or received. For a uniformly illuminated aperture, that is,
one for which the amplitude and phase are both constant over the physical area and one
that is lossless, the gain G is as follows:
G = ( 4π Ap ) λ 2 (5–60)
Equation (5–60) was shown by Silver (1949, p. 183) to be true by integrating the
total radiation pattern for a uniformly illuminated aperture and calculating D. The mea-
sured gain of a real antenna is of course less, depending on losses that can occur because
of discrepancies in the antenna pattern, impedance mismatches, and dissipative (I2R)
losses.
In sec. 3.3.4, chapter 3, directivity is defined as D = 4p/W, where W is the imaginary
solid angle that contains all radiated power if it were distributed uniformly and totally
within said solid angle. Thus, directivity and consequently gain are in approximate
inverse proportion to the products of 3 dB principal-plane beamwidths. Theoretically, the
half-power beamwidth qHP for a uniformly illuminated line source of length L
θ HP = 50.8λ L (5–61)
when qHP is expressed in degrees, and l and L are in the same units. Equation (5–59)
can be obtained from the table in sec. 6.10–2, chapter 6.
Now consider a rectangular aperture. Its gain can be determined with (5–61) by
expressing physical aperture Ap in terms of principal plane beamwidths, qHP and fHP, by
using equation (5–60). Then, the approximate gain G for a lossless, uniformly illuminated
aperture is
The approximate, “typical” gain of (5–63) is nearly 1.0 dB less than gain of (5–62)
determined by using theoretical beamwidths and for a lossless, uniformly illuminated
rectangular aperture. It is to be underscored that use of (5–63) is recommended only if
the antenna efficiency (“k” in G = kD) is known to be approximately 1. The point being
that directivity and thus gain are reduced by spurious radiation not readily detected and
thus not ordinarily accounted for when determining directivity. For example, if only the
usual principal-plane measurements are made, there may be unsuspected out-of-plane
spurious radiation.
Equation (5–63) is applicable to arrays as well as continuous aperture antennas (chapter
6). Stutzman (1998, pp. 7–11) cautions the reader regarding the use of the following
equation for directivity of arrays that appears often in antenna literature, namely:
D = De Di, (5–64)
where
De = directivity of each element in the array
Di = directivity of the array with isotropic elements
This formula is for large arrays without grating lobes. Even when the transmission effi-
ciency (“k” in G = kD) is large, Stutzman recommends the use of (5–63) instead of (5–64).
This he notes is because the accuracy of (5–64) can vary greatly with the details of array
geometry. For example, small changes in array element position can increase spurious
radiation above what is indicated by a theoretical calculation for directivity (even though
there may be no perceptible change in principal-plane beamwidths), thereby contributing
to a loss in gain.
Feed line
FIGURE 5–22.
Resonant V antenna and typical pattern.
l l
Feed line
Terminating
resistor
FIGURE 5–23.
Rhombic antenna and typical pattern.
lg /2
Beam direction (z-axis)
Slots
C
Input end z
Polarization Resistive
y (x-axis) termination
x
FIGURE 5–24.
Broadside array of longitudinal shunt slots in a waveguide.
lightweight and small. For these reasons, patch arrays are widely considered for airborne
and spacecraft applications. In addition, design and construction techniques exist that
exhibit low mutual coupling between array elements, thereby suppressing the likelihood
of blind angles, that is, blindness (Yang, Mosallaei, and Rahmat-Samii 2007).
In fact, there are conditions, sometimes observed, where a matched condition exists at
the broadside beam direction, but at some other angle most of the power is reflected and
thus a scan “blindness” results. In spite of the difficulty of dealing with mutual coupling,
especially with electronically steered arrays (chapter 8), well-functioning fixed-beam as
well as electronically steered arrays have been developed and operating for a number of
decades (Hansen 2007; Oliner and Knittel 1972).
References
Balanis, C. A., Antenna Theory, Wiley-Interscience, 2005.
Bodnar, D. G., “Materials and Design Data,” ch. 55 in Antenna Engineering Handbook,
4th edi, J. L. Volakis (ed.), McGraw-Hill, 2007.
Bregains, J. C., F. Ares, and E. Moreno, “Visualizing the 3D Polar Power Patterns and
Excitations of Planar Arrays with MATLAB,” IEEE Antennas and Propagation
Magazine, April 2004, pp. 108–112.
Brown, G. H., “Directional Antennas,” Proceedings IRE, vol. 25, January 1937,
pp. 78–145.
Dolph, C. L., “A Current Distribution for Broadside Arrays Which Optimizes the
Relationship between Beamwidth and Side-Lobe Level,” Proceedings IRE, vol. 34,
June 1946, 335–48.
Hansen, R. C., Microwave Scanning Antennas, vol. 1, Academic Press, 1964.
Hansen, R. C., “Phased Arrays,” ch. 20 in Antenna Engineering Handbook, 4th ed.,
J. L. Volakis (ed.), McGraw-Hill, 2007.
Hansen, W. W., and J. R. Woodyard, “A New Principle in Directional Antenna Design,”
Proceedings IRE, vol. 26, March 1938, 333–35.
Harris, F. J., “On the Use of Windows for Harmonic Analysis with the Discrete Fourier
Transform,” Proceedings of the IEEE, January 1978, pp. 51–83.
IEEE Standard Dictionary of Electrical and Electronics Terms, 5th ed., Institute of
Electrical and Electronics Engineers, 1993.
Kraus, J. D., Antennas, 2nd ed., McGraw-Hill, 1988.
Kraus, J. D., and R. J. Marhefka, Antennas, 3rd ed., McGraw-Hill, 2002.
Long, M. W., Airborne Early Warning System Concepts, SciTech Publishers, 2004,
pp. 235–36 and 405–08.
Ma, M. T., “Arrays of Discrete Elements,” ch. 3 in Antenna Engineering Handbook,
R. C. Johnson, 3rd ed., McGraw-Hill, 1993, p. 3–23.
Mailloux, R. J., Phased Array Antenna Handbook, Artech House, 1994.
Oliner, A. A., and G. H. Knittel, Phased Array Antennas, Artech House, 1972.
Reference Data for Radio Engineers, 6th ed., Howard W. Sams & Co, Inc.,
1979.
Sevgi, L., and Çagatay Uluişik, “A MATLAB-Based Visualization Package for Planar
Arrays of Isotropic Radiators,” IEEE Antennas and Propagation Magazine, vol. 47,
February 2005, pp. 156–63.
Problems and Exercises 223
of isotropic elements. (b) Substitute into this expression the appropriate values
for n, d, and a. (c) Apply the principle of pattern multiplication to obtain the
pattern versus q and f of the dipole array.
4. A uniform broadside array of twelve point-source isotropic elements has an
element spacing of 0.65 wavelength (d = 0.65l). (a) What is the approximate
beamwidth? (b) If the spacing is d = 0.5l (instead of 0.65l), what will the
beamwidth be? (c) What is the directivity for d = 0.5l? Will the directivity for
d = 0.65l be less than or greater than this value?
5. A five-element driven endfire dipole array has element spacing d = l /8. (a) What
should the phasing (a) of the elements be to obtain maximum directivity of the
beam? Express the result in both radians and degrees. (b) Using the value of a
found in part (a), calculate E(f) from equation (5–9) for f = 90° and f = 270°
(the two “endfire” directions). What is the front-to-back ratio of this antenna,
expressed as a ratio of electric field strengths? Express it also as a decibel value.
Note: In part (b), you will find that the initial calculation indicates a “negative”
electrical intensity for f = 270°, but recall that absolute-value brackets are used
in equation (5–9) to indicate that the calculated field is regarded as always
positive.)
6. Check the box that you think precedes the correct answer to the following
multiple-choice problems:
(a) A Yagi-Uda antenna has a beam that is
ⵧ unidirectional ⵧ omnidirectional
ⵧ bidirectional ⵧ multilobed (many equal-strength
lobes)
(b) If the total number of dipole elements in a Yagi-Uda antenna is five, the
usual number of these that will be parasitic directors is
ⵧ one ⵧ three
ⵧ two ⵧ four
(c) The principal advantage of a Yagi-Uda antenna over a fully driven endfire
array with the same number of elements is
ⵧ higher gain ⵧ simpler feed system
ⵧ easier adjustment of element ⵧ better polarization characteristics
length and spacing
7. It is desired to obtain a directivity of 1,000 with a large uniform planar dipole
array backed by a reflecting screen. The wavelength is l = 0.5 meter
( f = 600 MHz). (a) What must be the area of this array? If the array is square,
what is the length of each side? (b) If the dipoles of the array have their centers
spaced a half wavelength apart in both vertical and horizontal dimensions, with
dipoles located at each corner of the square, how many dipoles does the array
contain? (Disregard the fact that the edge dimension is not exactly an integral
multiple of a half wavelength; it is very nearly so.)
Problems and Exercises 225
It has been shown in sec. 5.1 that an array antenna can be used to achieve a directional
radiation pattern in which the radiated power is concentrated in a beam. This chapter
treats an entirely different method of achieving essentially the same result, by the use of
reflectors and lenses. Antennas using these devices achieve a directional effect most
readily explained in terms of optical principles described in sec. 1.2—the principles of
reflection and refraction.
The branch of optical science that deals with these phenomena in terms of rays and
wave fronts, rather than in terms of electromagnetic-wave theory, is called geometric
optics. The principles of geometric optics can be applied (as mentioned in sec. 1.2) only
when the dimensions of the optical surface are large compared with the wavelength. This
means, for example, that a reflector 3 meters in diameter would not behave in accordance
with a geometric-optics analysis at a frequency of 1 MHz, for which the wavelength is
300 meters. It would do so very well, however, at 10,000 MHz, for which the wavelength
is 3 cm. Generally speaking, therefore, antennas based on geometric-optics principles are
very-high-frequency devices. They are used mainly above 30 MHz, and arguably above
about 1 GHz they are more common than arrays.
Even when the dimensions of the antenna are large compared to the wavelength, the
principles of geometric optics cannot be applied to all aspects of its behavior. For
example, diffraction will occur at the edges of a lens or reflector and at small irregulari-
ties in the structure if such exist. Diffraction is not explainable by geometric optics. A
qualitative description may be given verbally or graphically in terms of Huygens’ prin-
ciple, as was done in sec. 1.2, but an exact description requires the mathematical expres-
sions of electromagnetic theory.
The diffraction effects are of secondary importance, and are ignored here, in the aper-
ture field formulation of lens and reflector antennas. However, they cannot be ignored
when an exact far-field analysis is desired. Diffraction must be considered, for example,
in analysis of the minor-lobe formation, and the main-beam pattern shape and width. The
method of geometric optics (ray-wave front analysis) is basically an approximate rather
than an exact method. But it is so nearly exact for many purposes, and is so much simpler
than the exact mathematics, that it is usually employed wherever possible. Also, it
conveys a useful intuitive conception of some aspects of antenna behavior.
227
228 Reflectors and Lenses
x 2 = 4 fz (6–1)
The focal point, or focus, at distance f from the vertex, is the point at which incoming
collimated rays will converge (as indicated in Fig. 6–1) or from which the diverging rays
of a point source will be collimated.
The bilateral symmetry of the curve is inherent in the x2 term of the equation; that is,
if a given value of z is chosen and the positive value of x is found that satisfies the equa-
tion, it will also be satisfied by minus the same value of x. The curve defined by (6–1)
extends to infinity in the +z-direction (and consequently also in the +x and −x directions).
* The word “reflector” in this chapter implies a nonplanar reflector—either one that is curved or a system of
plane reflectors set at different angles with respect to a reference plane. A simple plane reflector, such as
might be used in conjunction with an array, does not provide a focusing action.
Focusing and Collimation 229
* Since the symbol D is also used to denote the directivity of an antenna, the aperture linear dimension will
be designated Da in applications where it might be confused with directivity symbol. But often it is called
simply D because this symbol has traditionally been used, especially in referring to the f/D ratio.
230 Reflectors and Lenses
of constant phase) is created in the aperture plane. Hence the rays are parallel to the axis,
since rays are always perpendicular to a wavefront (Sec. 1–1).
x 2 + y 2 = 4 fz (6–2)
The intersection of any plane containing the z-axis with the paraboloidal surface is a
parabolic curve like the one shown in Fig. 6–2. (The intersecting plane in this figure is
simply the x-z plane.) The intersection of any plane perpendicular to the z-axis with the
paraboloidal surface is a circle. Thus the open mouth of a paraboloidal reflector, the
aperture, is circular (as in a conventional automobile headlight) if the reflector has
the same depth in all planes containing the parabolic axis, that is, the z axis.
It is possible, however, to “cut away” portions of a paraboloidal reflector in such a
way that it does not appear circular when viewed from a point on the parabolic axis.
Such cut paraboloids have certain advantages in some applications, as will be discussed
in sec. 6.3.
The second type of surface is the parabolic cylinder, formed by translating the
parabola of Fig. 6–2 in the direction of the y-axis, that is, by “moving it sideways.” The
intersections of all planes parallel to the x-z plane with the parabolic cylinder are parabo-
las like the one shown in Fig. 6–2. The intersections of all planes parallel to the y-z plane
with the parabolic cylinder are straight lines. If the cylindrical surface has a finite dimen-
sion in the y-direction, the reflector as viewed from a distant point on the z-axis will
appear rectangular, that is, it has a rectangular aperture. The parabolic cylinder has a
focal line, rather than a focal point, and a
vertex line. The approximate appearances
of the full paraboloid, a cut paraboloid,
and a parabolic cylinder are shown in
Fig. 6–3.
source of radiation. The strength of the radiation versus distance is described by a term
called space attenuation. Thus, knowledge of space attenuation is an important aspect of
the design of reflector and lens antennas. As may be seen in Fig. 1–2, chapter 1, rays
emanating from a point source diverge spherically. Thus, the wave energy from a point
source is constant within a tube of solid angle. Therefore, the power density varies in
proportion to 1/r2, where r is the distance from the source. On the other hand, as will
now be discussed, the power density varies in proportion to 1/r for radiation from a line
source.
Consider a line source that radiates radially with equal intensity perpendicular in every
direction from its axis. Now imagine a concentrically placed cylinder that intercepts the
radiated power. The cylinder’s surface area will vary in direct proportion to r, and hence
the power density at the cylinder would vary in direct proportion to 1/r. Therefore, as a
consequence of energy conservation, the power density from a line source varies with
distant r as 1/r; yet it varies as 1/r2 from a point source.
In Fig. 6–1, both focused (collimated) and unfocused rays are shown near the reflector.
The collimated (focused) rays are parallel, and thus the power density for focused rays
is independent of distance r near the parabola’s aperture (open mouth). Refer now to Fig.
6–2 to develop equations for the field strength (illumination) over the apertures of parabo-
loids (parabolas of revolution) and parabolic cylinders. Radiated rays diverge along the
distance r from the focal point to the reflector, but they are focused after reflection.
Consequently, for a paraboloid illuminated by a point source, the power density varies
as 1/r2, where r is the distance between the focal point and the reflector. Similarly, for
a parabolic cylinder illuminated by a line source, the power density varies as 1/r between
the focal line and the reflector.
In Fig. 6–2, note that zap − f is the distance that the aperture point zap is to the right
of the focal point. Recall that, for a focused antenna, all ray path lengths between the
focus and aperture are equal. Then, to determine r versus angle a by using Fig. 6–2, one
observes:
By equating (6–3) and (6–4), it can be seen that 2f = r + r cos a. Then, by using the
trigonometric identity (1 + cos a)/2 = cos2(a/2), one finds
f 1 + cos α α
= = cos2 ⎛⎜ ⎞⎟ (6–5)
ρ 2 ⎝ 2⎠
We now use the fact that, for a point source, power density varies as 1/r2. Then, from
(6–5), for a paraboloid with an isotropic radiator at its focus, power density pa on the
aperture versus angle a is
232 Reflectors and Lenses
2
pα ⎛ 1 + cos α ⎞ α
=⎜ ⎟ = cos4 ⎛⎜ ⎞⎟ (6–6)
p0 ⎝ 2 ⎠ ⎝ 2⎠
In (6–6), p0 is power density at the aperture center, where angle a is zero. Similarly, for
a parabolic cylinder, power density pa on the aperture versus angle a is
pα ⎛ 1 + cos α ⎞ 2⎛ α ⎞
=⎜ ⎟⎠ = cos ⎜⎝ ⎟⎠ (6–7)
p0 ⎝ 2 2
In (6–7), p0 is the power density at the center of the parabolic cylinder, where angle a
is zero.
To obtain electric field strength E versus angle a, the fact is used that E varies as
the square root of power density. Then, from (6–6) and (6–7), the variation of E versus
a along the aperture of a paraboloid and a parabolic cylinder can be expressed, respec-
tively, as
Eα 1 + cos α α
= = cos2 ⎛⎜ ⎞⎟ (for paraboloid ) (6–8)
E0 2 ⎝ 2⎠
and
1 + cos α α
= cos ⎛⎜ ⎞⎟
Eα
=
⎝
(for parabolic cylinder ) (6–9)
E0 2 2⎠
Equations (6–8) and (6–9) describe the variations in the aperture amplitude Ea versus
angle a caused by divergence of rays. Thus, (6–8) describes the amplitude on the aperture
of a paraboloid if the feed radiates isotropically, and (6–9) describes the amplitude dis-
tribution if a line source radiates uniformly at all angles about its axis.
Another useful expression is the attenuation, expressed in decibels, between focus or
focal line and the aperture, caused by the divergence of rays. From (6–8), this attenuation
in decibels for a paraboloid is
The equations in the present section describe the effects of space attenuation. These
equations provide a necessary step toward accurately determining the aperture strength
versus position along an aperture, and therefore often they are essential for calculating
the side-lobe levels and beamwidths available from reflector and lens antennas. Some-
Beamwidth and Directivity 233
times, however, as discussed in sec. 6.4, approximations are made for antenna designs
that use only the relative field strengths between the aperture center and its edges.
at the focus by the edges of the reflector in order for the entire surface to be illuminated.
The feed radiator is also known as the primary radiator, and its pattern is called the
primary pattern. The pattern of the entire antenna is called the secondary pattern. When
the term antenna pattern is used, the secondary pattern is meant. The primary pattern
may also be called the feed pattern.
6.2.3. Beamwidth
The secondary radiation pattern of a lens or reflector antenna is very similar to that of a
unidirectional planar array backed by a plane reflector, of the same size and aperture
shape. The beamwidth is given by a formula of the same type used in chapter 5:
λ
BW = k1 (6–12)
Da
where l is the wavelength, Da is the aperture width of the reflector or lens, and k1 is a
constant, depending on the units in which the beamwidth is to be expressed and on some
other factors. If the beamwidth (BW) is desired in degrees, and if both l and Da are
expressed in the same units of length, k1 is of the order of 60 to 70. This means, for
example, that if the aperture width Da is equal to 10 wavelengths, the beamwidth will
be about 6 or 7 degrees.
For a paraboloid of noncircular aperture, the value of Da to be used in (6–12) is the
aperture dimension in the plane in which the beamwidth is to be calculated. If the aperture
has different dimensions in different directions, the beam will have different widths in
these directions.
6.2.4. Directivity
The directivity (D) of a reflector antenna may be expressed in terms of the area (A) of
its aperture. The formula is the same as that for a large planar array:
4π A
D = k2 2 (6–13)
λ
The constant k2 will usually have a value between about 0.5 and 0.7, depending on the
shape of the aperture and the characteristics of the source of the radiation used.
The general relationships discussed in chapter 5 between beamwidths, directivity, and
gain are also applicable to continuous apertures. The factor k2 used above that relates
directivity with aperture area is commonly called aperture efficiency.
β β
τ1 = 40 log cos = − 40 logsec (6–14)
2 2
The taper due to the primary pattern, in decibels, is approximately given by the
formula:
β 2
τ 2 = −12
BW
(6–15)
where BW is the half-power beamwidth (as defined in sec. 3.5) of the feed radiator. (The
tapers are here expressed as power ratios less than one, or negative decibel values.)
Reflector Illumination 237
f/D 2b
0.10 273°
0.25 180°
0.50 106°
1.00 56°
(This table indicates why the f/D ratio is seldom smaller than 0.25 or larger than 0.50,
i.e., the required feed beamwidths become impractically large or small, respectively.)
A further tapering effect occurs if the aperture is not rectangular. Consider, for example,
an approximately hexagonal aperture as illustrated in Fig. 6–6, with height hc at the center
and he at the edge. Consider vertical strips of this reflector, of width w and height h. If
the illumination of the aperture is uniform, the total power radiated from a given strip is
proportional to its area, hw. Therefore the ratio of the power radiated from the center
strip to that from the edge strip is hc /he. This is equivalent to a horizontal illumination
taper (in the direction perpendicular to the strips) given in decibels by
hc
τ 3 = −10 log (6–17)
he
The analysis is of course more complicated for an aperture having a curved outline, or
for nonuniform illumination in the vertical direction, but the principles are the same. The
total illumination taper, expressed as the ratio of the edge illumination to the center illu-
mination in decibels, is the sum of the three component tapers, t1, t2, and t3.
An example illustrates the use of these principles and results. Suppose that it is desired
to design a circular-aperture paraboloidal-reflector antenna for a 30 dB side-lobe level.
A horn is to be used as the feed radiator, and the f/D ratio is 0.4.
From (6–16) it is first calculated that the value of b is given by
1
cot β = 0.8 − = 0.4875
3.2
∴ β = 64°
Hence the primary radiator must illuminate a reflector that subtends an angle of 128
degrees.
The taper due to reflector curvature is calculated from (6–14):
The table above indicates that for −30 dB side lobes a total taper of −14.5 dB is required.
The feed pattern must therefore provide a taper of −14.5 + 2.9 = −11.6 dB. This is the
required value of t2 in (6–15).
The next step is to determine the required primary-pattern beamwidth, BW. That is,
(6–15) must be solved for the value of BW corresponding to the given value of t2, which
is −11.6, and b = 64°. The result is
BW = 64 12 11.6 = 65°
This means that a feed horn with a 65-degree half-power beamwidth is required.
Equations (4–23a) and (4–23b) in chapter 4 provide for the design of a horn of a given
beamwidth. It is thus found that an optimum-flare pyramidal horn must have an E-plane
mouth dimension (dE) of 0.86l, and an H-plane dimension (dH) of 1.03l, where l is the
wavelength.
and also by the shape of the aperture outline. The formulas given are based on experi-
mental investigation rather than theoretical analysis.
Aperture tapers for arrays were discussed in chapter 5, where the similarities in the
effects of tapers on arrays and continuous apertures were briefly discussed. In addition,
Tables 6–1 and 6–2 in sec. 6.10 include a number of different continuous aperture dis-
tributions, and the beamwidths and relative gains that result therefrom.
In sec. 5.6.5 the Taylor distribution is said to be widely used in the design of efficient
and low side-lobe-level arrays, providing there are 20 or more elements (Taylor 1955).
Although widely used in the design of arrays having a large number of elements, the
Taylor distribution is basically a design for continuous apertures, and thus it provides an
efficient aperture design for continuous distributions. In other words, the Taylor distribu-
tion is generally considered to be an optimum aperture distribution that is applicable both
to arrays and continuous aperture antennas.
β
Parallel
τ1 = 20 log cos (6–18)
B
plates
2
obtained (i.e., narrowest beam in the desired direction, and greatest gain). Methods of
performing the necessary pattern and gain measurements are discussed in Ch. 9. A knowl-
edge of the approximate location of the phase center of the feed is of course helpful in
this procedure. In most cases it is at or near the geometric center of the feed. For a low
flare angle horn, it is near the center of the aperture plane (mouth). However, as the flare
is increased, the phase center moves inside the horn. For some types of feed radiators
the phase center may change if the frequency of operation is changed; thus this parameter
becomes another factor in the determination of the bandwidth of the antenna. (Certain
types of log-periodic radiators, discussed in sec. 7.2, are especially subject to this effect,
so that their otherwise excellent bandwidth properties may be severely limited when they
are used as feed radiators for reflectors and lenses. However, this is not true of all log-
periodic types of radiators.)
In addition to the problem of knowing where to position the feed, there is the com-
panion problem of insuring that it will remain at the correct position when subjected to
various stresses that may occur due to wind, gravity, ice loading, and so forth. The
obvious method of accomplishing this is to support the feed with a strong and rigid
mechanical structure. Unfortunately, in the case of a reflector antenna, there is the con-
flicting requirement that the feed support must not block too many of the reflected rays;
it must not be opaque if it is bulky, and if it is opaque it must not be bulky. A common
method of support in large paraboloidal reflectors is a tripod or quadrupod attached to
the reflector with its apex somewhat beyond the focal point, so that the feed may be
nested under the apex. The support booms may be either hollow metal tubes or solid
pieces of insulating material, such as fiberglass. When the reflector is small and of short
focal length, the waveguide or transmission line may furnish adequate support. Reflectors
with feeds supported by these methods are shown in Fig. 6–9.
The fact that the feed support (and the feed itself) blocks the aperture is a disadvantage
of many reflector antennas. However, if a paraboloid or parabolic cylinder is cut so that
the collimated rays do not “see” the feed radiator, the problem of feed blockage is thereby
solved. Figure 6–10 shows how this is done. The parabolic section actually used is shown
by the solid curve, while the full (symmetrical) parabolic section is shown dotted. This
arrangement avoids blockage but somewhat complicates providing the desired primary
pattern taper.
The feed-support is most severe for paraboloidal reflectors when the beam direction
must be steerable in any direction, or over a large range of angles. This means that some
method of moving the reflector (often called a “dish”) must be provided, and the feed
must be moved at the same time so as to keep it at the focal point; that is, the feed and
the dish must be mechanically connected by a rigid structure. The arrangements of Fig.
6–9 meet this requirement. Feed support is simplified when the antenna need not be
moved, that is, when the beam is always to point in a fixed direction. The feed support
can then be attached to the antenna base (e.g., the ground) rather than to the dish, and
light weight is not as important. This situation occurs, for example, with large reflector
antennas used for tropospheric-scatter point-to-point communication at VHF or UHF.
When the feed is located in the aperture of the dish, so that it blocks some of the
reflected radiation, the antenna pattern is adversely affected in two ways. The gain is
Reflector Illumination 243
Tripod
Waveguide-
boom
supported
horn feed
Feed
(waveguide runs
large trussed down boom)
mesh-surface
reflector
Small
solid-sheet
dish
FIGURE 6–9
Two methods of reflector feed support.
ignored. This is a reason for avoiding paraboloids with large f/D ratios, since a large f/D
means that a relatively narrow primary pattern is required, and this in turn requires the
feed horn or other radiator to be relatively large.
56
θE = degrees (6–19)
dE
246 Reflectors and Lenses
and
67
θH = degrees (6–20)
dH
where E and H refer to the horn’s E and H plane patterns. The symbols dE and dH are
the aperture dimensions (widths), expressed in wavelengths along the E and H planes.
Graphs of radiation patterns, developed from aperture analyses, known as universal
patterns, are available for a variety of horn types and dimensions (Balanis 2005; Love
1993, pp. 15–6–15–9, 15–35). These “universal” patterns provide details of the major
lobe shapes for aperture widths as small as 1.5 or 2l. However, designs for center-fed
reflectors and lenses often require feed-horn patterns with dimensions even smaller than
1.5l. Section 6.4.3 provides guidance on feed design for aperture widths less than 2l.
Sometimes identical E- and H-plane feed patterns are desired. However, for a conven-
tional tapered waveguide horn, these patterns are unequal and the side lobes are higher
in the E plane. This is because the E- and H-plane amplitude distributions of a waveguide
are different, with no amplitude taper in the E-plane. Horns designed with corrugated
inner walls near the horn’s mouth (aperture) reduce differences in E- and H-plane patterns
(Love 1993, pp. 15–43), but corrugated horns are heavier, larger, and more expensive
than smooth-walled horns.
2
β
−0.6
BW3
E (β ) = 10 (6–22)
Note that BW3 is the feed beamwidth between half-power points, but b is measured from
the beam center.
For primary feed patterns, antenna designers usually prefer working with the beam-
widths between the −10 dB points, instead of the 3 dB beamwidths. Reasons include
Radiation Patterns of Horn Antennas 247
(i) the −10 dB levels are nearer the reflector or lens edges, and the antenna
side-lobe levels are highly sensitive to the edge illumination, and
(ii) near the −10 dB angles the primary feed (horn) pattern changes more rapidly
with angle than at the −3 dB angle, causing overall antenna performance to
be more sensitive to the −10 dB locations.
Now let BW10 be defined as the full −10 dB beamwidth and let b0.1 be the angle
between the feed’s beam center and a −10 dB direction, that is, BW10 = 2b0.1. Then, the
relationship between BW10 and BW3 is as follows:
2
β 0.1
2
0.5 ⋅ BW10
20 log E (β 0.1 ) = −10 = −12 = −12
BW3
(6–23)
BW3
or
BW10 10
=2 ≈ 1.83 (6–24)
BW3 12
Finally, the amplitude E(b ) at angle b, measured from the beam center, in terms of the
full −10 dB beamwidth BW10 can be expressed as
2 2
β B
20 log E (β ) = −12 = − 40
BW10
(6–25)
BW3
or
2
β
−2
BW10
E (β ) = 10 (6–26)
on the continuous aperture are, at the coordinates of the array elements, the same as those
of the array elements.
A major difference in pattern design between arrays and reflector antennas is, however,
that reflectors are illuminated by a primary feed; thus, feed pattern shapes critically affect
the secondary (overall) patterns of reflector antennas.
The table of 2b (the total angle a paraboloid subtends) versus f/D ratio, that follows
equation (6–16), illustrates the range of feed pattern widths required to subtend the aper-
ture dimension D of a paraboloid versus its f/D ratio. However, as can be seen from
Fig. 6–10, smaller feed beamwidths are needed if a sectionalized (cut) parabolic cylinder
or a “cut” paraboloid is used. Larger f/D ratios (requiring smaller beamwidths) are usually
avoided because, for mechanical rigidity, short focal lengths are desired. On the other
hand, short focal lengths require wide feed patterns that require electrically small feeds;
and, as discussed in sec. 6.4.3, their patterns are generally less predictable than those of
larger feeds.
There are two general types of primary feeds: (a) the point source and (b) a line source.
A point source feed is used for illuminating a reflector having axial symmetry (e.g., a
paraboloid). Because of this symmetry and for effective use of the cross-sectional area,
the limits on compatible feed beamwidths relative to aperture widths constrain the ratio
of maximum-to-minimum principal plane patterns to approximately two or three. Con-
sequently, when point-like feeds are used, the range of possible azimuth-to-elevation
beamwidth ratios is quite limited. On the other hand, line-source feeds are usually used
where azimuth and elevation beamwidths differ greatly. For example, line-source feeds
for antennas having beamwidth ratios as large as 20 : 1 are commonly used with surveil-
lance radar.
As discussed for arrays in chapter 5, a pattern versus q and f is relatively easy to
determine, if the aperture distributions are separable in the principal planes and the linear
aperture patterns for the two aperture distributions are known. Then, the patterns may be
determined from the product of the two linear aperture patterns. Otherwise, as with
arrays, the complete pattern calculations require computational intensive, point-to-point
numerical calculations. For apertures with nonseparable distributions, patterns for the
principal planes are much simpler to calculate than the complete patterns. Then, also as
with arrays, the principal plane patterns may be calculated by using an equivalent line
aperture for each principal plane and integrating each line aperture distribution over the
relevant line aperture.
An example follows in the next section for calculating the illumination along the posi-
tive x-axis (in the upper half) of a parabola. In addition, the far-zone pattern in the x-z
plane is calculated, based on the calculated x-axis amplitude distribution. In this example,
the feed for the parabola is a line source. Thus, the space attenuation between feed and
reflector is caused by two-dimensional spreading of the radiation between feed and reflec-
tor. A line-source feed is commonly used to illuminate a parabolic cylinder. Then, the
illumination is separable because the x-axis illumination of the parabola’s aperture will
be independent of the illumination of the parabolic cylinder along the length of the line-
source feed (y-axis). Thus, as in the case of separable distributions for arrays, the normal-
ized far-zone pattern in the x-z plane is controlled by the x-axis distribution and the
250 Reflectors and Lenses
pattern in the y-z plane is controlled by the distribution by the y-axis distribution. Addi-
tionally as described in chapter 5 for arrays having independent (mathematically separa-
ble) x- and y-axis distributions, the pattern versus q and f may be determined as the
product of the two patterns created by the x- and y-axis distributions.
For a paraboloid or a sectional surface area thereof, the steps in calculating detailed
patterns are different than for a parabolic cylinder. This is so because the axial symmetry
of the reflector surface shape prohibits the illumination of the principal planes being
independent of one another. Therefore, as in the case of arrays having nonseparable dis-
tributions, a detailed principal plane pattern calculation uses an equivalent line aperture
distribution, and each line aperture distribution is integrated over the relevant line aper-
ture. Principal plane patterns are of course much simpler to determine than complete
patterns, which require point-to-point numerical calculations over the whole aperture.
Relative Strength in dB
metry in an aperture amplitude distribu-
tion along the x dimension.
–8
–12
6.5.3. Steps for Reflector
Antenna Design –16
Steps for developing a parabolic reflector
antenna follow. Similar steps are appro- –20
0 0.25 0.5 0.75 1
priate for other reflector shapes and Aperture X-Axis Coordinate
lenses.
Normalized Feed Pattern
A. Choose approximate overall
Space Attenuation
reflector aperture size. As first
Field Strength on Aperture
approximations:
• BW = 70l/D for overall FIGURE 6–13
aperture dimensions. BW is Relative feed pattern amplitude, space
a desired half-power prin- attenuation, and field strength in decibels
versus aperture x-axis coordinate.
cipal plane beamwidth in
degrees, D is antenna
dimension in plane of BW,
and l /D is dimensionless.
Normalized Pattern in dB
0
• G = 26,000/q1q2 for gain, –10
where q1 and q2 are princi- –20
pal plane beamwidths in –30
degrees. Special beam –40
shapes will result in smaller –50
–60
gain. For example, a cose- –30 –20 –10 0 10 20 30
cant-shaped pattern typi- Observation Angle Relative to Broadside
cally causes gain to be
FIGURE 6–14
reduced by factor of 0.7
(1.5 dB). Normalized far-zone pattern in x-y plane.
• Allow for increased
antenna size needed for structural integrity.
B. Select a general reflector shape: e.g., a paraboloid or a parabolic cylinder
• A simple pyramidal feed horn usually suffices for a paraboloid, if focal-
length to aperture-dimension ratios are within acceptable bounds, for
example, see table that follows equation (6–16). Consequently, for a parab-
oloidal reflector, the ratio of large-to-small principal-plane beamwidths is
ordinarily limited to less than 2 : 1.
• Parabolic cylinders require line source feeds, which permit large ratios of
principal plane beamwidths and wide-angle scanning. Line source feeds
are, however, more difficult to design and fabricate than horns.
252 Reflectors and Lenses
C. Decide where to locate the feed with respect to the aperture, i.e., whether or
not to use an offset feed for reducing aperture blockage and minimizing
impedance mismatch.
D. Perform trades between focal length and feed pattern beamwidths to furnish
reflector edge illumination that is at least 10 dB below reflector peak illumi-
nation, when effects of space attenuation are included. A larger edge illumina-
tion results in generally unacceptable reflector spillover and loss of gain. A
lower edge illumination results in lower side lobes, slightly wider beam-
widths and lower gain.
E. Choose feed configuration and dimensions based initially on the reflector
edge illumination. Based on estimated feed beamwidths, calculate expected
shape of feed’s major lobe versus angle, by using either (6–22) or (6–26). It
may be necessary to make feed measurements, especially if −30 dB or lower
side lobes are required or if the estimated feed dimensions are less than 2l.
F. Calculate illumination versus aperture coordinates based on expected shape
of feed’s major lobe and reflector geometry.
G. Calculate patterns in principal planes. Special requirements may also require
pattern specifications in other planes.
H. Continue the iterative process of selecting focal length and feed configuration
until acceptable secondary patterns are calculated. Before desired patterns
are obtained, it may be necessary to revise the original estimate of overall
reflector dimensions.
I. Patterns of the actual antenna should be measured for assurance, especially
if the side-lobe or beam-shape requirements are stringent.
England. When completed in 1957, it contained the largest fully steerable reflector. There
are several in the 45- to 60-m size range, and dishes of 20 and 25 meter diameter are
fairly common. However these are still “huge” antennas, especially if they are
steerable.
plane wavefront of the reflected radiation. But perfect accuracy is unattainable, and
construction of a precise surface is very expensive. It is desirable, therefore, to know
what degree of inaccuracy can be tolerated without serious sacrifice of performance.
Theoreticians have studied the question of how much deviation from a true parabolic
surface is permissible, assuming that the deviations are of a random nature—that is, that
they are not of a regular nature. An intuitive understanding of the effect of surface irregu-
larities can be obtained from Fig. 1–6 (chapter 1). As shown, the total reflection consists
of a specular (desired) component and a scattered or diffuse (undesired) component.
The generally accepted rule-of-thumb figure is that the permissible deviation is 1/32 to
1/16 wavelength. Thus the irregularity of the reflector surface in effect determines the
maximum frequency at which it can be used successfully; the higher the frequency of
operation desired, the more precisely accurate the surface must be. Appendix I discusses
relatively simple calculations on the effects of random phase errors on antenna patterns.
An analysis performed by Ruze (1966) gives detailed information on this matter. The
random deviations of the surface from a true parabolic shape cause diffuse reflection that
increases the side-lobe level and reduces the gain. Ruze has shown quantitatively how
these effects are related to the root-mean-square (rms) deviation of the reflector from the
exactly correct surface. He showed that a given rms deviation has a greater effect on
side-lobe level in a small reflector than in a large one, and also that the effect depends
on the correlation interval of the irregularities. Correlation interval is a statistical concept.
If the individual “bumps” of the surface extend over a small area, the correlation interval
is small. If they extend over a larger area, as would be true of a “wavy” rather than a
“bumpy” surface, the correlation interval is larger. In terms of these concepts, Ruze
showed that a given magnitude of rms deviation from the true surface causes a greater
increase of side lobes and reduction of gain when the correlation interval is large than
when it is small.
His paper contains curves giving the results for specific cases. As an illustration, a
paraboloidal reflector of circular aperture, having a diameter of 25 wavelengths and
typically tapered illumination, will have an additional side-lobe level of −20 dB when
the rms deviation of the reflector is 1/32 wavelength if the correlation interval is
one wavelength. But if the correlation level is only a quarter of the wavelength, the
additional side-lobe level will be only about −35 dB. By “additional side-lobe level” it
is meant that the relative power of the side lobes due to reflector inaccuracies will be
added to the relative power of whatever side lobes would be present if the reflector were
perfect. As an example of how this “addition” would be computed, suppose that the side-
lobe level with a perfectly accurate reflector would be −20 dB, which corresponds to a
power level of 1/100 or 0.01 of the main-beam level. Then an additional −20 dB side-
lobe level would add another relative power level of 0.01, giving a total level of 0.02.
Thus the combined effects give a net side-lobe level of −17 dB. On the other hand, an
additional −35 dB side-lobe contribution corresponds to an additional 1/3000, or roughly
0.0003, so that the total in this case becomes 0.0103, which is negligibly different from
−20 dB.
Another factor of great importance is the rigidity of the reflector—its resistance to
deformation due to wind or gravitational forces (including additional loading that may
Corner-Reflector Antennas 255
be imposed by ice in the colder climates). Gravitational forces are usually important only
in steerable reflectors that may point sometimes at the horizon and sometimes at the
zenith or intermediate elevation angles. If the reflector shape is paraboloidal at one of
these angles, the loading at other angles may be such that it will deform, that is, the edges
of the dish may sag differentially. This effect is important only for large heavy reflectors.
Deviations of this type will have a serious effect on the pattern, however, since they are
of the “large correlation interval” type. A related effect is the possible variable effect of
gravity on sag of the feed support structure, for different elevation angles of the antenna.
The net effect of reflector deformation and feed-support deformation can be separated
into two parts. One is a deterioration of the pattern—increase of side-lobe level and
reduction of gain; the other is a shift in the direction of the maximum of the beam, which
may cause angle-measurement errors in a radar or radio-astronomy system unless the
shift is known and corrected in interpretation of data.
Distortion of the antenna by wind forces can also be serious, and resulting shifts of
the beam direction cannot be corrected by a “calibration” as can be done for gravitational
shifts. Wind forces, as previously mentioned, can be reduced by using perforated or mesh
material for the reflector surface (“skin”). However, the most serious consideration in
the reduction of wind force is usually that the antenna may be overturned in a high wind,
or that the steering motors may be unable to overcome an opposing wind force. Most
large steerable antennas, in fact, will not operate in very high winds, and are usually
stowed (locked in a fixed position) when winds exceed 30 or 40 mph. Some types of
moderate-sized steerable antennas, however, such as those of radars for air traffic control
or for military and naval applications, are capable of operating in fairly high winds.
feed-to-vertex distance, d, is made equal to half the side length, l. Thus the design equa-
tions are
l
d= (6–27)
2
(The second equation is obtained from fact that the hypotenuse of a right triangle is the
square root of the sum of the squares of the other two sides.) The distance d is generally
made to be between 1/3 and 2/3 wavelength. A thorough discussion of corner-reflector
antenna design is given by Kraus and Marhefka (2002).
These equations are valid only within the stated range of values of Da. Within
this range, increasing the size does not greatly affect the beamwidth and gain, but the
radiation resistance and the bandwidth are increased. It is apparent from the fact that the
maximum allowable value of Da is about two wavelengths that the beamwidth in this
dimension is fairly broad; it is about 40 degrees. In the other dimension, as with the
parabolic cylinder the beamwidth is determined entirely by the feed radiator, which may
be a linear array with a fairly narrow beam. If a beamwidth in the corner-aperture dimen-
sion of appreciably less than 40 degrees is desired, a parabolic-cylinder reflector should
be used instead of the corner. (The same feed may be used, positioned on the focal line
of the parabolic cylinder.) The directivity of a corner-reflector antenna is approximately
10 dB greater than that of a dipole or of a linear dipole array by itself.
The surfaces of a corner reflector are frequently, in fact usually, made of spaced wires
or tubes parallel to the vertex, rather than solid sheet metal. The spacing is a small frac-
tion of a wavelength, and for polarization parallel to the wires the reflection is practically
as good as that from a solid reflector. If the feed is a dipole or dipoles parallel to the
vertex, as is almost always the case, the polarization will be correct. The use of spaced
wires or tubes reduces the weight and the wind resistance.
Corner reflectors with corner angles other than 90 degrees are sometimes used. A
60-degree corner has a slightly higher directivity than a 90-degree corner, but the
sides must be longer to realize the increase. (The
vertex-to-feed distance must also be greater.) It is
likely that the greater directivity may be obtained
Dipole with less total reflector surface by using a parabolic
cylinder, or by stacking two 90-degree corners as
Reflectors shown in Fig. 6–16. This arrangement is in effect an
Direction array of corner reflectors. The two line sources must
of beam
of course be fed in phase with each other. (Inciden-
Dipole
tally, parabolic cylinders may also be stacked in this
way. The gain will be about the same as that of a
single reflector of the same total aperture, but there
FIGURE 6–16 may be some advantages in reduction of reflector
Stacked corner reflectors. depth.)
Lens Antennas 257
Source
(focal point) Lens axis
FIGURE 6–17.
Collimation of rays by a simple converging lens.
levels, and of the relations between aperture size, beamwidth, and gain. Feed radiator
requirements are in general quite similar, except that aperture blocking effects do not
occur, as already noted. As will be discussed in sec. 6.7 in more detail, a collimated beam
may also be formed in a direction not parallel to the lens axis, by moving the source
(feed) away from the focus in the focal plane—that is, perpendicularly to the lens
axis—as was also indicated to be true for parabolic reflectors, for limited deviations from
the optical axis.
FIGURE 6–18.
Lens surface curvature combinations for n > 1.
source—must have at least one convex surface.) If the index of refraction is less than
one (n < 1), one or both of the curved surfaces will be concave, whereas for n > 1 one
or both will be convex. It is also possible to have one surface convex and one concave,
provided that the convex surface has the greater curvature for n > 1 and vice versa for
n < 1. The various possibilities are shown in Figs. 6–18 and 6–19. The curvatures are in
all cases calculated to produce the right amount of total ray bending at the two surfaces
so that the rays from a focal point or line will issue from the opposite lens surface all
parallel, that is, collimated.
Lenses having simple curved and plane surfaces may be very thick, and therefore
excessively bulky and heavy. Also, waves passing through the thickest portions of the
lens may suffer considerable loss of power, since the lens structure or material may be
dissipative. To offset these effects, lens surfaces can be zoned. That is, at definite dis-
tances from the center of the lens, the surface is “stepped” so that its total thickness is
reduced. The depth of each step is such that the rays passing through the lens on each
side of the step have path lengths that differ by a full wavelength; hence they are in-phase.
The wavefront issuing from the lens is therefore the same as if zoning were not used.
Figure 6–20 shows cross sections of zoned lenses corresponding to the unzoned types
of Fig. 6–18a. Although zoning provides the benefits mentioned, it also introduces
260 Reflectors and Lenses
a close to the cutoff value (half of the free-space wavelength) so that the phase velocity
will be high. The plates must operate in the fundamental TE mode (electric field parallel
to the plates), and thus such a lens will function properly only for a proper polarization
of the feed. However, it is also possible to have two sets of plates running in perpendicular
directions so that a set of square waveguide channels is formed. This is known as “egg
crate” construction, and it may be used with waves of any polarization.
n (r ) = 2 − (r R )2 (6–29)
where R is the radius of the sphere and r is the radial coordinate of any point within the
sphere. It is apparent that n = 1 at the surface of the sphere (r = R) and has a maximum
value of 2 at the center of the sphere (r = 0). Such a lens has an interesting and useful
property; it will collimate the rays from a feed source placed anywhere on its surface.
The collimated rays will emerge on the opposite side of the sphere from the feed point,
traveling in the direction of the line from the feed point through the center of the sphere.
Thus a beam can be caused to point in any direction by moving the feed to an appropriate
point on the lens. The behavior of the rays in a Luneberg lens is shown in Fig. 6–21. A
sphere having the required variation of refractive index can be made using either an
artificial dielectric medium or concentric spherical shells of solid dielectric material of
different indexes of refraction. The effective refractive index of a solid dielectric material
can be reduced to a desired value for this purpose by manufacturing it in the form of a
“foamed” solid containing air spaces.
There are also two-dimensional instead of spherical (three-dimensional) Luneberg
lenses. Two-dimensional variable-index-of-refraction lenses focus in only one plane
(Johnson 1993, p. 16–26). They have dielectric properties like a thin cylindrical slice cut
through the center of a 3-D Luneberg lens. An early type consisted of nearly flat plates
partially filled with dielectric material and it operated in the TE10 waveguide mode. In
addition, there have been flat-plate lenses that operated in the TEM mode. One of these
used concentrate rings of different dielectric constants and another used partially filled
262 Reflectors and Lenses
Collimated rays
Lens
Plane
Radiating source wavefront
(feed)
FIGURE 6–21.
Rays in a Luneberg lens.
dielectric plates that are tapered to zero thickness at the lens edges. Another 2-D type
lens is the geodesic Luneberg lens, described in sec. 6.8.8.
r PP L QQ1 Q1Q2
+ 2 = + + (6–30)
λ0 λ d λ0 λd λd
te
pto
ym
As
Note from Fig. 6–22 that PP2 = Q1Q2
and r cosb = (L + QQ1). Therefore, from r
P P2
where
l0 = wavelength in free space (air or
vacuum)
FIGURE 6–22.
ld = wavelength in the dielectric lens
Path lengths in a dielectric lens.
From equation (1–19), chapter 1, the
index of refraction n of a lossless sub-
stance is expressed as
c µε
n= = (6–32)
v µ0 ε 0
For most dielectric materials m = m0, and thus for a lossless material usually
n = ε ε0 (6–33)
where e/e0 is the relative permittivity of the material having index of refraction n.
Now, from the general relationship fl = v and (6–32), for a material having index of
refraction n, and with its wave velocity vd and wavelength ld
c f λ0 λ0
n= = = (6–34)
vd f λ d λ d
Then, multiplying (6–31) by l0, and solving for r the following is obtained
(n − 1)L
r= (6–35)
n cos β − 1
Equation (6–35) gives the shape of the lens, and it describes a hyperbola having
asymptotes at an angle b0 with respect to the axis of symmetry. With r very large, it
can be seen from (6–35) that ncosb is then approximately one, and thus cosb0 ⬵ (1/n)
and
264 Reflectors and Lenses
For convenience, a lens is said to be (a) “spherical” if the surfaces are generated by
a rotation about the line FQ2 and (b) “cylindrical” if it has symmetry along a focal line.
Risser (1949, ch. 11) addresses both “spherical” and “cylindrical” lenses. Parenthetically,
the F number of a lens is the ratio of the focal distance L to the lens diameter D; thus,
F number = L/D.
Effects of aperture tapering by primary feeds, and losses caused either by air-dielectric
reflections and absorption for dielectric lenses are addressed by Risser (1949, ch. 11).
λ0
λg = (6–37)
1 − (λ 0 2 a )2
λ0
n= = 1 − (λ 0 2 a )2 (6–38)
λd
L QQ1 r PP
+ = + 1
λ0 λg λ0 λ g
or
L r PP − QQ1 r L − r cos β
= + 1 = + (6–39)
λ0 λ0 λg λ0 λg
Finally, by substituting n for l0/lg in (6–39) and solving for r, one obtains
(1 − n ) L
r= (6–40)
1 − n cos β
where
l0 = wavelength in free space (air or vacuum)
lg = wavelength between the metal plates
The reader will notice that (6–40) for the metal plate lens is equal to (6–35) for the
dielectric lens. However, because n < 1, to make both the numerator and denominator
of (6–40) positive, one had to first multiply both by −1. As noted previously with n > 1,
(6–35) has the shape of a hyperbola. However, with n < 1, (6–40) has the shape of an
ellipse.
Generally, a metal plate lens is more frequency sensitive than a dielectric lens, in that
it focuses only for a relatively small bandwidth. Recall that (6–40) must be satisfied for
the lens to be perfectly focused. However, for waveguide, n is frequency dependent.
Thus, a metal plate lens can be perfectly focused only at one frequency. As with dielectric
lenses (Fig. 6–20), metal plate lenses are zoned (stepped) to reduce weight and to increase
bandwidth. Even so, depending on the desired side-lobe level, good metal-plate lens
performance is limited to bandwidths of
only a few percent (Kock 1946).
P P1
r
6.8.8. Geodesic Luneberg
Lenses
b
Fermat’s principle of least time is a fun- Q Q1
F
damental law of optics. It may be
expressed as follows: “When a ray passes L
constrained to follow a surface. Thus, the shortest path that a person can travel between
New York and London is a geodesic, because the shortest path is constrained to follow
the earth’s surface. Therefore, for a medium that provides a constant propagation velocity,
a natural result of Fermat’s principle is that the path of an electromagnetic wave is a
geodesic. Consequently, focusing occurs by forcing electromagnetic waves in air to travel
along curved surfaces.
To appreciate what a geodesic lens accomplishes, we first consider the length required
of a sectorial H-plane horn (Fig. 4–21, chapter 4), if it has an aperture width dH of 50l
at l = 0.1 m. The far-zone equation L = 2D2/l of Appendix F, that results in 22.5° phase
error across an aperture, indicates a required horn length L of 500 m. Obviously, this
length is impractical. However, as will now be discussed, a geodesic lens of 50l i.e., 5 m
diameter can provide a 5 m aperture with no phase error.
Geodesic lenses focus in one plane by having equal physical path lengths between a
point and a line, as now described. Advantages over dielectric lenses include very wide
bandwidth, and excellent impedance properties. Additionally, mechanical tolerances are
not severe, and appropriate manufacturing techniques coupled with good shop practice
can produce effective geodesic lenses at moderate cost.
A basic geodesic lens consists of two evenly spaced, rotational symmetric, conducting
surfaces. Ordinarily, propagation within the surfaces is in the TEM mode for which wave
velocity is that of free space. Thus, the focusing between a feed point and the aperture
results because all path lengths are equal, and consequently such a lens is focused over
a very wide bandwidth. Figure 6–25 shows an example lens with dashed lines indicating
two of the equal path lengths between a point and a straight line. The paths of the rays
as seen from a top view, when projected onto a planar surface, are the same as the paths
within a thin cylindrical slice cut through the center of a three-dimensional Luneberg
lens. Consequently, a geodesic lens focuses in only one plane.
Figure 6–26 shows cross sections of previously developed geodesic Luneberg lenses.
Note the input and output lips designated oa and ob at the edges (rims) of the lenses.
The tin-hat lens is symmetrical (ao and ob are identical) and has horizontal lips. Thus
focusing is maintained when a feed is rotated along the rim through 360°. However, the
input and output lips (ao and ob) of the helmet and clam-shell lenses are different and
thus destroy lens symmetry. These lips are provided to allow beam scanning with multi-
ple, rotating feeds over a limited angular
sector.
Line of
Constant Phase Figure 6–27 illustrates the use of a geo-
desic lens for exciting a line source reflec-
Focal tor feed for a scanning antenna. The
Point geodesic lens and nonfocusing flat-plate
extensions transform a point source at the
Conducting Surfaces lens feed horn into a linear wave front
FIGURE 6–25. along the reflector feed assembly. The
Paths through a geodesic lens for two of the linear wave front from the lens is directed
rays of a collimated beam (from Johnson diametrically opposite the lens feed point.
1962), courtesy of The Microwave Journal. Therefore, the angle of incidence at the
Lens Antennas 267
LENS AXIS
LENS AXIS
o o
o o
b b
a
a
TIN-HAT LENS HELMET LENS
LENS AXIS
o o
a b
CLAM-SHELL LENS
FIGURE 6–26.
Sectional views of mean surfaces for common types of geodesic lenses (from Johnson 1962),
courtesy of The Microwave Journal.
LENS AXIS
MAIN SPINNINGS
REINFORCING RINGS
INSERT
SPINNINGS
SP ACER BUSHING
FEED ARC
FIGURE 6–28.
A sectional view through a 16 GHz geodesic Luneberg lens (from Hollis and Long, 1957;
© 1957, IEEE).
Figure 6–28 shows a cross section of one of the 16 GHz clam-shell lens. Each lens
consists of two closely nested aluminum spinnings; each spinning is made of two sec-
tions: a main spinning, and an insert that replaces a section of the main spinning between
the periphery and the reinforcing ring. This insert forms an internal feed arc that permits
sectoral feed horns to rotate in a plane about the lens axis. The insert subtends 110° on
the lens periphery, to permit the lens to focus for all feed positions over the 40° scan
sector. The fabrication tolerance for the lens surfaces was set at 32 1 inch. Although this
tolerance was maintained over most of the lens surface areas, greater deviations did exist.
Pattern measurements indicated a maximum side-lobe level of −25 dB.
A narrow-beam geodesic Luneberg lens was used in the AN/MPS-29 (XE-1), a 70-
GHz radar (see sec. 7.7.2). The lens diameter is 350l, that is, 60 inches (152 cm), that
provides a 0.2° scanned beam (Long, Rivers, and Butterworth 1960). Spacing between
the lens conducting surfaces is about l.4l. This dimension is oversized for the TEM
mode, yet it was successfully used to minimize lens attenuation. Measured side-lobe level
of the lens-reflector combination did not exceed −22 dB for any of the feed horns or scan
directions. The lens was spun of mild steel, machined, copper-plated, and protected with
a thin coating of Irilac, a low loss plastic surface coating. Templates for the conducting
surfaces were made on a digitally controlled profiler, and the surfaces were machined on
a tracer controlled boring mill.
A folded geodesic Luneberg lens is an entirely different configuration than the lenses
of Fig. 6–26 (Goodman et al. 1967). Its height-to-diameter ratio is reduced by “folding”
the lens surfaces. A helmet-type lens (Fig. 6–26) with one fold, located symmetrically
about the lens axis, was made for 70 GHz. To affect the height reduction, the upper-dome
sections of the lens were reversed to become bowls, with their troughs along the center-
line of the domes. The lens aperture was approximately 23 inches (58 cm) and its
maximum mean-surface height was 3.6 inches (9.4 cm). Measured side lobes of the
folded lens were, surprisingly, smaller than for the unfolded lenses. The side lobes were
“lost” in the receiver noise, which was at least 36 dB below the peak of the main lobe.
The −3 dB beam width was 0.56°, the same as expected for an unfolded lens. However,
the folding broadened the base of the main lobe at angles wider than the −10 dB angles.
Thus, the folded geodesic lens provided reduced profile height and lowered side lobes,
Pattern Calculations for Continuous Apertures 269
and these were achieved with negligible degradation of antenna gain or beamwidth. This
lens was installed in an azimuth scanning, 70 GHz armored-vehicle mounted radar that
provided adequate vision for night driving (Dyer and Goodman 1972).
Geodesic Luneberg lenses have been used in other types of scanners. For example,
Allen (1956) designed a tin-hat lens that was sector-scanned with multiple waveguides
in a configuration called organ-pipe feed; and Schaufelberger (1960) employed the ultra-
wide-band millimeter wavelength feature of a geodesic Luneberg lens to provide the
detection sensitivity needed for target detection by radiometry.
Lens
q
Focal Lens
point axis
Plane
wavefront
Source Collimated rays
Spherical
wavefront
(a)
Plane
wavefront
Focal point
Reflection
q Parabolic
axis
Collimated rays
that are based on sound electromagnetic principles, yet certain assumptions called high-
frequency, scalar approximations are used (Silver 1949, pp. 169–99). High frequency
refers to the aperture being large enough that the effects of radiation from aperture edges
are negligible; and scalar refers to the field over the aperture being almost completely
linearly polarized, with only a small fraction of the energy being cross-polarized.
z
x
b
q
q
a
y a z
f
b y
f
x
(a)
(b)
FIGURE 6–30.
Coordinates for a rectangular aperture located in the x-y plane.
272 Reflectors and Lenses
given in Johnson (1993, p. 2–15), the normalized radiation pattern in the x-z plane of a
continuous line source on the x-axis may be expressed as
2π x sin θ ⎞ ⎤
A ( x ) exp ⎡⎢ j ⎛⎜
1 L2
E (θ ) = ∫ ⎟⎠ ⎥ dx (6–41)
L − L 2 ⎣ ⎝ λ ⎦
where E(q) = electric far-field pattern versus observation angle q, normalized for peak
amplitude of unity
q = observation angle measured from normal to aperture
L = overall length of aperture
x = position along the aperture: −L/2 ≥ x ≤ L/2
A(x) = a(x)eja(x)
a(x) = aperture amplitude versus position x
a(x) = aperture phase versus position x
The analogy between equations for linear arrays and linear continuous apertures may be
apparent. Equation (6–41) is a summation of amplitude and phase contributions, at
each observation angle q from aperture currents of infinitesimal length dx. For mathe-
matical convenience, solutions are usually expressed in terms of the variable u, where
πL
u= sin θ .
λ
The factor 1/L normalizes (6–41), so that the maximum of E(q) is unity. Thus,
an unnormalized pattern g(q) in the x-z plane for a continuous line source along the
x-axis is
2π x sin θ ⎞ ⎤
A ( x ) exp ⎡⎢ j ⎛⎜
L 2
g (θ ) = ∫ ⎟⎠ ⎥ dx (6–42)
−L 2 ⎣ ⎝ λ ⎦
To provide a feeling for the mathematics of (6–42), its solution will be obtained
when its integrand is simplified by letting A(x) = 1. Thus, the radiation pattern
will be determined for a line source having constant amplitude of unity and zero
phase along its length. To accomplish this, two mathematical relationships are used;
namely,
1
∫ exp (cz ) dz = c exp (cz )
and
πL
where u = sin θ . Now, since sin u/u equals unity when u is zero, the normalized
λ
amplitude pattern for a uniformly illuminated line source is
sin u
E (u ) = (for constant amplitude and phase ) (6–44)
u
Another line amplitude distribution frequently used is cosine amplitude and constant
phase. In this case, the amplitude is made unity at the aperture center and zero at its
edges. Then A(x) = a(x) = cos(px/L) for −L/2 ≤ x ≤ L/2. From Silver (1949, p. 187) and
other references, the pattern obtained using this cosine distribution with (6–41) is
cos u
E (u ) = 2
(for cosine amplitude distribution ) (6–45)
1−
2u
π
πL
where u = sin θ .
λ
Graphs of E(q) from (6–44) and (6–45), expressed in decibels, for the uniform and
cosine amplitude distributions are included in Fig. 6–31. There the first side lobes for
the uniform distribution are only 13.2 dB less than the peak of the major lobe. On the
other hand, the peak side lobes for the 0
Pattern strength (dB)
TABLE 6-1 Beamwidth, Maximum Side Lobes, and Relative Gain Versus Linear Aperture
Amplitude Distribution (from Silver 1949, p. 187)
that coordinate x is between −1 and +1. Notice that the first side lobes are appreciably
reduced by the tapering of amplitudes to small values at the aperture edges. The relative
gain in Table 6-1 is commonly called aperture efficiency. It is the gain relative to the
gain available if the aperture were illuminated with constant amplitude and constant
phase.
In the previous discussions, the phase over the aperture is assumed constant. A major
effect of aperture phase errors is to reduce the gain and broaden the major lobe. In addi-
tion, phase errors reduce the depth of pattern minima between the lobes and raise the
side lobes. Very large phase errors over the aperture can cause the main lobe to be split
into parts and cause large side lobes. For a detailed discussion on phase errors along a
continuous aperture, the reader is referred to Johnson (1993, pp. 2-21–2-27).
2π
a2 b2 j sin θ ( x cos φ + y sin φ )
g (θ , φ ) = ∫ A ( x, y ) e
a 2∫ b 2
λ dxdy (6–46)
− −
and it expresses the relative electric field strength at a fixed distance from the aperture
origin versus the angles q and f. Here a is the aperture length along the x-axis, and b is the
Pattern Calculations for Continuous Apertures 275
length along the y-axis. Note that the aperture distribution A(x, y) is a function of two vari-
ables, and thus it describes the amplitude and phase at the aperture’s x and y coordinates.
With reflector and lens antennas, often the dependencies of A(x, y) on x and y are interde-
pendent so that it is difficult or impossible to describe A(x, y) with a simple equation. Then,
it may be necessary to solve (6–46) by point-to-point integrations over the aperture.
The simplest g(q, f) calculation is one for a line source along either the x- or the y-
axis. Then from (6–46), for an x-axis continuous distribution
2π
a2 j x sin θ cos φ
g (θ , φ ) = ∫ A(x)e λ dx (6–47a)
−a 2
because all y values are zero. Notice that the integrand in (6–47a) is the same as in (6–42),
except that the product sin q cos f replaces sin q. Consequently, solutions of (6–47a) are
the same as solutions for (6–42), except that their solutions in terms of the variable u,
where u = (pL/l) ⋅ sin q, are replaced by
Likewise, for a y-axis continuous distribution, for which all x coordinates are zero
2π
b2 j y sin θ sin φ
g (θ , φ ) = ∫ A ( y) e λ dy (6–48a)
−b 2
Then, similar to solutions for (6–47a), solutions of (6–48a) are the same as solutions for
(6–42), except that the solutions in terms of the variable u, where u = (pL/l) ⋅ sin q, are
replaced by
2π 2π
a2 j x sin θ cos φ b2 j y sin θ sin φ
g (θ , φ ) = ∫ A1( x ) e λ
∫
dx ⋅ A2 ( y) e λ dy (6–50)
−a 2 −b 2
Equation (6–50) is simplified further if patterns only in the x-z and y-z planes are
considered. Then, the two principal plane patterns are dependent on aperture distributions
along the x- and y-axes, as follows:
2π
x sin θ
g (θ , 0 ) = ∫
j
A2 ( y )dy ⋅ ∫
b2 a2
A1( x ) e λ dx ( X -Z plane ) (6–51a)
−b 2 −a 2
2π
y sin θ
g (θ , 90 ) = ∫
j
A1( x )dx ⋅ ∫
a2 b2
A2 ( y )e λ dy ( y-z planee ) (6–51b)
−a 2 −b 2
Notice that the bracketed terms of (6–51a) and (6–51b) are constants. Also, notice that
the integrals outside the bracketed terms are similar in form to (6–42). Thus, except for
a multiplicity factor, (6–51a) and (6–51b) for planar apertures are describable by radia-
tion patterns of line sources. Therefore, knowledge of linear aperture patterns is often
beneficial to an understanding of the patterns of continuous planar apertures.
The simplest planar aperture distribution has unity amplitude and zero phase, and its
pattern versus q and f can be determined with relative ease. For this case,
A ( x, y ) = 1⋅ exp 0 × 1⋅ exp 0 = 1.
Then, since A(x, y) is separable, g(q, f) is the product of two integrals and each is the
pattern of a line source of constant amplitude and phase. Thus, from (6–47b), (6–48b),
and (6–50), the radiation pattern for a distribution of constant amplitude and phase
becomes
2π
x sin θ
g (θ , φ = 0°) = ∫
a2
b 2 A ( x, y ) dy ⋅ e j
− a 2 ∫− b 2
λ dx (6–53a)
Similarly, the pattern g(q, f = 90°) in the y-z principal plane for an equivalent line source
along the Y-axis, can be obtained by rearranging the terms in (6–46) as follows:
2π
y sin θ
g (θ , φ = 90°) = ∫
b2
a 2 A ( x, y ) dx ⋅ e j
− b 2 ∫− a 2
λ dy (6–53b)
Now compare (6–53a) with (6–42) for the x-z plane pattern of an x-axis linear aperture.
Note that the bracketed integral of (6–53a) functions for (6–42) as the equivalent linear
aperture distribution along the x-axis. However, this bracketed integral (the equivalent
linear aperture distribution) is a function of x and y. Thus, to evaluate the bracketed
integral at each x coordinate, A(x, y) is integrated over all y coordinates between −b/2
and b/2. Likewise, the pattern in the f = 90° plane is attained from (6–53b). The principles
of evaluating (6–53a) and (6–53b) are beneficial in envisioning the effects of aperture
tapers, as is discussed below.
A simple mathematical example is now given, which is that of determining the equiva-
lent linear aperture distribution and the pattern for a uniformly illuminated circular
aperture. Then, the aperture amplitude may be expressed as A(x, y) = 1 over the circular
278 Reflectors and Lenses
aperture, but it is zero beyond. For specificity, let the radius of the aperture be unity.
Then, the x and y coordinates of the outer aperture edges are interrelated by x2 + y2 = 1.
In other words, at each x-value, the y coordinate at the aperture edge is ye = (1 − x2)1/2.
Thus, for the bracketed term of (6–53a), the limits of integration b/2 and −b/2 may be
replaced by ye/2 and −ye/2. Consequently, the bracketed term of (6–53a) then becomes
1⋅ dy = ye = (1 − x 2 )
b2 ye 2 12
∫−b 2 A ( x, y ) dy = ∫− y e 2
(6–54)
Thus, even though the amplitude A(x,y) is unity when not zero, it is in fact a nonseparable
amplitude distribution. This is because the limits of integration, and therefore the inte-
grated amplitude along a y coordinate, are dependent on its x-value.
From (6–54) above, it is seen that the effective illumination along the x-axis is math-
ematically tapered toward each aperture edge. In other words, the bracketed term of
(6–53a), that is, equivalent linear aperture distribution, is (1 − x2)1/2. Thus, (6–53a)
becomes
2π
x sin θ
(1 − x )
a2 j
2 12
g (θ ) = ∫ e λ dx (6–55)
−a 2
Consistent with the original assumption, g(q) of (6–55) is the equation for a principal
plane pattern of a uniformly illuminated circular aperture. The beamwidth and maximum
side-lobe level for this aperture distribution are included in Table 6-2 (distribution:
f(r) = 1) of the next section. Parenthetically, by using variable instead of fixed integration
limits when determining an equivalent linear aperture distribution, the above example
shows that g(q) can be determined using (6–53a) or (6–53b) for planar apertures that are
not rectangular.
The determination of equivalent line-source apertures for nonseparable distributions
is usually more difficult than in the above example. However, the concept of equivalent
linear apertures helps to show how aperture amplitude shaping of planar apertures affects
side lobes and beamwidths, similar to how amplitude tapering affects the pattern of a
linear aperture. Relationships between the physical shaping of a planar aperture and the
tapering of an effective linear aperture are discussed also in connection with Fig. 6–6.
TABLE 6-2 Beamwidth, Maximum Side lobes, and Relative Gain Versus Circular
Aperture Distribution (from Silver 1949, p. 194)
E1(u ) =
sin u
u
(6–56)
πa
with u = sin θ for a pattern in the plane of the line source.
λ
The cosine-squared distribution along the y-axis is unity at the aperture center and
zero at each end, and the phase is assumed to be zero. Then, A2( y) = cos2(p y/b), with
−b/2 ≥ y ≤ b/2; and from Johnson (1993, p. 2–16), the normalized pattern for a line-source
with a cosine-squared distribution is
sin u π 2
E 2 (u ) =
u π 2 − u2
(6–57)
πb
with u = sin θ for a pattern in the plane of the line source.
λ
To obtain the pattern E(q, f) for a planar aperture having line source distributions in
accordance with (6–56) and (6–57),
and
E (θ , φ ) = E (U x , U y ) = E1(U x )⋅ E 2 (U y )
(6–58)
shapes of the aperture distributions. Consequently, the side lobes for f = 45° patterns are
not always smaller than those of both the f = 0° or f = 90° patterns. However, for some
operational scenarios, the effect of a side-lobe reduction in a plane, without loss of gain,
sometimes may be obtained by rotating the aperture 45° about its z-axis.
References
Adams, R. J., and K. S. Kelleher, “Pattern Calculation for Antennas of Elliptical Aperture,”
Proceedings IRE, vol. 38, September 1950, pp. 1052 ff. See also K. S. Kelleher,
“Reflector Antennas,” ch. 12 in Antenna Engineering Handbook, H. Jasik (ed.),
McGraw-Hill, 1961.
Allen, C. C., “Geodesic Lens With Organ Pipe Feed,” Record of the Georgia Tech-SCEL
Symposium on Scanning Antennas,” Georgia Institute of Technology, 18–19 December
1956, pp. 297–314.
Balanis, C. A., Antenna Theory: Analysis and Design, 3rd ed., John Wiley & Sons, 2005,
ch. 13.
Bodnar, D. G., “Lens Antennas,” ch. 18 in Antenna Engineering Handbook, 4th ed., J. L.
Volakis (ed.), McGraw-Hill, 2007.
Clayton, L., Jr., and J. S. Hollis, “Calculation of Microwave Antenna Radiation Systems
By the Fourier Integral Method,” Microwave Journal, Sept. 1960, pp. 59–66.
Diaz, L., and T. Milligan, Antenna Engineering Using Physical Optics, Artech House,
1996.
Dyer, F. B., and R. M. Goodman, Jr., “Vehicle Mounted Millimeter Wave Radar,” 18th
Tri-Service Radar Symposium Record, University of Michigan, 1972. Reprinted in
S. L. Johnston, Millimeter Wave Radar, Artech House, pp. 485–506, ch. 19.
Emrick, R., and J. L. Volakis, “Millimeter-Wave and Terahertz Antennas,” ch. 23 in
Antenna Engineering Handbook, 4th ed., J. L. Volakis (ed.), McGraw-Hill, 2007.
Goodman, R. M., Jr., R. C. Johnson, H. A. Ecker, and W. K. Rivers, Jr., “A Folded
Geodesic Luneberg Lens Antenna,” presented at the 17th Annual Symposium on USAF
Antenna Research and Development at the University of Illinois, November 1967.
Hannan, P. W., “Microwave Antennas Derived from the Cassegrain Telescope,” IRE
Transactions on Antennas and Propagation, March 1961, 140–53.
Hollis, J. S., “A Fourier Integral Computer for Calculation of Antenna Radiation Patterns,”
MSEE Thesis, Georgia Institute of Technology, 1956.
Hollis, J. S., and M. W. Long, “Luneberg Lens Scanning System,” IRE Transactions on
Antennas and Propagation, vol. AP-5, January 1957, pp. 21–25.
Imbriale, W. A., and D. L. Jones, “Radio-Telescope Antennas,” ch. 49 in Antenna
Engineering Handbook, 4th ed., J. L. Volakis (ed.), McGraw-Hill, 2007.
Johnson, R. C., “The Geodesic Luneberg Lens,” Microwave Journal, vol. 5, August 1962,
pp. 76–85.
Johnson, R. C., Antenna Engineering Handbook, 3rd ed., McGraw-Hill, 1993.
Johnson, R. C., and R. M. Goodman, Jr., “Geodesic Lenses for Radar Antennas,” Record
of IEEE EASCON ’68, Washington, DC, 1968, pp. 64–69.
Kock, W. E., “Metal Plate Antennas,” Proceedings of the IRE, November 1946, pp.
828–36.
Long, M. W., W. K. Rivers, Jr., and J. C. Butterworth, “Combat Surveillance Radar
AN/MPS-29 (XE-1),” Record of the Sixth Annual Radar Symposium, University of
Michigan, 1960. Reprinted in S. L. Johnston, Millimeter Wave Radar, Artech House,
pp. 461–78, 1980.
Problems and Exercises 283
Chapter 7 discusses properties of antennas and antenna analysis techniques not addressed
in other chapters. Antennas are used in many different applications, and consequently
there are numerous special properties needed to satisfy their requirements. The range of
topics addressed in this chapter is wide, and includes techniques for providing wide
bandwidths, multiple polarizations, low receiver noise, and extremely low side lobes. A
specific application may, of course, employ more than one of these techniques. There is
also a section solely on antennas for direction finding and another on antennas that
accomplish beam scanning mechanically. The concept of synthetic-aperture antennas,
although perhaps more of a signal processing than an antenna topic, is also discussed
here, because it is a uniquely effective way of attaining very narrow antenna beamwidths
through use of a relatively small antenna on a moving platform. Finally, the chapter closes
with a section on seemingly unrelated subjects, namely, geometrical theory of diffraction
(GTD), method of moments (MoM), and fractals. These topics are, however, connected
by the subject of radiation from sharp edges (curvatures small compared to a wavelength),
which currently cannot be analyzed by exact theoretical methods.
GTD and MoM analyses are computer programmed for solutions to complex antenna
configurations. GTD uses ray-optics representation of electromagnetic propagation, and
it also incorporates diffraction theory and surface waves for surfaces large compared to
a wavelength. MoM, on the other hand, is applicable primarily to wire antennas (e.g.,
dipoles and longer wire radiators), but it has also been used for patch antennas. MoM
programs compute currents, from which radiation can be calculated. In contrast, fractals
are computer-generated random shapes proven useful as novel antenna configurations,
and for which MoM programs are used to analyze the radiation from their sharp edges.
285
286 Antennas with Special Properties
better ratio of radiation resistance to reactance. In effect, the radiative loading lowers the
Q. Moreover, other methods of improving bandwidth have been discovered. These
methods may be separated into two classes—those that apply to basic radiators, and those
that apply to complex antennas, such as arrays, reflector-feed combinations, and lens
antennas (in which the refracting element as well as the feed may be frequency-sensitive).
This separation is not absolute, but it is of some value.
The methods that apply to basic radiators, such as dipoles, are mostly geometric in
nature. That is, the employment of certain shapes, and certain size ratios, have been found
to be beneficial. A dramatic extension of this class of methods has been made in the
development of the spiral and log-periodic antennas, or frequency-independent antennas
as they are also called.
Other methods apply to complex antennas. In some cases it is possible to combine
elements whose variations of impedance (or other properties) are complementary in such
a way that they tend to offset each other, as the frequency is varied. The bandwidth of
an array is, naturally, improved by employing broad-band radiators as the elements. It is
important to note, however, that the element bandwidth in the array environment may
be less than it would be when isolated, because of mutual coupling effects.
Another important factor in broad-band array design is the transmission-line or wave-
guide feed system that distributes the rf power of the transmitter, to provide the required
amplitude and phase to each array element. Certain possible arrangements are much less
frequency-sensitive than others. The two basic methods are shown in Fig. 2–8, chapter
2. For broadside arrays, the method of Fig. 2–8a is much better than that of Fig. 2–8b,
because it preserves the same total length of line from the common input point to each
element of the array; therefore all elements will be in-phase regardless of the frequency.
But with the arrangement of Fig. 2–8b, the elements will be in-phase only at frequencies
for which the interconnecting line length, d, is an integral multiple of a wavelength. The
arrangement with all input-to-element path lengths equal is called a corporate structure
feed system.
It is often necessary to provide impedance-matching transformers at an antenna’s input
terminals or at the elements of an array. Baluns (sec. 4.11) may also be required. The
bandwidth of the transformers and baluns is an important factor in determining the overall
impedance bandwidth. It is desirable to employ a transformer and balun whose reactance
variations with frequency are in such a direction that they tend to cancel the reactance
variation of the antenna or the array element. For example, a dipole fed at the center has
inductive reactance at frequencies above the resonant frequency, and capacitive reactance
at lower frequencies. A possible method of adjusting the dipole input impedance is to
make it shorter than the resonant length, so that it is capacitive; then a short-circuited
section of transmission line (shorted stub) is connected across the antenna input termi-
nals, and its length is adjusted until it has an inductive reactance just sufficient to coun-
teract the capacitive reactance of the dipole. For example, the required stub length is
obtainable by rearranging equation (2–36), chapter 2. The resulting input resistance will
have different values for different amounts of shortening of the dipole from the resonant
length. The reactances of the stub and of the dipole will tend to cancel as the frequency
is varied up or down. The cancellation will be imperfect, but there will be less net
288 Antennas with Special Properties
reactance change with frequency than there would be for a resonant dipole without the
stub. This is a relatively simple illustration of a principle. More elaborate and effective
methods may be used in practice.
These observations serve mainly to convey some of the general ideas and a few par-
ticulars of the art of making complex antennas broad-band. The remainder of this section
is devoted to a discussion of the design of individual radiators of broad bandwidth.
FIGURE 7–2.
7.1.4. Monopoles
Dipoles, whether cylindrical, biconical, or of other shape, are symmetrical radiators when
fed at their centers. All dipole types can also be used as monopoles, which were discussed
for specific dipole forms in secs. 4.1 and 4.3. A monopole is half of a dipole operated in
conjunction with its image in a conducting ground plane perpendicular to it. Monopole
designs corresponding to the various broad-band dipoles that have been discussed are
shown in Fig. 7–5. There are innumerable variants of the shapes shown. The feed-point
(input) impedance of a monopole is exactly half the value of the corresponding dipole,
if the ground plane is a perfect conductor
and of infinite extent, and the pattern cor-
responds exactly to half the pattern of
the dipole. Since for a given input power
the radiation is concentrated into half the
q solid angle occupied by the dipole pattern,
Transmission the directive gain of the monopole is
line twice that of the dipole in free space. (See
h
discussion of directivity in sec. 3.3 and of
vertical radiators in secs. 4.1 and 4.3.) In
s
practice, of course, the ground plane will
not be a perfect conductor of infinite
extent, but the properties of the monopole
will approximate those just described if
the ground plane is a good conductor
FIGURE 7–4. extending to a distance several times as
Biconical antenna. great as the height of the monopole.
Ground
plane
Monopoles are especially useful as vehicle antennas where the ground plane is the
“skin” of the vehicle (ship, boat, automobile, tank, aircraft, spacecraft). On aircraft the
monopole may for aerodynamic reasons take the shape of a streamlined “blade” rather
than the circular-cross-section shapes commonly used in other applications. Another
advantage of the monopole is that it is conveniently fed by a coaxial line. (The advantages
of coaxial lines are discussed in sec. 2.4.) A monopole has the same bandwidth properties
as the corresponding dipole.
l /2
l /4
l /4
(Flat
sheets)
(a) (b) (c)
l /4
l /4 (Disc size
0. 3l variable)
∼30°
0.3 l
l /3
Coaxial
feed line
Coaxial Coaxial
feed line feed line
(d) (e) (f)
FIGURE 7–10.
Miscellaneous dipoles and related types. (a) Triangular dipole. (b) Fan dipole. (c) Ellipsoidal
dipole. (d) Folded monopole with simulated ground plane. (e) Conical skirt monopole.
(f) Discone antenna.
comparison with more conventional antennas. They may have unidirectional or bidirec-
tional patterns of low to moderate directive gain. Higher directive gain may be obtained
by using them as elements of an array or as feeds for parabolic reflectors and lenses, in
which use, however, a design must be chosen that maintains a constant or nearly constant
phase center. (Many log-periodic types do not.) Log-periodic array design presents
special problems not encountered in ordinary arrays. Excellent discussions of the subject
have been published by Deschamps and DuHamel (1961), Jordan, Deschamps, Dyson,
and Mayes (1964), and Rumsey (1966).
Beam direction
R6
R5
l6
l5
Transmission
line
Dipoles
FIGURE 7–11.
Log-periodic dipole array.
perhaps the nearest that a log-periodic design can come to resembling a conventional
antenna. (There is a marked similarity to the end-fire dipole array described in
sec. 5.2.)
This particular form of dipole array was devised by Isbell (1960). It consists of a
number of dipoles of different length and spacing, fed by a two-wire line that is trans-
posed between each adjacent pair of dipoles. The array is fed at the small end of the
structure, and the maximum radiation is toward this end. The lengths of the dipoles and
their spacing are graduated in such a way that certain dimensions of adjacent elements
bear a constant ratio to each other; these are the quantities designated R and l in Fig. 7–11.
That is, if this design ratio is designated by t (a number less than one), then
R2 R3 R4 1 l2 l3 l4
= = = = = = (7–1)
R1 R2 R3 τ l1 l2 l3
It is apparent that these conditions cause the ends of the dipoles to lie along straight lines
that meet at an angle designated a. It is a characteristic of the frequency-independent
antennas that certain aspects of their structures can be specified in terms of an angle or
angles.
The result of these structural conditions is that if a plot is made of the input impedance
as a function of frequency, the variation will be found to be repetitive. If the plot is made
against the logarithm of the frequency, rather than the frequency itself, this variation will
be periodic—that is, the impedance will go through cycles of variation in such a way
that each cycle is exactly like the preceding one. A typical plot is shown in Fig. 7–12. It
is this behavior that gives rise to the name “log periodic”; the impedance is a logarithmi-
cally periodic function of the frequency. (It is not implied, however, that the variation is
necessarily sinusoidal.) Moreover, all the electrical properties of the antenna undergo a
296 Antennas with Special Properties
* In certain cases the impedance (but not the pattern) may vary with a period half as large (that is, with
1 1
period 2 log (1/t)). But these cases do not violate Eq. 7–2, since a variation with periodicity 2 log (1/t) is
also periodic with period log (1/t).
Frequency-Independent Antennas 297
be of properly increasing thickness as the distance from the apex of the structure increases.
The width of the gaps at the dipole centers must also increase. However, these relatively
minor factors can be ignored without serious effect if the overall bandwidth (ratio of
maximum to minimum cutoff frequencies) does not exceed about 10 : 1.
When a log-periodic antenna is fed at a particular frequency, it is found that the radia-
tion occurs from a certain portion of the structure, and that other portions do not radiate.
This “active” region of the antenna of Fig. 7–11 is the region in which one or more of
the dipoles is nearly half a wavelength. The cutoff frequencies are those at which the
shortest and longest dipoles are approximately half a wavelength. Thus the active region
of the antenna is near the apex for the highest frequencies radiated, near the large end
for the lowest frequencies, and at intermediate positions for frequencies in between. This
behavior may also be expressed by saying that the phase center of the antenna progresses
from the large end to the small end as the frequency goes from minimum to maximum
of the band-width range.
Log-periodic antennas may be either unidirectional or bidirectional in their pattern
characteristics. The bidirectional types are characterized by two active regions that move
apart as the frequency is decreased; in these types the phase center remains fixed and
coincides with the feed point.
For the antenna of Fig. 7–11 typical values of the design ratio and apex angle are
t = 0.8 and a = 30°. For any given value of a, there is a minimum permissible value of
t (which is always a number between 0 and 1). Larger values of a and smaller values
of t go together and result in more compact design for a given bandwidth. On the other
hand, smaller values of a and larger values of t result in improved performance—smaller
variation of impedance and pattern, and higher gain, at the cost of a larger structure. The
reason for this effect may be intuitively understood by realizing that for small a and large
t, at any given frequency, there are more dipoles of nearly half-wavelength; hence the
active or radiating region of the antenna encompasses more dipoles than for large a and
smaller t. As is always true, the gain increases with the number of radiating elements
and with the size of the radiating region. Also, the impedance and pattern variations are
smaller because of the smoother transition in the configuration of the radiating region
when it encompasses several dipole elements.
Sheet
metal
Feed
line
a
(a)
Support
Beam
direction
g
Wires
Feed
line Support
(b)
Beam direction
q Wires
Feed
line
(c)
FIGURE 7–13.
Some practical log-periodic antenna designs. (a) Trapezoidal tooth structure. (b) Wire
trapezoidal tooth antenna. (c) Wire triangular tooth antenna.
Frequency-Independent Antennas 299
spiral makes with a circumference, or its complement that the spiral makes with a radial
line from the apex. There are two interleaved spiral conductors lying on a conical (non-
conducting) surface. Each spiral is connected at the apex to one side of the feed line.
The radiation is unidirectional, going in the direction toward which the apex of the cone
points. The peak of the beam is polarized elliptically, becoming virtually circularly polar-
ized as the number of turns per unit length is increased.
¼-wavelength loop
of line (from AA to BB)
Dipoles
B
A
Two-wire
transmission
line
Line
input
FIGURE 7–17.
Arrangement of crossed dipoles for radiation of circular polarization (turnstile antenna).
the radiation will have varying degree of ellipticity, going from circular on the perpen-
dicular axis to linear polarization in the plane of the dipoles.
This arrangement of dipoles and the feed system is shown in Fig. 7–17. It is an espe-
cially important arrangement for two reasons. First, it is a basic method of producing
circular polarization. The pattern in this polarization is bidirectional but can readily be
made unidirectional by placing a reflector behind the dipoles, or combining several pairs
in an endfire array (with the cross-dipole planes perpendicular to the line of the array).
Second, it is useful when approximately uniform radiation is desired in a horizontal plane
with horizontal polarization.
This configuration is known as a turnstile antenna. To provide increased gain in the
horizontal plane and to reduce or eliminate the circularly polarized upward and downward
radiation, several turnstiles can be stacked vertically and fed in broadside-array fashion.
This arrangement is much used for television and FM broadcasting, using broad-band
dipoles. When a particular form of slotted-sheet dipole is used, the arrangement is known
as the superturnstile antenna, as was mentioned in the discussion of broad-band dipoles
(see Fig. 7–9). The quarter-wave phase-delay line section, if this phasing method is used,
304 Antennas with Special Properties
should be arranged to produce as little obstruction as possible of the radiated fields and
also to have as little capacitive or inductive coupling to the dipoles as possible.
It is also possible to produce the necessary 90-degree phase difference by separately
feeding each dipole from a common source with parallel lines whose lengths differ by
a quarter wavelength, provided that both lines are matched by the dipole impedances
(no standing waves). The necessary phase shift between the dipoles can also be obtained
with a coaxial- or waveguide-type of phase shifter, and at low enough frequencies even
lumped-circuit elements (coils and condensers) may be used. Care must be taken, in
designing the feed system, to assure that the impedance relationships are such as to feed
equal power to each dipole and to provide a proper impedance match to the main feed
line.
Circular polarization can also be obtained with a horn radiator, either with a square
waveguide and a square-aperture pyramidal horn, or a circular waveguide and a conical
horn. In either case, the requirement is to excite a circularly polarized mode of propaga-
tion in the waveguide. This can be done in a variety of ways. A common way is to
introduce both polarizations into a rectangular waveguide, by means of in-phase probes
(Fig. 2–15, chapter 2) in two adjacent walls. Because of the rectangularity of the guide,
the two mutually perpendicular polarizations will have different phase velocities (this
assumes the waveguide is not square, but its height and width must be large enough to
permit propagation of perpendicular polarizations). The length of the rectangular guide
(measured from the probe position) is made just long enough to result in a 90-degree
phase difference between the two polarizations at the guide output.
A horn or other radiator with circularly polarized radiation can be used to illuminate
a paraboloidal reflector to produce a high-gain circularly polarized beam. Pyramidal
horns that are not square can be used to illuminate noncircular apertures, but special
methods must be used to compensate for the differences in phase velocities of the two
polarizations in the horn flare (since the flares are different for the different polarizations).
Lenses can also be used with such horns if the refracting medium is capable of passing
both polarizations with the same phase velocities.
A linearly polarized wave can be directly converted to circular polarization by means
of special transmission or reflection devices. An example of such a transmission medium
is a set of straight parallel vanes or plates with their planes parallel to the direction of
propagation and their direction inclined at 45 degrees to the linear polarization, so that
the wave in the regions between the plates can be considered to have two mutually per-
pendicular components, one parallel to the plates and one perpendicular to them. The
latter component will travel through the region at free-space velocity, but the former will
have a phase velocity that is determined by the plate spacing, according to the principle
of the metal-plate waveguide lens (except that the boundary surfaces of the parallel-plate
region are flat, or plane, instead of curved). If the thickness of this “lens” is properly
chosen, the two component waves can be caused to have a 90-degree phase difference
and the wave that emerges will be circularly polarized. This device is known as a quarter
wave plate, since it produces a quarter-wavelength effective path difference for the two
polarizations. A reflecting device that accomplishes the same result can be constructed
by using a similar set of vanes of half the depth, backed by a plane reflector.
Antennas for Multiple Polarizations 305
• waveguide dimensions are small enough so that only one waveguide mode
(the dominant mode) is propagated
• a thin dielectric slab, as depicted in Fig. 7–18, is aligned at +45° to vertical
• the length of the slab is long enough that the phase delay of the +45° polarized
wave is 90° greater than that of the −45° wave
306 Antennas with Special Properties
This situation occurs most often at low frequencies, where an antenna of wavelength
dimensions is extremely large. Then it becomes important to consider antennas that are
physically and electrically small. The emphasis of the discussion is on their electrical
smallness, since mere physical smallness does not by itself make an antenna unusual. An
electrically small antenna is one that is small compared to the wavelength.
Some basic forms of electrically small antennas have already been discussed in secs.
4.1 and 4.5—the short dipole and the small loop. Their inefficiency was noted, particu-
larly in connection with the use of vertical grounded antennas for transmitting at LF and
VLF. These are actually monopoles or half dipoles operated in conjunction with their
ground-reflection images. They may sometimes be one hundred or more meters high, yet
they are electrically small because the wavelength may be several kilometers. Their inef-
ficiency is due to a combination of factors. The radiation resistance is very low, so that
a high current is required. At the same time the input impedance includes a high capaci-
tive reactance. If the required high current is to be obtained with a reasonable generator
(transmitter) rf voltage, this capacitive reactance must be tuned out with a large series
inductance (coil). The coil, in turn, has some resistance, in which appreciable loss can
occur due to the heavy current. The heavy current also flows into the ground so that
further losses will occur unless the ground conduction is excellent.
There are various means of reducing these losses. Capacitive top loading is used to
increase the radiation resistance, thus lowering the current required for a given radiated
power. Extensive systems of buried wires are used to increase the effective ground con-
ductivity. Loading coils (tuning inductors) are made of large conductors to reduce resis-
tance. These measures are necessary when the electrically short, grounded antenna is
used for transmitting. Consequently, at higher frequencies where efficient antennas may
be of reasonable size, electrically small antennas are not generally used in transmitting
applications.
There are times, at the low and medium frequencies especially, where electrically
small antennas may be necessary for receiving. Moreover, it is fortuitous that in receiving
applications at these frequencies the antenna efficiency may be unimportant, so that the
electrically small antenna may function very effectively. This may seem to be a contra-
diction of the “reciprocity” that has been ascribed to antennas in their transmitting and
receiving behavior, but it really is not. The reciprocity principle is correct and does hold
for electrically small antennas as long as only the radiated and received signal is consid-
ered. In reception, however, the important consideration is the signal-to-noise ratio.
The electrically small antenna of low efficiency may be effective when the alternative
is a larger antenna of about the same directivity, and when the predominant noise is
external, that is, enters the receiver via the antenna. These circumstances occur primarily
at frequencies below about 10 MHz. The predominant noise is atmospheric static, and
high-directivity antennas are usually out of the question, especially below 2 or 3 MHz.
Thus the alternative antennas may be a tall tower, possibly top-loaded, or a very much
shorter vertical monopole. The patterns, and hence the directivities, will be practically
the same in either case. The difference being considered between antennas is efficiency
(the factor k defined in sec. 3.3.5) and thus gain, and it can be shown that, within wide
limits, this will not affect the signal-to-noise ratio under the stated conditions.
310 Antennas with Special Properties
Very common examples of electrically small antennas are those used for reception in
the broadcast band—535 to 1,605 kHz (approximately 200 to 600 meters wavelength).
As is well known, a short random length of wire serves very well as an antenna for local
reception, and at night even distant reception is possible. Automobile radios typically use
a short vertical monopole, about one meter in length, or about 0.01 wavelength or less.
Small “table radios” and also some larger sets often have built-in small-loop antennas
that are resonated by means of a parallel variable condenser; the combination then serves
as the tuned input circuit of the receiver rf amplifier. The ultimate in smallness is the tiny
“ferrite rod” antenna used in transistor pocket portable radios. This antenna is actually a
small-diameter elongated coil wound on a rod of ferrite material up to perhaps 10 cm in
length, connected in parallel with the rf amplifier tuning capacitor; it thus serves as a
combination antenna and tuned circuit as does the somewhat larger loop antenna of the
table radio. The ferrite core increases the rf flux in the coil. It also allows the necessary
inductance to be obtained with relatively few turns of wire so that the resistance is kept
low; hence the coil Q is high. Consequently both the induced voltage and the selectivity
of the circuit are greater than with an air-wound coil of the same size.
Electrically small loop antennas are also very commonly used for receiving at MF,
LF, and VLF, again because efficiency is not important. Also, since the wave polarization
is always vertical at the low frequencies (below 2 MHz), advantage may be taken of the
nulls that occur in the directions perpendicular to the plane of the loop, to eliminate
interfering signals from a particular direction, if the loop is made rotatable.
Sometimes an electrically small antenna may be desirable for transmitting at frequen-
cies where a quarter-wavelength vertical antenna would be possible but fairly expensive
and perhaps inconveniently large. A “short monopole” antenna will, in this case, have
just as good directivity, but the radiation resistance will be very low so that the efficiency
may be poor. A method of improving the efficiency, which has been incorporated in a
number of commercial designs, is the use of a resonant top-loading structure. In effect,
the antenna consists of a short vertical section that radiates (with vertical polarization)
and a horizontal section that is nonradiating (because of cancellation due to ground
reflection) but is of sufficient length, inductance, and capacitance to make the entire
structure resonant. That is, the input impedance is nonreactive and the radiation resistance
is much higher than it would be for a monopole without top loading. A metallic ground
plane is also employed. Therefore such antennas may be reasonably efficient even though
electrically small.
will be attempted here is a brief sketch of the antenna types and techniques, since the
basic operation of the antennas has already been covered in chapters 4, 5, and 6.
locating them at a distance from the receiver, with transmission-line connections. Crossed
Adcock antennas can also be used with a goniometer.
difference signal, since the amplitudes of the two lobes are equal. If the signal is coming
from B (an “off-axis” direction), there will be a difference, or error, signal, as shown by
the fact that the dashed line crosses the two lobe patterns at different levels. The sense
(direction) of the error is found by comparison of the phases of the sum and difference
signals in special receiver circuits that convert the error signal into a d-c voltage; the
polarity of this voltage will reverse as the signal direction changes from right-of-center
to left-of-center. The sum signal is also used, in radar systems, for signal detection and
range measurement. In the sequential-lobing scheme a single receiver is used, and the
received signals from the two beams are switched so that their amplitudes can be
compared.
It is apparent that the “difference pattern” of the two lobes (the difference of the pattern
amplitudes plotted as a function of angle) will be very sharp, rather than blunt like the
nose of a lobe. Because of this fact the accuracy possible by lobing is much greater than
by simply observing the maximum direction of a beam.
The two displaced beams can be generated in a number of ways. With an array antenna,
changing the phasing of the elements will produce the desired result. With a paraboloidal
reflector or lens antenna, the beam may be shifted in the desired manner by moving the
feed slightly off the optical axis, as shown in Fig. 6–3b, chapter 6. In a sequential lobing
system this shift may be made by mechanically moving the feed off-axis first in one
direction, then in the opposite direction, at a rapid rate. In either a sequential or a simul-
taneous lobing system, two feeds may be used, each displaced a small amount from the
optical axis, in opposite directions. In the simultaneous lobing system both feeds are
connected to the receiver through a hybrid junction (coupler) (sec. 2.6) so that a differ-
ence (error) signal is received. For sequential lobing, the receiver is switched between
the lobes.
Thus far, lobing only in one plane has been discussed. It may be desired, however, to
measure the incoming signal direction in two mutually perpendicular planes, for example,
vertically and horizontally. This can be accomplished by generating four lobes—left,
right, up, and down with respect to the nominal beam axis. With a paraboloidal reflector
or a rotational lens this requires four fixed feeds. A typical arrangement of four horns is
shown in Fig. 7–23. Their waveguide outputs are combined by means of hybrid junctions
in such a way that one receiver is fed
the difference between the signals of the
upper and lower pairs, and another the Four-horn
difference between the left and right pairs. feed cluster
Paraboloidal
A third receiver is fed the sum of all four Zero axis of
four-lobe
Paraboloid reflector
axis
horns. pattern
The sequential-lobing equivalent of the
four-lobe simultaneous system is accom- Four waveguides
plished by moving a feed (of a reflector To hybrid
or lens) in a circle around the zero axis. junctions
with small microwave antennas, where the required mechanical motion is not very
great.
The simultaneous-lobe-comparison method, when used in a pulse radar system, is
known as monopulse lobing, since the signal amplitudes can be compared on a single-
pulse basis; the sequential-lobing system requires at least two successive pulses for its
operation. The method that has been described is also called amplitude monopulse, since
it is based on comparing the signal amplitudes in two angularly displaced lobes of a
single antenna. Another method, phase monopulse, uses two or four antennas placed side
by side, with their beams pointing in the same direction but with their phase centers
separated. If a signal arrives from an off-axis direction it will be received with equal
amplitude in each beam but with a phase difference, which may be used to generate an
error signal. When this phase method is used with nonpulsed systems it is referred to as
phase-comparison lobing or as an interferometer system. It is capable of greater accuracy
with a given antenna size because the sharpness of the error response is determined by
the separation of the antennas rather than by their individual beamwidths. The amplitude-
comparison lobing system, on the other hand, is more compact and more readily steerable
as a unit, for the same antenna gain and beamwidth.
In the phase-comparison or interferometer system, the interference pattern between
the two beams becomes multilobed as the separation of the antennas is increased. The
principle is identical to that of the interference between an antenna and its image in a
reflecting surface, as discussed in sec. 1.4 (see also Figs. 1–14 and 1–15). Therefore
ambiguity can exist if the individual beamwidths are appreciably greater than the lobe
widths, but the ambiguity can usually be resolved by one of a number of methods. This
technique is used in VHF satellite tracking systems. Interferometers are also commonly
used in radio astronomy applications.
Aircraft landing systems use a technique similar to the amplitude lobing principle,
where antenna patterns are transmitted from fixed ground stations and received by aircraft
receivers. For lateral guidance, ILS (Instrument Low-Approach System) radiates, from
the far end of a runway, two identical beams in the high-VHF region that are separated
in azimuth, but modulated at different frequencies. When the signal strengths of the dif-
ferently modulated VHF signals are equal, the aircraft is on lateral course. For vertical
guidance, ILS radiates, from the near end of a runway, two identical beams in the low
UHF region that are separated in elevation, but modulated at different frequencies. When
the UHF received signals of the differently moduated signals are equal, the aircraft is on
the vertical (i.e., proper glide) course.
Lobing is also used for passive direction finding at high frequencies, where it is supe-
rior to the loop or Adcock antenna methods.
angle sector from the horizon to the zenith (90°). Sometimes a more limited sector can
be scanned in repetitive fashion, an extreme example being the conical scan lobing
system, which was described in sec. 7.6. Scanning in the horizontal plane is known as
azimuth scanning; vertical-plane scanning is called elevation scanning. The two motions
may sometimes be combined when it is desired to scan a large solid angle with a narrow
beam; typically, the azimuth scan is performed at a relatively slow rate and the vertical
scan is a rapid saw-tooth motion, with at least one vertical sawtooth scan occurring during
the time it takes for the azimuth scan to progress one beamwidth. (This insures that there
will be no gaps in the scan coverage.)
Relatively slow azimuth scanning is commonly accomplished by simply rotating the
entire antenna about a vertical axis at a constant rate. Typical speeds for radar antennas
are 1 to 30 rpm, depending on the antenna size. Limited sectors can be scanned with the
antenna as a whole remaining stationary, either by moving the feed of a reflector or lens,
or by varying the phasing of an array. These beam steering methods permit higher scan-
ning speeds. In fact, for many applications inertial-less scanning by electronic phasing
of an array is the preferred method for rapid beam scanning and/or tracking. Conse-
quently, chapter 8 is devoted entirely to the concepts of electronically scanned antennas
(ESA).
14 ft
OFFSET PARABOLIC-
CYLINDER REFLECTOR
9 ft
PRIMARY FEED
32˝ 30˝
PRIMARY FEED
MOUNTING ANGLES
flicker (that occurs with slow scan rate radars). The back-to-back reflectors rotated about
a vertical axis. Advantages of multiple reflectors include increased target revisit rate for
a given antenna spindle rotation rate, and increased beam-time on target for a given target
revisit rate. An azimuth sector of about 110 or 80 degrees was viewed by a radar, depend-
ing whether the scanner had three or four reflectors.
Figure 7–24 shows the principal components of the MILORD antenna (Holliman and
Hollis, 1956). MILORD is an acronym formed from the words Mine Impact LOcating
Radar Device.
Mechanical Scan Antennas 317
INPUT WAVEGUIDE
PRIMARY
FEED
PRIMARY FEED
PHASE FRONT
LATUS
α
α RECTUM
REFLECTOR
FOCAL LINE
2α
VERTICAL
REFLECTOR PHASE FRONT
REFLECTOR
FIGURE 7–25.
Exaggerated cross section through MILORD offset parabolic cylinder reflector and primary
feed. Reflector tilt angle a is 2°. From Holliman and Hollis (1956).
upward to permit mounting the lenses beneath the reflector. The two beams are separated
in elevation by means of line source feed horns that are located on opposite sides of the
focal line of the reflector. A sandwich-type radome covers the line source feeds.
The scanning mechanism is located in the cavity formed between the two lenses. Four
sectoral lens feed horns are spaced at intervals of 90° on two feed wheels, one for the
upper- and the other for the lower-elevation beam. The wheels are mounted coaxially
and the lens feed horns of one wheel are spaced 45° in azimuth from those of the other.
The feed-switching mechanism includes a four-way switch at the center of each feed
Mechanical Scan Antennas 319
TRANSMITTED
ENERGY
2 1
6 5
that is expressed in terms of its ratio to the noise level. Consequently, it is important to
keep the noise level as low as possible.
Noise in radio systems is of various types and origins in different parts of the spectrum.
In the frequency region below about 10 MHz the principal noise is usually “static,” which
is generated in the atmosphere by lightning discharges and propagates for great distances
by ionospheric reflection. It enters the receiver via the antenna along with the desired
signal, and little or nothing can be done to minimize it. (A high-gain antenna will help
by increasing the received signal while the received noise remains virtually unchanged.
But at the very lowest frequencies, directive antennas are not feasible.)
In some situations, especially in the HF through UHF regions, man-made noise
from electrical machinery, appliances, power lines, and automotive ignition is the prime
offender. Much of this noise tends to be vertically polarized and to be stronger closer to
the ground. For the first of these reasons, horizontal polarization may be advantageous,
and is therefore used for FM and TV broadcasting. It is also beneficial, in these frequency
regions, to employ an elevated antenna connected to the receiver via a balanced or coaxial
transmission line, so that all reception takes place well above the ground.
From about 30 to 300 MHz, in the absence of man-made noise, cosmic noise predomi-
nates. This noise originates in the galactic system and is heard in audio-output receivers
as a “hiss” when the receiver gain is sufficiently high; on a radar A-scope it is called
“grass,” and on an analog TV screen it is sometimes referred to as “snow.” These terms
describe its approximate appearance. It is not seen or heard when a strong signal is
received because the receiver gain is then reduced (usually automatically). Cosmic noise
is, like static, unavoidable; it is received via the antenna along with the signal. (A high-
gain antenna is beneficial, however, because it increases the signal strength without
increasing the noise. This is true because the signal comes from a single direction whereas
the cosmic noise comes from virtually all directions of the sky.)
The cosmic noise decreases with increasing frequency, so that above about 1 GHz it
is of negligible importance. Some noise is radiated by the earth’s own atmosphere, and
by the ground. The atmospheric noise does not become significant until the frequency is
about 3 GHz or higher, and is not severe below about 10 GHz. The ground noise, on the
other hand, is relatively constant at all frequencies. The result of all these effects is that
there is a region of minimum noise between about 1 and 10 GHz, with cosmic noise and
static predominating at lower frequencies and atmospheric noise predominating at higher
frequencies. (See also sec. 9.13.2).
The foregoing discussion refers solely to noise that enters the receiver via the antenna.
Additional noise is generated within the receiver itself. Techniques exist, however, for
reducing this internal noise to such a level that the system performance is largely deter-
mined by the external noise, even in the low-external-noise frequency region. (Masers,
parametric amplifiers, and tunnel diodes and low-noise transistors are the basis of espe-
cially low noise amplifiers.) When the predominant noise is the noise that enters from
external sources, via the antenna, it is desirable to consider antenna designs that reduce
the reception of noise relative to the received signal reception.
In the frequency region UHF and above where the noise from the ground radiation is
a significant portion of the total noise, certain aspects of the antenna design do affect the
322 Antennas with Special Properties
system noise level. As has been mentioned in chapters 3, 5, and 6, directional antennas
always have some side lobes and back lobes, through which some reception of ground
noise can occur even when the main-beam is pointed upward. Therefore, the objective
in designing a low-noise antenna is primarily to reduce the level of these lobes as much
as possible. It is especially important to do this when the antenna will be used for such
applications as space communication, radar, or telemetry with a low-noise receiver in
the frequency range somewhat below 1 GHz to somewhat above 10 GHz.
The design techniques that minimize side and back lobes have been discussed in sec.
5.6 and 6.3. There is, however, one further factor to be considered; namely, the ohmic
losses between the waves in space and the input terminals of the first amplifier of the
receiver. Because noise will be generated by any loss that exists in this part of the system
the length of transmission line or waveguide must be kept to a minimum. For this reason,
arrays are not ordinarily used as low-noise antennas, even though their side-lobe and
back-lobe characteristics can be controlled more readily than those of reflectors. The
losses in a typical array feed system (especially if the array is large) may generate more
thermal noise than would be avoided by total elimination of the side and back lobes.
(Thermal noise generation is discussed in sec. 9.13.)
The effect of these array losses can be circumvented by placing a low-noise receiver-
preamplifier at each element of the array, ahead of the feed-system losses. The
signal phase must be preserved in the preamplifier, so that the signals from the various
elements can be combined in correct phase relationship in the usual way, to form a
beam in the desired direction. By definition (IEEE 1993), an active antenna array is an
array which all or some of the elements are equipped with their own transmitter or
receiver or both. Usually there is not a complete transmitter or receiver located at an
element. Instead, located at the elements must of course be at least those components
necessary for minimizing losses and for providing the proper array element phases.
Generally, the active array antenna technique is very expensive, and it is therefore used
only in special applications such as in electronically beam-steered arrays for radar
(chapter 8).
Lenses can be designed for low side lobes, and also they offer the possibility of locat-
ing the receiver near the feed. For very high gain applications, however, lenses become
unwieldy in the low-noise frequency region below 10 GHz. Therefore, except for elec-
tronic beam scanning radar, the greatest emphasis in low-noise antenna design is on the
use of paraboloidal reflectors.
It is well known that a highly tapered aperture illumination will reduce the side-lobe
level, and somewhat greater tapering is usual in low-noise designs than when maximum
gain is the paramount consideration. At least equally important for reflector antennas is
the minimization of illumination spillover, because spillover causes back lobes directed
toward the ground. For the same reason it is important to minimize leakage through the
openings of a mesh reflector surface; that is, the openings must be very small compared
to the wavelength.
The Cassegrain feed system (sec. 6.3) is frequently employed to minimize the trans-
mission line length, to reduce or eliminate noise due to transmission-line loss. The
Cassegrain system is also less likely to result in back lobes due to spillover. The spillover
Low-Noise Receiving Antennas 323
Collimated radiation
reflector
aperture
Horn aperture
Plane of rotation Paraboloid focus and
about a phase center of horn
horizontal axis
Typical ray paths
rf input Reflector (offset
paraboloidal
section)
Horn walls
Paraboloid
axis
Continuation of
parabolic curve
Paraboloid
vertex
FIGURE 7–29.
Sectional view of a conical-horn-reflector antenna. Metallic enclosure (heavy line) is complete
except for opening at reflector aperture.
that occurs at the subreflector causes side lobes that look at the sky rather than at the
ground; hence they result in little or no additional ground-noise. When a conventional
focal-point feed is used, if low noise is important it is desirable to locate a receiver pre-
amplifier at the apex of the feed support. Care should be taken, when this is done, to
minimize aperture blockage, which also causes higher side-lobe levels. Parenthetically,
offset-fed reflectors are commonly used to reduce feed blockage and side-lobe levels for
home-use direct-broadcast satellite TV reception.
An antenna design that is perhaps ultimate in low-noise performance is a very large,
conical-horn-reflector antenna used by the Telstar satellite-communication ground termi-
nal at Andover, Maine. Figure 7–29 is a diagram of this type of antenna. Its design dra-
matically reduces spillover and reflector leakage, and also permits the receiver and
transmitter to be located close to the feed point, at the apex of the horn. The axis of the
horn is horizontal, but the mouth of the antenna can be pointed upward. This is possible
because a section of a paraboloidal reflector is inserted in front of the mouth of the horn,
so that the wavefront is redirected at a 90-degree angle from the horn axis. The apex of
the horn is at the focus of the paraboloid. The paraboloidal aperture is circular, 20.7
meters in diameter. The antenna is used for reception at a frequency of 4.08 GHz and
provides high gain and a narrow beam in combination with extremely low noise. The
gain is approximately 58 dB, and the beamwidth is about 0.23 degree. The antenna is
324 Antennas with Special Properties
also used for transmitting at 6.39 GHz, where the gain is 62 dB and the beamwidth is
0.15 degree. The entire antenna, enclosed in a 210-foot-diameter inflatable radome, is
mounted on a turntable about 50 meters in diameter. The antenna rotates about the horn
axis, and thus the aperture can be directed toward any part of the sky. A detailed descrip-
tion of this antenna has been published (Hines, Li and Turrin 1963). A modification of
the design that results in a more compact structure has been devised. It is called the triply
folded horn reflector antenna (Giger and Turrin 1965).
There are three SAR modes of operation: stripmap, spotlight, and scan. In stripmap
SAR, the beam remains at a constant squint angle, which is usually perpendicular to the
flight. Spotlight is used to obtain an improved angular resolution of a known location of
a target of interest. As the platform moves, the beam-pointing direction is changed so as
to keep pointing at the target. Lastly, the scan mode is seldom used; but by means of
beam direction changes, it does permit observing a straight path that is not parallel to
the flight path. Even though there are different SAR modes of operation, the technique
employs side-looking from an elevated platform, and the resultant improvement in
angular resolution is one-dimensional, in the direction parallel to the vehicular motion.
The required resolution in the orthogonal (range) direction is obtainable because the
radar employs pulses. By means of pulse compression, a relatively long-duration trans-
mitted pulse can be transformed into a very short received pulse, and the range-dimension
resolution is determined by this received pulse length. Pulse compression is accomplished
by transmitting a very wide-band coded pulse, which is decoded and compressed by an
appropriate receiver filter. The pulse at the filter output is much shorter in duration than
the transmitted pulse, and the received signal sensitivity is dependent only on the trans-
mitted pulse energy; and the range resolution obtained depends only on the compressed
pulse length.
This combination of angle resolution by means of the synthetic aperture technique and
range resolution by pulse compression allows very high-resolution mapping of the surface
of the earth and other planets, monitoring of the positions of ships at sea, and study of
changes in the configuration of the earth or sea surface over time (Elachi 1987).
An interesting aspect of the SAR technique from the antenna-theory viewpoint is that
the best performance is obtained by processing the data so that, in effect, the antenna
beam is focused on the earth surface at the point of reflection of the radar pulse. Ordinary
arrays, as well as reflectors and lenses, usually produce a collimated beam, in which the
rays are parallel; in effect they are “focused at infinity.” This is the appropriate strategy
for transmission distances that are in the far-field region of the antenna. The simplest
types of SAR also produce a collimated beam. But by suitable data processing the SAR
beam can in effect be focused so that the rays converge to a point at the earth’s surface
(and this can be done at all ranges, since the data are separately processed for different
target distances). The advantage of this type of processing is that it permits a much larger
synthetic aperture to be formed than would be possible if a parallel-ray beam were
used.
For this synthetic-aperture procedure, the ordinary concept of a beamwidth is not
applicable. The relevant consideration in mapping is the azimuthal resolution, which is
expressed as a distance rather than an angle. For ordinary radar, and for far-field targets,
the resolution is of the order of qbR, where qb is the azimuthal beamwidth in radians and
R is the target range. But, for sufficiently large synthetic apertures, targets are not neces-
sarily in the far-field. It is shown by Cutrona (1990, pp. 21.4–21.7) that for the focused
case, and for a sufficiently large synthetic aperture D along the flight path, the azimuthal
resolution is equal to D/2, independent of the range and also independent of the frequency
(or wavelength) of the radar. Cutrona also shows that for the unfocused synthetic aper-
ture, the resolution is λ R 2 where l is the wavelength.
326 Antennas with Special Properties
than waveguide slots, thus demonstrating that the technology is not confined to slot
arrays.
Although it is generally considered that better control of side-lobe levels can be
achieved with array antennas than with reflectors, side-lobe levels in the 40 to 50 dB
range have been reported for reflector antennas. These results are accomplished using
computer-aided design and precision construction. In addition to his review of low-side-
lobe array technology, Schrank also published a review article on low-side-lobe reflector
antennas (Schrank, 1985).
The decibel side-lobe level is conventionally expressed as the ratio applicable to the
highest side-lobe level. In some contexts, the average side-lobe level may be of greater
importance than the level of the highest side lobe. The average side-lobe level is usually
expressed in terms of its ratio to the isotropic level, denoted in decibels by the notation
dBi. For high-gain (narrow-beam) antennas, this ratio is approximately equal to the frac-
tion of the total power radiated (in transmitting-antenna terms) in the side lobes. Thus,
−20 dBi means that approximately one percent of the total power radiated is side-lobe
radiation. Since the main-beam power gain is also referred to the isotropic level, this
means that if the main beam gain is 35 dBi, for example, and the average side-lobe level
is −5 dBi, then the average side-lobe level is 40 dB below the main beam radiation. (Note
that in the dBi notation, positive numbers refer to levels above isotropic and negative
numbers refer to dB below isotropic level.)
In addition to airborne radar applications, low side-lobe design is especially
important for reflector or lens antennas used on satellites for earth communication when
frequency reuse is an objective. A paper by Burdine and Wilkinson (1980) discusses this
matter and describes techniques for minimizing the side lobes of reflectors and lenses.
(Many of the principles discussed in that paper are also described in chapter 6 of this
book.).
As discussed in secs. 7.8 and 9.13.2, as major objective in designing a low noise
antenna is to reduce the side lobes and back lobes as much as possible. The well-known
horn-parabola, a successful technique for reducing wide-angle and rear lobes, is dis-
cussed in sec. 7.8. Interestingly, a conical horn-parabola design provided one of the
lowest noise temperatures ever achieved (see sec. 9.13.2). In addition, this low-noise
satellite-communication ground terminal type of antenna has been also commonly used
for commercial microwave ground links, where the extremely low, wide-angle side lobes
allow high back-to-back isolation between antennas. Even so, strictly speaking, the horn-
parabola antenna is not an extremely low side-lobe type, because, according to Schrank
(1985), the near-in side lobes are not very low (about −20 dB).
An important aspect of the low side-lobe problem is measurement. Great care must
be taken to ensure that the side-lobe level measured on an antenna range is the true far-
field side-lobe level. As discussed in chapter 3, sec. 3.2.5, the conventional criterion for
an acceptable far-field measurement distance is 2D2/l, where D is the largest aperture
dimension and l is the wavelength. However, it has been pointed out by Hacker and
Schrank (1982) and Hansen (1984) that this criterion is inadequate for some types of
measurement, in particular for the measurement of very low side-lobe levels. The paper
by Hansen (1984) shows that for a Taylor-type pattern having a 50 dB first side-lobe
328 Antennas with Special Properties
level, accurate measurement of this level to within 0.5 dB requires separation of the
measured and measuring antennas by at least 12D2/l. (However, according to Hacker
and Schrank (1982), measurement at lesser distance initially affects the measurement
accuracy for only the side-lobes nearest to the main-beam.)
equations (Appendix D). Although GTD uses the ray-optics representation of electro-
magnetic propagation, it also incorporates diffraction theory and surface waves to account
for the effects of edge and surface discontinuities and surface-wave propagation. Surface
waves were briefly introduced in sec. 1.4.2.
Effects of each of the different electromagnetic phenomena used in GTD analyses are
discussed in the literature and are computed as coefficients (Hansen 1981). Then, each
of the coefficients, after being computed appropriately with algorithms, is used as a step
in the computational process for a specific antenna analysis. Depending on the geometry,
in GTD analyses there may be higher order effects like multiple reflections and diffrac-
tions. These occur in a complex environment, like an antenna on an automobile, near
ship structure, or near aircraft wings and engine nacelles. Burnside and Marhefka (1988)
published an extensive, one-hundred page treatment of GTD theory. Burnside, Rudduck,
and Marhefka (1980) published a summary of GTD computer codes developed at the
Ohio State University, and Marhefka (2000) is an example of more recently developed
GTD code. Examples of successful uses of GTD in calculating patterns of antennas
mounted on modeled aircraft and investigating the effects of the aircraft on the patterns
are addressed by Oortman and Ryan (2007).
An extensive set of antenna pattern calculations versus beam pointing direction for a
large, low-side-lobe beam-steered L-band (approximately 1.3 GHz) array, that includes
effects of aircraft interactions, is given by Allen (1992; 2004). The objective of the study
was to determine the optimum antenna location, when antenna performance and flight
dynamics are included. The required computations were very time intensive when made
with a major computer system at Lockheed Martin. Even so, the GTD method of locating
and moving an array on an aircraft was determined to be more cost-effective than a
program of measurements using a full-scale antenna and airframe or using a scale model
of antenna and aircraft.
element size increases the distance between element edges, and this reduces the mutual
coupling between elements. Second, reduced element size permits smaller center-to-
center spacing and thus more elements for a given array width, thereby permitting a wider
off-axis scan angle without creating grating lobes. Either type of use of reduced element
size can be effective in improving antenna patterns for off-normal beam directions.
Finally, the electromagnetics phenomena common to GTD, MoM, and fractal antennas
is radiation from sharp edges. Exact mathematical electromagnetic solutions cannot be
obtained for antenna configurations containing sharp edges, but it has been demonstrated
that accurate approximations are possible by using GTD and MoM. GTD or MoM is
used when dimensions are large or small, respectively, compared to a wavelength. The
mathematics of fractals are relatively easy to computer-generate, but fractal shapes have
sharp edges. However, experience shows that the electromagnetic characteristics of
fractal wire and patch antennas can be calculated with MoM techniques. Thus, it is appar-
ent that the antenna technology that has evolved from GTD, MoM, and fractals could
not have been developed without digital computers, because reduction to practice would
have been a mathematically impracticable procedure.
References
Allen, W. P., “Aircraft Interference Effects on AEW Antenna Patterns,” pp. 326–44 of
M. W. Long, Airborne Early Warning System Concepts, Artech House, 1992 and
SciTech, 2004.
Balanis, C. A., Antenna Theory, 3rd ed., John Wiley & Sons, 2005.
Booker, H. G., “Slot Aerials and Their Relation to Complementary Wire Aerials,” Journal
of the Institution of Electrical Engineers (London), 93, pt. IIIA, no. 4, 1946.
Booker, H. G., and P. C. Clemmow, “The Concept of an Angular Spectrum of Plane
Waves, and Its Relation to That of a Polar Diagram and Amplitude Distribution,”
Proeedings of the Institution of Electrical Engineers, vol. 97, pt. III, January 1950,
pp. 11–17.
Burdine, B. H., and E. J. Wilkinson, “A Low-Sidelobe Earth-Station Antenna for the
4/6 GHz B and,” Microwave Journal, vol. 23, no. II, November 1980, pp. 53–68.
Burnside, W. D., and R. J. Marhefka, “Antennas on Aircraft, Ships, or Any Large Complex
Environment,” Ch. 20 in Antenna Handbook—Theory, Applications, and Design, Lo,
Y. T. and S. W. Lee, eds., Van Nostrand, Reinhold Co., 1988.
Burnside, W. D., R. C. Rudduck, and R. J. Marhefka, “Summary of GTD Codes
Developed at the Ohio State University,” IEEE Transactions on Electromagnetic
Compatibility,” EMC-22, November 1980, pp. 238–243.
Campbell, D. V., “Personal Computer Applications of MININEC”, IEEE Antennas and
Propagation Society Newsletter, February 1984.
Cutrona, L. J., “Synthetic Aperture Radar,” ch. 21 in Radar Handbook, M. I. Skolnik (ed.),
McGraw-Hill, New York, 1990.
Deschamps, G. A., and R. H. DuHamel, ch. 18 in Antenna Engineering Handbook, H.
Jasik (ed.), McGraw-Hill, New York, 1961.
References 333
Electronically Steered
Arrays
This chapter discusses electronically steered arrays (ESAs), also commonly called phased
arrays. These antennas have beams that can be electronically steered in pointing direction
or beam shape, without mechanical motion, and accomplished in microseconds. Although
the operation of all arrays depends on the proper phasing of the individual array elements,
the terms ESA and phased array have by usage come to mean an array in which the
beam is steered in direction and shape by varying the phasing of the elements. Present-
day ESA (or phased array) technology includes adaptive beam forming (ABF), which
merges rapid beam pointing and shaping with digital signal processing separately. With
ABF arrays, the phasing and time delaying of the signals from individual elements are
jointly adaptively adjusted and then summed to rapidly beam point, beam shape, and
Doppler filter desired received signals. In this way, the received signal output, from the
summed individual array signals, may be optimized, simultaneously, for both the strength
of the desired signal relative to thermal noise and relative to interference from undesired
signals.
The advancements in ESA (i.e., phase array) technology has resulted from intensive
and continuing research and development in the following areas:
(i) Improvement of the devices and circuitry used for producing the required
phase shifts and power division among the array elements;
(ii) Use of microprocessors or special-purpose computers to determine the
optimum phasing and power division to carry out a particular beam-steering
command;
(iii) Better understanding of factors affecting side-lobe levels, mutual coupling
effects on array-element patterns (and methods of minimizing the deleteri-
ous effects of such coupling), and effects of phase errors on side lobes and
main-beam gain;
(iv) Development of adaptive receiving arrays that are capable of automatically
adjusting their overall patterns to optimize signal-to-interference ratio by
placing pattern nulls in the directions of interfering signals;
(v) Development of solid-state amplifiers suitable for use at transmit-receive
array elements, making the use of amplifiers distributed among the array
elements more economically feasible and the arrays more efficient; and
337
338 Electronically Steered Arrays
Lc
os
θL
W
av
ef
dn
ro
nt
(lin
e
of
n
eq
δ
θL
ua
lp
ha
se
L
θL
Distance to wavefront: δn = dnsinθL
Element phase (radians): αn = (2π/λ) δn
inθ
L
Ls
FIGURE 8–1.
Effective aperture (line of equal phase, LcosqL) versus scan angle qL.
340 Electronically Steered Arrays
top element relative to the bottom element. Appropriate, decreasingly less time
delay is required for the lower elements for scan angle of qL. Clearly, for the beam to be
scanned in the −qL direction, the larger time delays will be needed for the lower
elements.
Consider now the following example, with qL = 30° and L/l = 20. Then focusing will
occur if a time delay is added to the top element that corresponds to the time it takes for
a wave in free space to traverse a distance of 10l. This delay can in principle be obtained
by adding a transmission line. However, for each scan angle a different length of trans-
mission line would be required—and the lengths would be different for each array
element. Now imagine an array with several hundred (or perhaps thousands of) elements;
and the capability to change to a different length of transmission line at each element
and each scan angle. Although time delayers permit an array to be focused over a wide
bandwidth, time delayers are bulky, heavy, and expensive. Therefore, time delayers are
used only when very broad-band performance is required. Instead, phase shifters (often
called phasers) are commonly used to provide agile beam-pointing capabilities.
As already noted, for a focused array the radiation from each element at a distant
observation point is in-phase. Now suppose there is a phase shifter (phaser) located at
each array element. Then, for narrow bandwidth operation, the required range of phase
shift is between zero and 2p, that is, between zero and 360°. However, each phaser must
change for each change in either scan angle or operating frequency. For focusing with
an element spacing d, there must a progressive decrease, from top to bottom in Fig. 8–1,
in phase delay a between elements as follows:
2π d
α = sin θ L
λ
(8–1)
Now assume a of (8–1) equals Df + 2pI, with I being any integer. Then, the phase
delay needed by a phaser is only Df. This mathematical unwinding of phase is called
modulo 2p. Thus, Df is expressed as mod (2p) in radians or as mod (360°) in degrees.
Recall that for perfect phasing of an array, the required phase shift for each element is
different, and all of the phases must change with a change in either scan angle qL or fre-
quency. Even so, to retain perfect array phasing, the needed phase delay of the phasers
never exceeds 360°.
Radiating Elements
Input Termination
S
Serpentine Feed
FIGURE 8–2.
Frequency scan linear array.
2π d
α = sin θ L + 2π I
λ
(8–2)
342 Electronically Steered Arrays
where l is free space wavelength and I is an integer. Letting ls denote the wavelength
within the transmission line of length s between elements, the interelement phase delay
is 2ps/ls. This phase delay must equal a of (8–2) if the array is focused. Thus,
2π d 2π s
sin θ L + 2π I = (8–3)
λ λs
Usually the connecting transmission line is air-filled waveguide operated in the TE01
mode. Its waveguide wavelength is given by (2–69) of chapter 2, and is repeated here
as follows:
c
λg = (8–4)
f 2 − fc2
In (8–4), c is the free space velocity of propagation and fc is the waveguide cutoff fre-
quency. If waveguide is used for the transmission line, ls = lg. Then from (8–3), if qL
is zero
s
I= (8–5)
λg0
Here lg0 is the wavelength within each connecting line of length s at the frequency f0
for which the beam is pointed in the broadside direction, that is, qL = 0. In other words,
I is the integer number of wavelengths along the transmission line at the wavelength
lg0.
Using (8–3), (8–4), and (8–5), it can be shown that
Iλ f 2 − fc2 cI f 2 − fc2 1 2
12
sin θ L = 1 − = 1 − (8–6)
d f02 − fc2 fd f02 − fc2
In (8–6), only values of f within the interval for which |sin qL| ≤ 1 are permitted.
spectrum, and therefore the shape of a short pulse. Category 2, called tunable bandwidth,
is one where there is enough time available for setting antenna parameters prior to a fre-
quency change. Instantaneous bandwidth (category 1) presents the more difficult problem,
because the antenna must function properly, without adjustments, at any frequency within
the band.
As already discussed, arrays having broad instantaneous bandwidth must use time
delayers. Time delayers are heavy and expensive. For these reasons, most scanning arrays
use phase shifters for scanning, and consequently they have only limited instantaneous
bandwidths. Array elements and phase shifters (phasers) can limit overall bandwidth.
However, array elements usually can perform over bandwidths of 10% without major
changes in impedance or beamwidth. Recall that elements of a focused array require
equal time delay to a point in space, and phase delay for a given time delay is directly
proportional to frequency. Therefore, bandwidth is limited by phase shifters, because
their phase delay is nearly independent of frequency.
The beam direction of an array that is steered by phase shifters will automatically shift
as the frequency is changed. This beam shifting limits the array bandwidth. Most modern
arrays can, however, be “tuned” (by resetting the phase shifters) over a bandwidth of five
to perhaps ten percent. Frank (1972) gives an estimated first guess at bandwidth, without
tuning, as
In other words, a phased array will have roughly a one percent instantaneous bandwidth
if it has a broadside −3 dB beamwidth of one degree. The actual bandwidth will depend
on the feed architecture, transmitted waveform, and scan angle. According to Frank, (8–7)
above will provide a bandwidth estimate that is rarely off by a factor of more than
two.
Determining a more accurate estimate requires a thorough examination of how the
array will be used. For example, the bandwidth decreases with increases in scan angle.
In addition, signals that are modulated, for example, pulsed signals, have a widened
spectrum, and this must be allowed in estimating needed bandwidth. In other words,
because of the different frequencies within a modulated signal, an array beam is spread
in angle. This spreading reduces the effective gain.
For a scanned beam to remain perfectly focused when frequency is changed, the
element phases must be changed. Without resetting, phase shifters are essentially constant
phase devices, with changes of only a few degrees over their operating bandwidths.
For (8–8), (8–9), and (8–10) that follow, it is assumed that the transmission lines to the
array elements are identical and phases of the phase shifters are independent of
frequency.
From (8–1), the beam is steered to a direction qL with a phase of
2π x 2π x
φ= sin θ L = f1 sin θ L (8–8)
λ1 c
344 Electronically Steered Arrays
at an element located a distance x from the array center. With the same phase setting, a
frequency f2 steers the beam to a new position qL + DqL. Then,
2π x 2π x
φ= f1 sin θ L = f2 sin (θ L + Δθ L ) (8–9)
c c
It can be shown that, for small changes in Df, where Df = f2 − f1, the change in direc-
tion of the beam DqL is small, and then (8–9) can be approximated as
Δf
Δθ L = tan θ L (8–10)
f
The relationship (8–10) indicates that the amount of beam shift DqL increases in
proportion to the product of fractional change in frequency and tangent of scan angle.
However, (8–10) is of limited usefulness, because it is valid only for small changes in
frequency.
To make practical estimates of beam shift, one must resort to approximations using
(8–9). Furthermore, the loss in signal strength for a given beam squint will necessarily
depend on beamwidth, and hence on aperture size. Frank gives graphs of loss in gain as
a function of beamwidth, scan angle, and spectral width. Equation (8–7), a commonly
used estimate for bandwidth, gives a gain loss of approximately 0.8 dB (due to the beam
shift) for an unmodulated signal with the beam scanned 60° from broadside. For a signal
with a broad spectrum, such as a short pulse, effective bandwidth may be less than indi-
cated by (8–7). This is because an array must be focused at all frequencies contained
within the spectral bandwidth of the signal being either transmitted or received.
Interestingly, antenna performance with short pulses can also be analyzed in terms of
the time required for a received pulsed wave to reach all of the array elements (Frank,
1972). This time is called the array fill time, which is (L sin qL)/c. Thus, fill time is the
time it takes for a wave traveling at speed c to traverse the distance L sin qL. The role of
a time delayer is, of course, to add delay that equals the fill time that will, in principle,
result in focusing that is independent of frequency.
g (θ , φ ) = ∑ ∑ a ( x, y ) exp [ jΨ ( x, y )] (8–11)
where
a(x, y) = the amplitude excitation of each element
Y(x, y) = a − (2p/l)d
Phased Array Principles 345
with
a = phase excitation of each element and
d = pathlength difference of each element (radiator) to the observation point relative
to pathlength from the array center, with the observation point in the far zone (see
Appendix H),
d = −sin q (x cos f + y sin f)
Usually, the elements of rectangular arrays are energized so that their amplitudes and
phases along the x-axis aperture coordinates are independent of the y-axis coordinates,
and conversely. In that case, the element amplitudes and phases can be expressed as
products of terms that depend only on one coordinate, either x or y, as follow:
a ( x, y ) = a ( x ) a ( y )
α ( x, y ) = α ( x ) α ( y )
Then, the integrand of (8–11) is separable as two products, one that is a function of the
x coordinates and the other a function of the y coordinates. In addition, electronically
steered arrays are usually designed so the separations between elements are evenly
spaced, but not necessarily the same along the x and y axes. Then, the x and y coordinates
can be expressed as xm = mdx and yn = ndy, where dx and dy are the element spacings in
the x and y directions. Similarly, a(x), a(y), a(x), and a(y) can be expressed as am, an,
am, and an, respectively. Then (8–11) can be written as the following product of two
summations
where far-zone amplitude patterns are calculated by summing the excitation of the ele-
ments over all m and n (x and y) values. The reader will recognize that each of the two
product terms of (8–12) is an array factor for a linear array, and thus each can be con-
sidered separately as if describing the radiation from a linear array.
Effects of beam steering are now considered, recognizing that scanning may be accom-
plished separately in each plane. An array beam is focused if the phases of the fields
from its elements are equal at a great distance. In other words, a focused antenna provides
a field of equal phase along a line normal to the beam-pointing direction, as in Fig. 8–1.
Consider now the pattern in the x-z plane for a line source along the axis. Then, since
there are no elements along the y axis an and an are zero. In addition, since the pattern
in the x-z plane is being considered, f = 0 and cos f = 1. Therefore, from (8–12), the
pattern in the x-z plane for a line source along the x-axis versus observation angle q is
M −1
2π
g (θ ) = ∑ am exp j α m + λ
md x sin θ
(8–13)
m=0
346 Electronically Steered Arrays
where M is the number of elements and there are equally spaced elements located at
coordinates mdx.
The pattern amplitude |g(q)| is maximized when in (8–13) the term
α + 2π md
m x sin θ = 0 . Thus, for the beam to be steered for its peak amplitude in direc-
λ
tion q = qL, each element phase am must be as follows:
2π
αm = − md x sin θ L (8–14)
λ
Then, when the phases are set in accordance with (8–14) to maximize the pattern in
direction qL, the pattern is as follows:
M −1
2π
g (θ ) = ∑ am exp j λ
md x (sin θ − sin θ L )
(8–15)
m=0
The reader should recall that qL in (8–15) is the pointing direction of the major lobe and
q is the observation angle in an arbitrary direction.
2π
md x (sin θ − sin θ L ) = ± 2πΙ (8–16)
λ
where I is an integer. When (8–16) is satisfied, the maxima of g(q) are grating lobes. As
a consequence of (8–16), element spacing must be less than l for an array that is scanned.
In particular, to prohibit grating lobes when the major lobe is pointed in directions other
than broadside, the maximum acceptable element spacing dmax is
λ
dmax = (8–17)
1 + sin ∆θ
From (8–17), for Dq equaling 90°, dmax is l/2. Because of loss in effective aperture
due to pointing off of broadside, ESA designs typically limit the scan to a maximum of
60° off-broadside. Then, sin Dq = 0.866 and dmax = 0.54 l. Consequently, the element
spacings for ESAs are usually about l/2, or slightly less, to avoid the risk of creating
part or all of a full grating lobe.
Thus far it has been assumed that the array elements are arranged in a square grid with
separation in height and width of dmax between elements. Now consider a lattice configu-
ration of equilateral triangles, with each triangle leg having length dmax. Such a configu-
ration, when considering only the avoidance of grating lobes, will allow (depending on
maximum scan angle off-broadside) a 10 to 15 percent reduction in the overall number
of required elements (Sharp 1961). However, the triangular configuration creates a some-
what more stringent requirement on the fineness of phase quantization for controlling
side-lobe levels (Nelson 1969).
Some ferrite phase shifters require a continuous magnetizing current, while others are
of the “latching” type, in which a relatively short pulse establishes a magnetization level;
this level is maintained until it is reset by another dc pulse. However, when the magne-
tization level is to be changed for either the continuous-current or the latching type of
ferrite phase shifter, it is necessary first to set the magnetization momentarily at the satu-
ration level and then to apply the current required to set the desired level. This avoids
errors that would otherwise occur due to the hysteresis effect.
Ferrite phase shifters can be reciprocal or nonreciprocal. In the former case, the same
amount and direction of phase shift occurs for either direction of propagation. In the
latter case, the sign of the phase shift is reversed when the direction of propagation is
reversed. When a nonreciprocal phase shifter is used for both transmission and reception
with the same antenna (as is the case for many radar systems), the phase shifter must be
switched rapidly between the transmitting and receiving modes of operation. Despite this
requirement, nonreciprocal phase shifters are sometimes used in radar applications
because of other desirable characteristics. Switching times of a few microseconds are
achievable.
Ferrite phase shifters can be either analog or digital, and even when an intrinsically
analog shifter is used, it may be digitally controlled. These phase shifters become increas-
ingly inefficient and bulky as the frequency is decreased. They are not generally used
below about 1 GHz, although they have been used at frequencies as low as 200 MHz.
On the other hand, ferrite phase shifters are usable at frequencies of the order of 10 GHz
or higher.
The primary electronic method employed for varying the length of a transmission line
is diode switching. Diode phase shifters, unlike ferrite phase shifters, become increas-
ingly efficient as the frequency decreases, although they also can be used up to 10 GHz.
For purposes of basic understanding of its phase-shifting capability, the diode can be
regarded as a switch, although in one type of phase-shifter it behaves as a variable capaci-
tive susceptance. The simplest type of diode phase shifter is the switched line phase
shifter, consisting of elements in which diode switches select either of two transmission
line lengths. This is a digital device, as are most diode phase shifters. As an example of
its action, if the two transmission-line segments through which the signal can be alter-
nately switched differ in length by a quarter wavelength, the resulting phase shift will
be 90 degrees (since a 360-degree phase change occurs for each wavelength of a propa-
gation path). Other well-known types of diode phase shifters are the hybrid-coupled
phase shifter and the periodically-loaded-line phase shifter (Tang and Burns, 1993).
A digital phase shifter is constructed by cascading several units, each of which can be
switched between two fixed phase-shift values. The incremental phase shift of each suc-
cessive unit is twice as large as that of the next-smaller unit, and the largest increment
is 180 degrees. This permits all possible phase shifts to be achieved, in increments of the
smallest unit, by switching the appropriate elements from one state to the other. For
example, a 4-bit phase shifter has incremental phase-shift elements of 22.5, 45, 90, and
180 degrees (and zero), as shown in Fig. 8–3. This permits phase delays of 22.5, 45,
67.5, 90, 112.5, 135, 157.5, 180, 202.5, . . . , and so on. (For an n-bit phase shifter, the
number of possible delays is 2n, including zero phase shift which is equivalent to 360
Beam-Steering Technology 349
effects when subarrays are used. For details of these matters, see Cheston and Frank
(1970, pp. 11–37).
avoiding major flaws in design, which include detailed preliminary design, simulation,
computer analyses, and measurements.
division among the elements; it is difficult to accomplish this for a many-element array
by branching from a single transmission line or waveguide into many lines all at one
division point. The corporate feed scheme, shown in chapter 2, Fig. 2–8(a), avoids
this problem while still preserving equal line lengths by the branching method.
At each branch, equal power division or any desired tapering can be obtained by suitable
impedance relationships at the junctions. Alternatively, hybrid couplers can be used to
minimize the coupling between element transmission lines, caused by impedance
mismatches.
If a beam at some direction other than broadside is wanted, then the lengths of the
lines to the various elements can be made unequal to provide the desired phase progres-
sion, or alternatively, phase shifters and/or time delayers can be inserted in the lines.
ELEMENTS
PHASE SHIFTERS
HYBRID JUNCTIONS
RF LOADS
TIME DELAYERS
CORPORATE FEED
FIGURE 8–5.
Corporate feed for an eight-element array, consisting of eight phase shifters and two time
delayers.
lens in which a space-fed array of waveguides are energized by an E-plane feed. In this
example, the antenna contains twenty-four waveguides (8 columns by 3 rows), each
having a phase shifter or a time delayer so that the beam may be steered in both azimuth
and elevation. The reader will note the similarity with the metal waveguide lens of Fig.
6–23, except in that case the beam cannot be electronically steered.
Pwr
ampl
Transmit
Transmit/ Transmit/
Phase
receive receive Antenna
shifter
switch switch
Receive
LNA
FIGURE 8–7.
Simplified diagram of a T/R module that serves as both transmitter and receiver at each
element of an active array. From Merrill I. Skolnik, Introduction to Radar Systems, 3rd ed.,
© 2001, p. 600, with permission of The McGraw-Hill Companies.
Phased Array Feed System Technology 355
in conformal arrays than in planar arrays because the element patterns in different parts
of the conformal array are oriented differently.
In the side-lobe canceller, the signals received by the auxiliary elements are processed
with the main-antenna signals in such a way that they create nulls of the main antenna
pattern in the jammer directions. The theory of this processing involves the concept of
the “degrees of freedom” available, which are equal to the number of low-gain elements
in the auxiliary array. This number in turn determines (generally is equal to) the number
of separate jamming-signal directions that can be nulled.
The mathematical algorithms of this signal processing and control tend to be very
complicated because they are partly based on multivariate estimation theory. Implemen-
tation of some of these algorithms can be either analog or digital, but in modern systems
digital technology is predominant. The adaptive process usually involves a “convergence
time,” and some of the simpler algorithms may converge relatively slowly, especially for
certain angular distributions of the interfering signals. However, rapid convergence times
are obtainable with the more sophisticated algorithms and techniques.
The side-lobe canceller, using a relatively few auxiliary array elements, is a “partial”
adaptive array. A fully adaptive array is one in which the entire system is an array and
all of the array elements contribute equally to the operation of the adaptive processor.
This allows much greater flexibility and control of the total pattern (i.e., there are more
degrees of freedom). However, the cost of the processor is much higher for arrays having
a very large number of elements, such as those used in many radar systems.
The principle of a fully adaptive array was demonstrated with computer computations,
however, in the above-referenced paper by Gabriel. In that 1983 paper, Gabriel’s array
had sixteen elements and was assumed to operate at S-band (approximately 3,000 MHz).
Further assumptions included a separate receiver for each array element and each of the
received signals, after conversion to an intermediate frequency (IF), was split into
two baseband I and Q signals (see Fig. 8–11). As is customary, I and Q denote the
in-phase and phase-quadrature components of an IF signal, and they preserve the phase
information in the “zero frequency” (baseband) channels that is necessary for digital
processing.
Figures 8–9(a) and 8–9(b) show results
of Gabriel’s computer analyses. In Fig. 0
8–9(a), the “quiescent” pattern of the
POWER IN DECIDELS
possible to retain the quiescent side-lobe levels in the adaptive patterns, but this requires
the application of additional computations in the adaptive algorithm (Frost 1972).
The technology of fully adaptive arrays is still in a state of intensive development,
and this was also the case as pointed out by Gabriel (1983). Some of the other early
classic papers include Gabriel (1976), Childers (1978), Gabriel (1980), Brennan and Reed
(1973), Brennan, Reed, and Swerling (1974), and Monzingo and Miller (1980). As
already indicated, development of adaptive or “smart” arrays continues to be an active
field, having its origins in the suppression of side lobes of a main antenna by use of
several smaller auxiliary antennas (Gabriel 1976). Today pattern optimization is done
with real-time weighting of received signals in such a way that the array on reception
can adapt to the interference environment. In principle it can also be done by weighting
the transmit antenna pattern, but this gen-
0 erally is not done. An extensive list of
up-to-date references is given by Gross
POWER IN DECIDELS
–10 (2007).
Rx ADC
ELEMENTS
Rx ADC
ARRAY
ADAPTED
+ OUTPUT
Rx ADC
WEIGHTS
ALGORITHMS
CONTROLLER
FIGURE 8–10.
Basic components of an adaptive array.
Adaptive Array Antennas 361
IF
MIXER
LOW
RF PASS A/D Q
MIXER FILTER
BAND
IF PASS
AMP FILTER IF
MIXER
LOW
PASS A/D I
RF LO FILTER
90°
IF LO
FIGURE 8–11.
Typical low-noise coherent (phase and amplitude) I and Q detection receiver.
computer algorithms for establishing phase and amplitude weights that modify the indi-
vidual received signals. Then, the modified signals are summed for providing interfer-
ence-suppressed outputs. The outputs can be from multiple signals that may have
simultaneously arrived at the array elements. For some algorithms, a sample of the
summed signals is fed back to the weight controller, as suggested by the dashed line in
Fig. 8–10.
Figure 8–11 illustrates the principal features of a receiver that supplies, in digital
format, the I and Q signals that contain phase and amplitude information, where ampli-
tude “a” is
a = I 2 + Q2
and phase j is
ϕ = arc tan ⎛⎜ ⎞⎟
Q
⎝I⎠
The mathematics of pattern optimization uses complex matrix theory. There is cur-
rently much activity in the development in adaptive algorithms, in computational capa-
bilities, and in improved array hardware (see, e.g., Balanis 2005). Presently used
algorithms are generally based on the early work of (1) the Howells-Applebaum method
(Applebaum 1976) and (2) a procedure due to Widrow and McCool (1976) that mini-
mizes the least square mean (LMS) between the array output signal and a reference
signal. The Howells-Applebaum method is commonly used with radar systems to affect
signal-to-noise optimization, subject to an array pattern formed in the absence of interfer-
ence. This requires sampling over a large number of range and Doppler cells to obtain
362 Electronically Steered Arrays
data to which the array system is optimized. The LMS method is used in communications
systems, for which a reference signal can use a replica of the received signal format.
Sometimes the LMS method is used with radar, and then the waveform of the transmit
signal can be used as the reference.
T w11 T T wN1
T w12 T wN2
w13 wN3
v (t)
FIGURE 8–12.
Generalized Space-Time Processor having N element channels and M delay taps per channel,
for cooperative space and time adaptive processing. From Barile, Fante, and Torres (1992);
© 1992, IEEE).
Adaptive Array Antennas 363
variety of frequency filtering options. For example, the availability of complex (ampli-
tude and phase) time-delayed signals permits fast Fourier transform (FFT) processing
that provides, in essence, signals from multiple band-pass filters. Finally, it is to be noted
that, in principle, the Fig. 8–12 configuration permits optimized frequency filtering and
beam shaping to be accomplished independently, or cooperatively (i.e., optimized
collectively).
As may be imagined, there are numerous design options. For example, the Fig. 8–12
configuration permits each antenna element to have a phase response that varies with
frequency. A phase response can compensate for the fact that a given path length l gives
a phase delay (2pl)/l = (2pl)f/c that depends on frequency. In other words, the tapped
delay-line may perform as an equalizer, to make the array have approximately the same
pattern over a range of frequencies. Wireless arrays using this phase feature are known
as wideband arrays and they are said to employ space-time, spatio-temporal, or two-
dimensional processors.
Thus, the wireless “two-dimensional” adaptive array provides the structure for sup-
pressing interfering signals based on frequency differences and on spatial separations
cooperatively. This configuration is said to be particularly valuable in heterogeneous
communications environments where signals of different bandwidths and different carrier
frequencies may be incident on the array (Liberti and Rappaport 1999, p. 102).
Features of a wideband wireless array are also of advantage for narrow band signals.
Consider, for example, multipath signal components that may or may not be separated
in angle or by significant time delays. The two-dimensional adaptive array can capture
multipath components separated in either time or angle, and thereby combine features of
both a spatial processor and a temporal equalizer (Liberti and Rappaport 1999, p. 102).
Computer throughput requirements and ultimate system performance can be altered
considerably, depending on how the element signals are processed. Processing may be
done on the signals from each array element, and then the separately processed element
signals can be jointly optimized. For example, one might first frequency-filter the M
time-delayed signals from each of the elements separately and then optimize the N
frequency-filtered element signals by adaptive beam-shaping and pointing. Alternatively,
the processing might first optimize beam shape and pointing direction for each of the M
different time delayed signals of the N elements separately, and then frequency-filter the
N signals that had previously been optimized by beam shaping and pointing.
Some readers may recognize that tapped delay-line filtering is commonly used in pulse
Doppler radar for suppressing clutter and noise. This filtering when combined with the
spatial adaptive features of Fig. 8–12 offers improved clutter and noise suppression
capabilities that are especially beneficial for radars on moving platforms (Skolnik 2001,
pp. 168–71).
Finally, fully adaptive arrays are costly and complex, but they offer many advantages
including
beamwidths of 1.6 and 3.2 degrees in elevation and azimuth, respectively. The elevation
coverage is from 0 to 19 degrees, with azimuth scan rates of either 6 or 12 rpm (option-
ally). The gain is 38.9 dB. Radiating elements are arranged in rows and columns, and
each of the fifty-four rows is fed by solid-state transmitter amplifiers; the needed rela-
tively large average power is made possible by the use of many low-power transmitter
amplifiers. Advantages of this technology are that failure of one or even a few of the
amplifiers does not totally disable the radar, and no portion of the system is subjected to
extremely high power levels with attendant problems of heating and voltage breakdown.
Receiver solid-state amplifiers and low-power-level phase shifters are similarly distrib-
uted. Both the TPS-32 and the TPS-59, incidentally, are transportable radars which can
be set up for temporary operation at field sites.
The Navy AN/SPY-1 shipboard multifunction S-band radar, a part of the Aegis weapon
system, has a phased-array antenna with true phase scanning in both azimuth and eleva-
tion (Bookner 1977; Scudder and Sheppard 1974). Four separate arrays are mounted on
four faces of the ship superstructure so that each covers an azimuth quadrant. The vertical
coverage of each array is sufficient so that the four arrays provide virtually hemispheric
coverage for the radar. Each array is of approximately hexagonal shape, about 12 feet in
width and 12 feet 7 inches in height. The vertical and horizontal beamwidths are of the
order of 2 degrees.
The radiating elements are rectangular horns (chapter 4, sec. 4.7), grouped into sub-
arrays as described in the preceding section of this chapter. The subarray arrangement is
somewhat different for transmitting and receiving. There are 32 transmit subarrays, each
containing 128 horn radiators. For receiving, the same horns are grouped into 64 sub-
arrays of 64 horns each. However, on reception four more subarrays are added to the
complete array, that is, these four subarrays are utilized for reception only. Thus there is
a total of 4,096 horns in the transmitting array and 4,352 horns in the receiving array.
The purpose of the additional four subarrays used on reception is to improve the sym-
metry of the array monopulse difference patterns, which are generated simultaneously
with the normal “sum” pattern, as described in Sec 8.3.3. The additional reception sub-
arrays also contribute to an improved side-lobe level on reception.
The phasing of the individual horns of the array is accomplished by ferrite (garnet)
nonreciprocal digitally-controlled latching phase shifters. The phasing is done on an
individual-element basis; the steering commands are based on the so-called row-column
method. The subarrays are not involved in this aspect of the antenna operation.
This description applies to the first operational model of the antenna, which differs
from the earlier experimental model described by Scudder and Sheppard (1974). As with
most major systems, newer antenna designs have been developed since the first opera-
tional model.
Each transmit subarray is fed by a separate high-power amplifier, whose output power
is then divided equally among the 32 subarray horns. On reception, the individual sub-
arrays do not have corresponding rf amplifiers; but there is a low-noise rf amplifier
mounted on the back of the array for each of the three monopulse channels (sum, azimuth
difference, and elevation difference). This does not eliminate the signal-to-noise reduc-
tion caused by losses in the on-array transmission-line components and phase shifters,
366 Electronically Steered Arrays
References
Ajioka, J. S., “Frequency-Scan Antennas,” ch.19 in R. C. Johnson, Antenna Engineering
Handbook, 3rd Edition, McGraw-Hill, 1993.
Allen, J. L., “Gain and Impedance Variation in Scanned Dipole Arrays,” IEEE
Transactions on Antennas and Propagation, vol. AP-10, September 1962, no. 5,
567–69.
Applebaum, S. J., “Adaptive Arrays,” IEEE Transactions on Antennas and Propagation,
Sept. 1976, pp. 585–98.
Archer, D., “Lens-Fed Multiple-Beam Arrays,” Microwave Journal, vol. 18, no. 10,
October 1975.
Balanis, C. A., Antenna Theory, Analysis and Design, 3rd ed., Wiley-Interscience, 2005,
ch. 16.
Barile, E. C., R. L. Fante, and J. A. Torres, “Some Limitations on the Effectiveness of
Airborne Adaptive Systems,” IEEE Transactions on Aerospace and Electronic Systems,
vol. 28, no. 4, October 1992, pp. 1015–32.
Brennan, L. E., and I. S. Reed, “Theory of Adaptive Radar,” IEEE Transactions on
Aerospace and Electronic Systems, AES-19, no. 2, March 1973, pp. 237–52.
Brennan, L. E., I. S. Reed, and P. Swerling, “Adaptive Arrays,” Microwave Journal,
vol. 17, no. 5, May 1974, pp. 43 ff.
Brookner, E., Radar Technology, Artech House, 1977.
Butler, J. L., “Digital, Matrix, and Intermediate Frequency Scanning,” in Microwave
Scanning Antennas, vol. III, R. C. Hansen, ed., Ch. 3, Academic Press, 1966.
Butler, J., and R. Lowe, “Beamforming Matrix Simplifies Design of Electronically
Scanned Antennas,” Electronics Design, April 1961, pp. 170–73.
Carlson, B. D., L. M. Goodman, J. Austin, M. W. Ganz,and L. O. Upton, “ An
Ultralow-Sidelobe Adaptive Array Antenna,” The Lincoln Laboratory Journal, Vol. 3,
No. 2, 1990, pp. 291–310.
Cheston, T. C., and J. Frank, “Array Antennas,” ch. 11 in Radar Handbook, M. I. Skolnik
(ed.), McGraw-Hill, 1970.
Childers, D. G., Modern Spectrum Analysis, IEEE Press, 1978.
Frank, J., “Bandwidth Criteria for Phased Array Antennas,” in A. A Oliner and G. H.
Knittel, Phased Array Antennas, Artech House, 1972, pp. 243–53.
Frost III, O.L., “An Algorithm for Linearly Constrained Adaptive Array Processing,”
Proceedings of the IEEE, vol. 60, no. 8, August 1972, pp. 926–35.
Gabriel, W. F., “Adaptive Arrays: An Introduction,” Proceedings of the IEEE, Feb. 1976,
pp. 239–72.
Gabriel, W. F., “Spectral Analysis and Adaptive Arrays Superresolution Techniques,”
Proceedings of the IEEE, vol. 68, June 1980, pp. 654–66.
Gabriel, W. F., “Adaptive Processing Antenna Systems,” IEEE Antennas and Propagation
Newsletter, October 1983.
Gross, F. B., “Smart/Antennas,” ch. 25 in Antenna Engineering Handbook, 4th ed., J. L.
Volakis (ed.), McGraw-Hill, 2007.
Hansen, R. C. (ed.), Significant Phased Array Papers, Artech House, 1973.
References 369
Hansen, R. C., “Phased Arrays,” ch. 20 in Antenna Engineering Handbook, 4th ed., J. L.
Volakis (ed.), McGraw-Hill, 2007.
Hendrix, R., “Aerospace System Improvements Enabled by Modern Phased Array Radar—
2008,” Proceedings of the 2008 Radar Conference, Rome, Italy, May 26–30, 2008,
pp. 275–280.
Johnson, R. C., Antenna Engineering Handbook, 3rd ed., McGraw-Hill, 1993.
Johnson, R.C., and H. Jasik, Antenna Engineering Handbook, 2nd ed., McGraw-Hill, 1984.
Liberti, J. C., Jr., and T. S. Rappaport, Smart Antennas for Wireless Communications,
Prentice Hall PTR, 1999, p. 102.
Mailloux, R. J., “Phased Array Theory and Technology,” Proceedings of the IEEE, vol. 70,
no. 3, March 1982, pp. 263–64.
Mailloux, R. J., Phased Array Antenna Handbook, Artech House, 2005.
Miller, C. J., “Minimizing the Effects of Phase Quantization Errors in an Electronically
Scanned Array,” Proceedings of the 1964 Symposium, Rome Air Development Center
Document RADC-TDR-64–225, vol. 1, pp. 17–38.
Monzingo, R. A., and T. W. Miller, Introduction to Adaptive Arrays, John Wiley and Sons,
1980.
Nelson, E. A., “Quanization Sidelobes of a Phased Array with a Triangular Element
Arrangement,” IEEE Transactions on Antennas and Propagation, vol. 17 (May 1969),
pp. 363–65.
Oliner, A. A., and G. H. Knittel, Phased Array Antennas, Artech House, 1972.
Romanofsky, R. R., “Array Phase Shifters: Theory and Technology,” ch. 21 in J. L.
Volakis, Antenna Engineering Handbook, 4th ed., McGraw-Hill, 2007.
Rotman, W., and R. F. Turner, “Wide-Angle Microwave Lens for Line-Source
Applications,” IEEE Transactions on Antennas and Propagation, AP-11, no. 11,
November 1963, pp. 623–32.
Scudder, R. M., and W. H. Sheppard, “AN/SPY-1 Phased-Array Antenna,” Microwave
Journal, vol. 17, no. 5, May 1974.
Sharp, E. D., “A Triangular Arrangement of Planar-Array Elements that Reduces the
Number Needed,” IEEE Transactions on Antennas and Propagation, AP-9, March 1961,
pp. 126–29.
Shelton, J. P., and K. S. Kelleher, “Multiple Beams from Linear Arrays,” IEEE
Transactions on Antennas and Propagation, March 1961, pp. 154–61.
Skolnik, M. I., Introduction to Radar Systems, 3rd ed., McGraw-Hill, 2001.
Stark, L., “Microwave Theory of Phased Array Antennas: A Review,” Proceedings of the
IEEE, December 1974, pp. 1661–1701.
Steyskal, H., “Digital Beamforming at Rome Laboratory,” Microwave Journal, Feb. 1996,
pp. 100–124.
Tang, R., and R. W. Burns, “Phased Arrays,” pp. 20–31 to 20–48, in R. C. Johnson,
Antenna Engineering Handbook, 3rd ed., McGraw-Hill, 1993.
Thomas, D. T., “Multiple Beam Synthesis of Low Sidelobe Patterns in Lens Fed Arrays,”
IEEE Transactions on Antennas and Propagation, AP-26, no. 6, November 1978,
pp. 883–86.
Volakis, J. L., Antenna Engineering Handbook, 4th ed., McGraw-Hill, 2007.
370 Electronically Steered Arrays
Antenna Measurements
Every experienced antenna engineer knows that successful antenna design requires a
knowledge of antenna theory, and except for the very simplest antennas, a certain amount
of “cut and try” is necessary before an initial design will perform satisfactorily. An
essential part of this cut-and-try process is measurement of the antenna’s performance.
The majority of antenna measurements lie within two basic categories: impedance
measurements and pattern measurements. The first category deals with one of the most
important antenna parameters—the input impedance. The second category is a very broad
and equally important one, with many subcategories, such as measurements of beam-
width, minor lobe level, gain, and polarization characteristics. Measurements of effi-
ciency and noise may also be desired in some instances.
Not all these possible measurements need be made in every situation. It is seldom that
the complete antenna pattern is measured, including side lobes and polarization charac-
teristics in all directions. Often, at the higher frequencies, it can be assumed that antenna
ohmic losses are negligible, and therefore the radiation efficiency need not be measured.
The input impedance is practically always important, however. The beamwidth, gain,
and side-lobe levels are also usually important, especially at the higher frequencies where
directional antennas are often used. Detailed polarization measurements are important in
special cases. For example, dual transmit and receive polarizations may permit doubling
the number of available communications channels. Measurement of antenna bandwidth
is not actually a separate measurement category; it consists of measuring impedance,
pattern characteristics, and other critical parameters over a band of frequencies.
As discussed in sec. 7.8, the antenna may affect the noise level of a receiving system
in special cases. Consequently, the technique of antenna noise measurements is especially
important in the lower atmospheric noise region of about 1 to 10 GHz. Therefore, the
basic technique of antenna noise measurement and the effect of antenna noise on overall
receiving system noise are described in the present chapter.
371
372 Antenna Measurements
q = 0°
Z
q
Ef
ANTENNA
POSITION
f)
N
TIO
(q,
q = 90°
EC
f = 270° Eq
DIR
EOU q = 90°
ATO Y
R f f = 90°
X
q = 90°
f = 0°
q = 180°
FIGURE 9–1.
Standard Spherical Coordinate System Used in Antenna Measurements. Source (IEEE 1979,
p. 4); © 1979, IEEE.
satisfying this criterion, referred to as far-zone measurements, are discussed in sec. 9.2.
In addition, well-developed measurement methods that permit the use of much shorter
separation distances are described in later sections and include compact ranges (sec. 9.3),
near-field ranges (sec. 9.4), and scale models (sec. 9.6).
A complete pattern measurement consists of measuring field strength and electric field
direction (polarization) for many different values of the angles q and f for all values, in
principle. In practice, the number of q and f directions in which measurements must be
made depends on the complexity of the pattern and the need for detailed information in
the particular application.
Since complete three-dimensional patterns are difficult to plot on a plain sheet of paper,
and since the patterns in particular planes usually provide adequate information, patterns
are usually measured and plotted in planes. Sometimes the pattern in only one plane
conveys the information of AUT interest for a particular antenna. For example, in most
earth-to-earth communication or broadcasting systems, the horizontal-plane pattern is the
major consideration. Vertical-plane patterns in these cases usually have significance only
to the extent that they affect the gain of the antenna in the horizontal plane (i.e., the less
power that is radiated uselessly upward, the more there is available for radiation in the
horizontal plane). Exceptions to this statement include earth-to-earth systems that involve
propagation via an elevated reflection point or medium.
The vertical pattern is often of interest more directly, in applications involving an
earth-based antenna and some elevated receiving point (or vice versa), as in radio
astronomy, space communications, and in radar intended for air or space target detection
and tracking. The patterns in these two planes—horizontal and vertical (azimuth and
elevation)—suffice for practically all applications. The main-lobe pattern in other planes
can usually be adequately estimated from these principal-plane patterns. However, if the
detailed side-lobe patterns are of concern, as they may be in some radar applications and
in other special cases, oblique-plane patterns will be of interest, for the side lobes in these
planes cannot be inferred from the principal-plane patterns.
0
RELATIVE RADIATION INTENSITY IN DB
MAJOR LOBE
–10
–20
VESTIGIAL LOBE WIDE ANGLE
RADIATION
MINOR LOBES REARWARD
–30
RADIATION
–40
–50
20 0 20 40 60 80 100 120 140 160 180
DEGREES OFF AXIS
FIGURE 9–2.
Measured wide-angle far-zone amplitude (in dB) pattern of a reflector antenna. From Cutler,
King, and Kock (1947); © 1947, IEEE.
vestigial lobes occur because of a misaligned feed. Although patterns that include the
major lobe and close-in minor lobes may be calculated, ordinarily it is impossible to
precisely calculate the effects of wide-angle and rearward radiation caused by such
factors as aperture blockage, spillover, and edge diffraction. Obviously, effects of random
mechanical errors may be computed (accounted for) only in a statistical sense.
The radiating near- and far-field patterns of the close-in minor lobes, including the
vestigial lobe, may be calculated following the discussions of Appendix G. In addition,
the accompanying website includes example pattern calculations using Mathcad soft-
ware. These calculations include the more easily defined effects of aperture amplitude
and phase on the major lobe and close-in side lobes. However, measurements are required
in order to determine effects of aperture blocking, feed pattern spillover, edge diffraction,
and reflector surface inaccuracies. In other words, calculations cannot determine all of
the factors that affect side lobes, and thus the final appraisal of antenna performance
usually must be made based on measurements.
Although far zone patterns are nearly always needed, sometimes it is important to
measure and understand field behavior in the radiating near-field. Results of calculations
for the field distributions near an aperture are now discussed, where the phase and ampli-
tude of the field are strongly influenced by distance to the observation point. As a con-
sequence, Figs. 9–3 and 9–4 that follow show major effects of radiation from an aperture,
per se; that is, effects of edge radiation, spillover, and random phase errors are neglected.
Thus, the figures show only how reduced separation may affect amplitudes near the center
of radiation patterns. The calculations were made by following the methodology used in
the supplementary materials SM 4.4 and SM 4.5.
Antenna Patterns, General 375
Amplitude in Decibels
where the observation point is separated –10
a distance K(D2/l) from the aperture –20
center, affects pattern shape. The reader –30
will recall that, from the rule of thumb of –40
sec. 3.2.5, the far-radiating field for elec-
–50
trically large antennas exists where K is
–60
two or greater. The effect of increased –10 –5 0 5 10
phase error, with reduced K (i.e., dis- Angle from Broadside in Degrees
tance), on filling the first pattern null may K approaches infinity
be seen. This null-filling effect due to K=1
K = 0.2
phase error is closely related to the cre-
FIGURE 9–3.
ation of the vestigial lobe of Fig. 9–2.
Figure 9–4 shows an entirely different Calculated radiation for three separation
distances K(2D2/l) versus angle from
aspect of separation distance than does
broadside.
Fig. 9–3. It shows how the amplitude of
the radiated field, near an aperture, varies
1
as a function of position along the
Field Strength (dB)
0.5
aperture. It depicts the near-field region
where radiation approximates a nearly 0
Figure 9-4 shows calculated radiating near fields along a line parallel to and separated
from the aperture by 6 meters. The aperture width and wavelength are 3.25m and 0.1m,
giving about 211m for 2D2/λ. Clearly the 6-meter separation is electromagnetically very
near the aperture. The close-in field within a compact range, called the quiet zone, is the
region that is used to place a test antenna for measuring its radiating far-field patterns
(see sec. 9.3). There are two field strength curves in Fig. 9-4; one for which the element
amplitudes are equal, and the other for a cosine amplitude distribution. Near the aperture
center, the variation in field strength for the cosine tapered aperture is small. There,
however, for the array with equal amplitude elements, the variation is about 2 dB. These
differences in field strength variations near the aperture center underscore the fact that
aperture amplitude tapering of compact range antennas will improve the uniformity of
amplitude within the quiet zone.
It is appropriate to mention that the above-mentioned compact range quiet zone,
although being close to its radiating source, is at a distance great enough so that the
magnitude of the reactive near-field is negligible. In sec. 3.2.5, it is stated that l /2p is
the generally accepted boundary between the reactive and the radiating near-fields for
antennas that are short compared to l, yet there is not a commonly accepted guideline
for larger antennas. According to Yaghjian (1986, p. 33), from experience with near-field
probe measurements on aperture antennas, a distance of “l or so” appears reasonable as
an outer boundary for the reactive near-field. More specifically, according to E. B. Joy,
at a distance of 4 or 5l from a typical antenna under test (AUT), the reactive near-field
is less than 100 dB below the radiating near-field. This guide to relative strengths of
reactive and radiating near-fields was attained through personal communications with E.
B. Joy, and it is reportedly based on studies by him that follow equations (10) and (11)
of Joy and Paris (1972).
The near-field measurement technique (sec. 9.4) is entirely different from a compact
range measurement. It uses a sampling probe to measure the near-field radiation on a
planar, cylindrical, or spherical surface enclosing the antenna-under-test (AUT), and the
results are used to accurately determine far-field patterns of the AUT. To accomplish this,
detailed computer calculations are made based on carefully measured point-to-point
amplitude and phase measurements over the near-field measurement surface. The mea-
surements are made near the aperture, but ordinarily at a great enough distance so that
the reactive near-field is negligible, typically four to five wavelengths from the AUT.
Preferably, the measurements are made with a small sampling probe, to minimize probe-
aperture multiple reflections. In addition, the computer calculations commonly include
corrections to compensate for the probe pattern, gain, and polarization.
°
4°
8°
°
°
0
°
°
8°
4°
0°
0
72
36
36
72
18
14
10
10
14
18
be useful for predicting radio-frequency
interference. This occurs because the FIGURE 9–5.
statistical distributions of antenna pat- 360 degree pattern through azimuth plane of
terns taken over wide angles are nearly AN/SPS-10 antenna. From Cain and Byers
(1968); © 1968, IEEE.
invariant to observation distance, whether
within either the near- or far-fields. In
other words, for a given antenna, the probability of its gain at any angle exceeding a
given value is dominated principally by the radiation patterns over a very wide range of
observation angles, almost exclusive of the major lobe or close-in minor (Johnson 1963;
Long 1960).
From the discussion above, the concept of statistical gain has meaning, if there is a
wide range of unknown but likely observation angles. In other words, the likelihood of
gain exceeding a given level is a useful means for estimating the susceptibility to inter-
ference, especially for narrow-beamwidth antennas. In fact, the statistics of radiation
patterns are useful for predicting near- and far-field interference from antennas, at in-band
or out of-band frequencies, and from antennas in the presence of interfering objects (Cain
and Byers 1969; Cain, Cofer, and Ecker 1970; Cain, Ryan, and Cown 1972; Cain,
Weaver, and Duffy 1974).
Figures 9–5 and 9–6 show results of gain measurements on the AN/SPS-10 radar
antenna (C-band). From Fig. 9–5, the gain versus angle of the antenna is above isotropic
(0 dB) for only about 5.3 degrees. Thus, the probability that the gain is less than isotropic
is 0.98, as illustrated in Fig. 9–6. Similarly, one can see from Fig. 9–5 that the gain is
less than −10 dB for more than half the angles. This is illustrated in Fig. 9–6, where the
probability is 0.74 that the gain is less than −10 dB. Although the peak gain of the antenna
is almost 30 dB, the median gain (the gain at which the cumulative probability is 0.5) is
only −13 dB. Thus, there is a significant difference between median gain and average
gain. (Average gain for a lossless antenna is unity, i.e., 0 dB).
1.0
PROBABILITY THAT THE GAIN IS LESS
0.8
THAN THE ABSCISSA
0.6
0.4
0.2
0.0
–50 –40 –30 –20 –10 0 +10 +20 +30 +40
GAIN (dB)
FIGURE 9–6.
Cumulative gain distribution in azimuth plane for AN/SPS-10 antenna. From Cain and Byers
(1968); © 1968, IEEE.
other receives. Because of the reciprocity principle, the antenna whose pattern is being
measured can be either the transmitting or the receiving member of the pair. The mea-
sured pattern will be the same in either case. In the following discussion the antenna
whose pattern is being measured will be called the antenna under test (AUT), and the
one used as the other terminal of the transmit-receive path will be called the range
antenna, regardless of which one transmits and which receives.
It is essential that the range antenna have the same polarization as the AUT antenna.
Ideally, the polarization of the range antenna should be controllably variable, so that the
pattern of the AUT antenna can be investigated for all polarizations (see Appendix B).
This is especially important in applications where very low side lobes are required,
because side lobes sometime have polarization different from that of the major lobe.
Typically, antennas transmit or receive electromagnetic waves over long distances.
Thus, knowledge of antenna characteristics with antennas in the radiating far-field is
needed, which is accomplished by placing the antennas under test in a plane wave of
uniform amplitude, phase, and polarization. The traditional method employs a free space
range, that is, one for which the separation between the AUT and range antennas meets
or exceeds the far-field requirements previously outlined.
The needed separation distances are generally large and often require the measure-
ments be made in out-of-doors environments. Thus, a major concern is the need to sup-
press unwanted reflections from the ground and nearby surrounding objects, even for
narrow-beam antennas. Typical free-space range geometries minimize effects of unwanted
reflections by using elevated transmit and receive sites, slanting the propagation path, or
by housing the range in an anechoic chamber lined with absorbing material.
Far-Zone Pattern Measurements 379
Hollis, Lyon, and Clayton (1970) and IEEE (1979) include comprehensive discussions
regarding far-zone ranges. Methods that permit smaller separation distances between
antennas and thereby permit indoor pattern measurements include compact (sec. 9.3),
near-field (sec. 9.4), and scale model (sec. 9.6) ranges.
n
tio
ec
dir
am
Be
Antenna
under
test Range
a b
antenna
FIGURE 9–7.
Effect of nearby reflecting object on pattern measurement.
that are in the clear—not too close to large buildings or power and telephone lines, for
instance.
When the pattern is to be measured by rotating the AUT antenna, both antennas should
be located so that they have an unobstructed view of each other, and also have the
required separation to insure a far-field measurement. A further requirement is that the
area between the antennas be clear of sizable reflecting objects, not only in the direct
line between them but for an appreciable distance on both sides. This requirement is
important if an accurate measurement of low-amplitude side lobes is to be made. Figure
9–7 illustrates the reason.
The range antenna is indicated to be the transmitting antenna, which is the customary
arrangement because it permits all the measurements (direction and signal strength) to
be made at a single location, that is, at the AUT antenna. If a large reflecting object, as
shown, is illuminated by the range antenna (whose beamwidth is assumed to be greater
than 2b ), some signal will be reflected toward the AUT antenna, arriving at the angle a
off the in-line direction between the two antennas. (The AUT antenna beamwidth is
assumed to be less than 2a.) This signal will be considerably less than the in-line signal
and will likely not seriously affect the measurement in the main lobe of the AUT antenna,
when the reflected signal will be received in the side-lobe portion of the AUT antenna
pattern. When the AUT antenna is rotated to allow measurement of its side-lobe pattern
at the angle a off axis, however, its main lobe will point directly at the reflecting object,
as shown in the diagram. Then the reflected signal received in this way may be compa-
rable to or even in excess of the in-line signal received in the side lobe, so that a consid-
erably erroneous measurement will result.
This effect is minimized by using a range antenna that is fairly directional, with high
gain in the direction of the AUT antenna and considerably reduced gain—or perhaps
even a null—in the direction of the reflecting object. However, this requires that the range
antenna be quite sizable and expensive. If there is just one major reflecting object in a
Far-Zone Pattern Measurements 381
troublesome position, a null of the range antenna may be directed toward the object, if
this is possible without too greatly reducing the radiation in the desired direction or giving
it an incorrect polarization. As already noted, ideally, the polarization of the range antenna
should be controllably variable, so that the pattern of the AUT antenna can be investigated
for all polarizations. This is because the side lobes may have different polarization
properties than those of the main lobe.
Other possible remedies for the reflecting-object problem exist. One is to interpose
absorbing material (available commercially) between the range antenna and the object,
or between the object and the AUT antenna. Another is to erect a reflecting barrier that
will intercept the radiation going to or from the object and reflect it in some harmless
direction. This barrier is usually a flat sheet of solid metal or mesh material set at an
angle that will direct the reflected waves away from the AUT antenna.
directly between antennas. For this case (Hemming and Heaton 1973), the use of hori-
zontal polarization is preferred to vertical because ground reflections are less sensitive
to the reflection angle at the air-ground interface.
In general, the mean surface of terrain undulates and consequently there may also be
increased beamwidth broadening caused by terrain slope changes at areas of surface
reflections. Then, a simple multipath calculation may not be feasible. When adequate
terrain flatness is not present, a possible solution to the problem is to set up reflecting
“fences” at appropriate points to intercept the waves that would otherwise result in an
interference pattern, and reflect them in harmless directions. Or the terrain may be delib-
erately roughened to such an extent that no appreciable, specular reflection will occur.
Another option is to select a site so that a “valley” exists between the antenna locations;
then, if the range antenna has a fairly narrow vertical-plane beamwidth, there will be no
appreciable reflection from the “floor of the valley.” In an industrial or built-up area this
effect may be achieved by locating the AUT and range antennas on the roofs of tall
buildings separated by relatively clear areas, or on tall towers. In rural areas, two hilltops
may be used.
The foregoing discussion has been concerned with horizontal (i.e., azimuth) plane
pattern measurement. Elevation-angle patterns are virtually impossible to measure in the
presence of a reflection-interference lobe structure. One method of circumventing this
difficulty is to turn the AUT antenna on its side so that its “vertical plane” pattern can
actually be measured in a horizontal plane. The vertical patterns of antennas that can be
tilted upward can be measured from a helicopter or an airplane, if some method of mea-
suring the aircraft direction angle and range can be provided. Use of a precision tracking
radar or a global position system (GPS) may be ideal for this purpose.
is used to minimize diffraction from the reflector edge for improving the uniformity of
the field across the aperture. Shaped serrated reflector edges are used that assisted in
providing quiet zone performance suitable for measuring antenna/vehicle radiation with
vehicle dimensions up to fifty feet (Joy and Wilson 1987).
Another major design problem for large compact ranges, when the test items are heavy,
is the AUT positioner, which is located approximately above the focal point. With a
compact range, the item being measured must be within the quiet zone, which is centered
roughly 40 feet (12.2 m) above ground. A positioner had to be designed and built, because
a suitable one was not available commercially. The positioner, with its base on the
ground, is designed to handle weights up to 140,000 pounds (69,300 kg), at heights
exceeding 40 feet, and move the test item from −1 to +91 degrees in elevation and con-
tinuously in azimuth.
A relatively new electronic timing means appears to have much promise for improving
quiet zone performance, by reducing the reflections from the compact range reflector
edges and from surrounding objects. The technique is, in essence, a short-pulse-length
radar range gating technique (Chang, Liao, and Wu 2004). Although additional electronic
instrumentation is required, radar range gating (also called impulse time-domain mea-
surements) may offer improved quiet zone performance as well as simplifications and
cost reductions in the traditionally used methods of reflector edge treatment and anechoic
chamber construction.
Z
9.4.3. Spherical Near-Field
Measurements
Spherical near-field measurements -X
(Hansen 1988; Leach and Paris 1973)
measure the near-field on the surface of a FIGURE 9–14.
sphere that surrounds the antenna as illus- Typical cylindrical near-field measured
trated in Figure 9–16. Data is collected at geometry (courtesy MI Technologies).
equal increments in q and j (the standard
spherical coordinate angles). This data is
then probe corrected and transformed to the far-field to obtain the far-field of the antenna
under test. Any antenna can be measured with a spherical system (e.g., a low-gain, broad-
beam antenna as well as fan-beam and pencil-beam antennas) since all of the energy
388 Antenna Measurements
9.6. Scale-Model Z
Measurements
Because of the practical difficulties of
-X
making measurements on large antennas,
and the expense of constructing experi-
mental designs of large size, the scale-
model technique is very useful in antenna FIGURE 9–16.
development and research work. The Typical spherical near-field measured geometry
principle of the scale model is that if two (courtesy MI Technologies).
antennas are of exactly the same shape
but differ in size by a scale factor S, their
electromagnetic behavior will be the same
if the smaller antenna is operated at a
frequency S times as great as that at which
the larger antenna is operated.
This principle is self-evident for simple
antennas such as a half-wave dipole, but
in fact it holds for all types of antennas.
For example, if two paraboloidal reflec-
tors are of exactly the same form (same
f/D ratio and same aperture shape) but one
is ten times the size of the other, and if
they have identically shaped feeds that
also differ in size by a 10 : 1 ratio, then if
the larger one is operated at 1 GHz and
the smaller one at 10 GHz, their operating
parameters will be identical. The gains,
beamwidths, side-lobe levels, and input
impedances will be the same; in fact the
patterns will be exactly alike. Also, the
percentage bandwidths will be the same.
The same would be true of a third antenna
five times as large as the larger of the first
two, operated at 200 MHz. Thus a large FIGURE 9–17.
antenna can be scaled down, or modeled, Spherical near-field antenna measurement
by any desired factor by operating it at a system employing an arch and azimuth
sufficiently high frequency. turntable (courtesy MI Technologies).
390 Antenna Measurements
Frequencies in the region of 10 GHz are very popular for scale-model work, because
the scaled antennas in this region (known as X-band) are of very manageable size,
usually, and also the necessary apparatus is readily available at this frequency: wave-
guides and fittings, signal sources, receivers, and special measurement equipment.
However, some antennas if scaled to so high a frequency would be too tiny and delicate
for practical handling. For example, a half-wave dipole is only about 0.6 inch long.
Another popular frequency region, when an X-band model would be too small, is S-band
(the region of frequency around 3 GHz).
The scale-model principle is exact if the scaling is exact. Not only the size, but also
the conductivity of the model must be scaled, for exact results. The conductivity of the
model must be greater than that of the full-sized antenna by the factor S. However, if
the conductivities of both antennas are quite high, as is usually the case, and if the con-
ductors of the model do not become too minute, the conductivity scaling is not of much
importance. Models are sometimes silver-plated to increase their conductivity. If parts
of the antenna involve dielectric or magnetic materials, the sizes of these parts are scaled,
but the dielectric constants and the permeabilities are not; they should be of the same
values in both the model and the full-sized antenna; however, their losses, which often
depend on conductivity, must be scaled.
The exactness of scaling of the linear dimensions of the antenna is the most important
scaling factor. The scaling requirement applies to all dimensions, such as the thickness of
the walls of a waveguide horn, the openings of the mesh of a reflector surface, and even the
deviations from a true paraboloidal surface—all of these must be smaller in the model by
the factor 1/S, if the results are to be completely accurate. On the other hand, if only approxi-
mate results are needed, the scaling need not be applied to some of the details of the antenna,
if they are very small compared with the wavelength. Inexact scaling affects the accuracy
of impedance measurements more than it does pattern measurements and gain.
An advantage of the use of scale models is that it reduces not only the size of the
antenna but also the required size of a pattern range, or separation of AUT and range
antennas for any far-field measurements. This can be seen by considering the separation
criterion 2D2/l. Since D is decreased by the scale factor 1/S, and l is reduced by the
same factor, the net effect is to reduce the quantity 2D2/l by the factor 1/S.
Because of this reduction scale-model pattern and gain measurements are sometimes
made in an anechoic chamber, which is an enclosure whose walls, floors, and ceilings
are lined with a special material that absorbs radio waves almost completely. Thus no
reflections occur, and the room behaves as if it were free space. To make an enclosure
of this type of suitable dimensions for use with full-scale antennas in the VHF and UHF
regions would be prohibitively expensive. But for model antennas at S-band or X-band
or small full-sized antennas it becomes quite practical. The absorbent material is com-
mercially available.
SIGNAL
TEST SOURCE SOURCE
ANTENNA ANTENNA
TEST
POLARIZATION
POSITIONER
POSITIONER
SOURCE
TOWER
PATTERN
RECORDER
FIGURE 9–18.
Block Diagram of a Typical Antenna Measurement System. Source (IEEE 1979, p. 19);
© 1979, IEEE.
f ROTATIONAL AXIS
q ROTATIONAL
A AXIS
q ROTATIONAL Z
AXIS X f ROTATIONAL
AXIS
q f
O O
Y Z
A R S q
f
S
X
Y
Z A
f ROTATIONAL q ROTATIONAL
q ROTATIONAL q AXIS AXIS
Y O f ROTATIONAL
AXIS O S AXIS
X
A Z
S
(a) (b)
FIGURE 9–19.
Antenna Positioners: (a) azimuth-over-elevation positioner (b) elevation-over-azimuth
positioner. Source (IEEE 1979, p. 25); © 1979, IEEE.
printed recordings. Even complete, or near complete, measurement systems that include
a central control console may be purchased. Usually, personal computers are used to
control the automated measurement system and perform needed calculations.
The AUT positioner, its control system, and its position indicators constitute a key
subsystem, which must be selected based on the sizes and weights of the antennas to be
measured and on the types of tests to be made. The positioner must of course have the
needed movements and yet to be highly stable, handle the requisite weights, and with-
stand prevalent environmental conditions. Additionally, the position controls and indica-
tors must be accurate and reliable. A wide variety of positioner types, along with their
position controls and position indicators, are commercially available. Perhaps the sim-
plest positioners are those that rotate continuously in either azimuth or in elevation.
Figure 9–19 includes sketches of two often-used movements, namely azimuth-over-
elevation and elevation-over-azimuth.
Positioners are commercially available in various sizes and weight-handling capaci-
ties. Figure 9–20 shows a large positioner that includes a bearing having a diameter of
85 inches (about 2.2 m). According to Gerald Hickman, formerly of Scientific Atlanta,
this model was used for positioning an F-111 aircraft. Smaller, conceptually similar
positioners having platform diameters as small as about 6 inches (15 cm) are also
commercially available.
Directivity and Gain Measurements 393
Therefore,
E2
4π R 2 E 2 4π R 2 pantenna
G = 377 = = (9–2)
Pin 377 Pin Pin
4π R 2
Efficiency k is Pout/Pin, and thus from (9–2) the directivity can be expressed as
4π R 2 E 2 4π R 2 pantenna
D= = (9–3)
377 Pout Pout
4π R 2 E 2 4π R 2 pantenna
G= = (9–4)
377 Pin F 2 Pin F 2
Pr 4π Pr 4π Pr
pr = = = (9–5)
Ar kDr λ 2
Gr λ 2
where Pr, Dr, and Gr refer to the received power, directivity and gain of the receiving
antenna. (For a review of receiving cross section area, see chapter 3, sec. 3.4).
G2λ 2
Pr = Pin F2
(4π R ) 2
4π R Pr
G= (9–6)
λF Pin
This procedure is likely to be most accurate when the propagation factor F is unity,
that is, under effectively free-space conditions with minimal earth-reflection interference
effects. Equation (9–6) can also be applied successfully under conditions that permit
accurate calculation of F, for example, when reflection occurs from a smooth water
surface located between the two antennas.
The gain calibration procedure described above, that uses (9–6), is known as the two-
antenna method. This method and the three-antenna method, to now be described, are
usually performed at a site where the ground and other reflections are negligible, so that
F is assumed to be unity. For the three-antenna method, three sets of two-antenna mea-
surements are performed with three antennas that may have different gains, designated
here as GA, GB, and GC. Then, by expressing the Friis equation in logarithmic form and
setting F = 1, three simultaneous linear equations are obtained that follow:
4π R ⎞ ⎛P ⎞
10 log G A + 10 log GB = 20 log ⎛⎜ ⎟ + 10 log ⎜ r ⎟
⎝ λ ⎠ ⎝ Pin ⎠ AB
396 Antenna Measurements
4π R ⎞ ⎛P ⎞
10 log G A + 10 log GC = 20 log ⎛⎜ + 10 log ⎜ r ⎟
⎝ λ ⎟⎠ ⎝ Pin ⎠ AC
4π R ⎞ ⎛P ⎞
10 log GB + 10 log GC = 20 log ⎛⎜ ⎟ + 10 log ⎜ r ⎟
⎝ λ ⎠ ⎝ Pin ⎠ BC
Thus, the three gains may be determined from the above simultaneous equations.
Since the gain of the unknown antenna is ordinarily higher than the gain of the gain
standard, the standard antenna is first connected to the receiver (in the manner of an AUT
antenna in pattern measurement), and aimed at the range antenna. The receiver gain is
adjusted to give a convenient output meter indication (after it is determined that the level
of the signal is well above noise level). Then the antenna whose gain is to be measured
is connected in place of the standard-gain antenna, and attenuation is introduced into the
transmission line between the antenna and receiver until the output indication is the same
as it was with the gain-standard antenna. If the attenuation factor, expressed as a power
ratio greater than one, is L, the gain of the unknown antenna, Ga, is
Ga = LGs (9–7)
where Gs is the standard-antenna gain. Inasmuch as antenna gains and attenuator calibra-
tions are often expressed in decibels, it is frequently convenient to make the calculation
in decibels, in which multiplication is replaced by addition:
In the unlikely event that the unknown antenna has a smaller gain than the standard, L
in (9–7) is expressed as a number less than one, and the decibel value of L in (9–8) is
negative.) The basic set-up for gain measurement by the comparison method is dia-
grammed in Fig. 9–21. It is essential in this method of gain measurement that both the
AUT and the standard antenna are equivalently impedance-matched to the load presented
to them by the transmission line. The best way to insure this is to make VSWR measure-
ments with each of them connected in turn, and adjust the matching of each for a flat
line (VSWR = 1). This method basically compares antenna power gains, but directivities
may also be determined if the antenna radiation-efficiency factors are known.
Standard-gain horn
Signal
Output source
Receiver
meter
Attenuator
FIGURE 9–21.
Set-up for gain measurement by comparison method.
398 Antenna Measurements
PR
kR = (9–9)
Pin
Pin − PD
kR = (9–10)
Pin
G
kR = (9–11)
D
The first equation requires direct measurement of the total radiated power, which is pos-
sible only in special cases. Measurement of the total input power to the antenna Pin is
not difficult, since this power flows in the transmission line connecting the transmitter
to the antenna.
The second equation requires measurement of the dissipated power PD. This can
sometimes be done, especially at low frequencies, by measuring the resistance of conduc-
tors in which current flows, and multiplying these resistances by the square of the current.
At LF and VLF, the components of the antenna system whose rf resistances must be
measured are the ground system, the loading coil, and the conductors of the antenna
itself.
The third equation is not directly useful, but it may be combined with (9–4) to
obtain
4π R 2 E 2
kR = (9–12)
377 Pin F 2 D
This equation is especially useful at VLF with short vertical grounded radiators (mono-
poles). For these antennas, D may be taken as twice the directivity of the corresponding
short dipole; hence, ordinarily, D is approximately 3. Within line-of-sight distances and
over water or moist ground, it can be assumed that F = 1. Therefore, if measurements
Radiation Resistance 399
are available of both the total input power to the antenna Pin and field strength E at dis-
tance R from the antenna, the radiation efficiency can be determined. Since the definition
of kR requires E to be the far-field strength, R must be a distance satisfying the far-field
criteria, as discussed in sec. 3.2.5.
4π A
D = k A 2 (9–13)
λ
where A is the geometric area of the aperture (for antennas that have a physical area such
as those listed immediately above) and l is the wavelength (both expressed in the same
system of units). In this context, kA is called the aperture efficiency. If the field intensity
over the aperture of an antenna is uniform (no “taper”), with no spillover or leakage,
kA = 1. This is the largest value of kA practically attainable, ruling out “supergain” anten-
nas, which are not feasible except to a limited degree in a few cases. Typical values of
kA range from somewhat less than 0.5 to nearly 1.0. The measurement of kA is virtually
synonymous with measurement of directivity D, since from (9–13):
Dλ 2
kA = (9–14)
4π A
2
I feedpoint
Rr (loop) = Rr (feedpoint ) (9–15)
I maximum
where I denotes the currents at the points indicated by the subscript notation.
If there is appreciable ohmic loss, so that the antenna radiation efficiency factor kR is
less than one, the radiation resistance is found from equations (3–26) and (3–29) of
chapter 3 to be
Rr = k R Ri (9–16)
where Ri is the input resistance, that is, the resistive component of the input
impedance.
It is apparent that the radiation resistance is sometimes a rather nebulous concept and
not always easily measured. In general, it is a useful concept only when it is readily
measurable. It has no meaning for antennas in which there is no clearly defined current
value to which it can be referred.
slotted line was commercially available for measuring standing wave ratios within coaxial
transmission lines or waveguides, up to frequencies of about 100 GHz. As already noted,
today the preferred impedance measurement method is the network analyzer, for which
multiband designs are commercially available for almost all radio frequencies, from
10 kHz to 100 GHz. The basics of network analyzers are given in sec. 9.11.2. However,
the standing wave method is first discussed in sec. 9.11.1, because its principles are basic
to the understanding of power transfer to antennas and their related components.
Voltage
Vmax V pattern
Vmin
0
λ d
Source of Transmission
voltage Line conductors ZL line
FIGURE 9–22.
Diagram showing quantities to be measured in standing-wave method of impedance
determination.
402 Antenna Measurements
The basic procedure for measuring the position of the voltage minimum, and the
VSWR, is to move a short sensing probe along a transmission line and find the positions
of maximum and minimum field strengths. To meet these requirements, it is possible to
employ a special slotted line—a horizontal section of coaxial line or waveguide that has
a long narrow slot cut into the top of its outer wall, along its length. A short metal pin,
the sensing probe, projects through the center of the slot and into the coaxial line or
waveguide (without touching the walls). The upper end of the probe connects to a detec-
tor (usually a crystal diode). Because the probe is of such small dimensions, it does
not disturb the field appreciably, and the detector does not constitute an appreciable
load.
Prior to about 1975, slotted lines were commercially available over a wide range of
frequencies, but because of the development of highly successful network analyzers, they
seem to have become almost extinct. Even so, the slotted section remains a low-cost
instrument for making VSWR and impedance measurements. For frequencies of several
hundred megahertz up to perhaps 3 GHz, coaxial slotted lines of 50-ohm characteristic
impedance were standard. At higher frequencies and up to about 100 GHz, slotted wave-
guides were used. Another method of obtaining VSWR is to use a directional coupler,
such as the one represented in Fig. 2–19, chapter 2. If the voltage outputs of both ports
2 and 4 of this coupler are separately measured, one measurement will represent the
amplitude of the forward-traveling wave in the line, and the other will represent the
reflected-wave amplitude. The ratio of the sum of these amplitudes to their difference
gives the VSWR, in accordance with equation (2–43), chapter 2. A measurement setup
that uses a directional coupler in this way is called a reflectometer. If also the phase dif-
ference of the two voltages is measured and the distance from the coupler to the load is
known, the load impedance can be computed.
A R B
DEVICE
FREQUENCY
UNDER LOAD
SOURCE
TEST
A R B
RECEIVER
PROCESSOR/
DISPLAY
FIGURE 9–23.
Schematic of a basic network analyzer.
the simple single-frequency network analyzer that compared incident and reflected waves
within the transmission line and “automatically” provided calculated impedance results.
This automation began in the late 1960s, and rapidly evolved with the subsequent devel-
opment of tunable, wideband, solid-state sources. The technology of network analyzers
expanded from the early, single low-radio-frequency input and transfer impedance mea-
surements to the present state-of-art whereby amplitude and phase determinations are
made from the lower radio frequencies up to the microwave and millimeter wavelength
regions.
Features of modern network analyzers include
• rapid tunable, wideband solid-state signal sources
• rapid-phase and amplitude measurements
• automatic measurements and computer calculations using the measured
results
• multiple visual displays of operating parameters and measured and calculated
results.
family go through it vertically. The first X/Z0 = –0.5 X/Z = 0 X/Z0 = +0.5
0 R/Z0 = 0.33
of these families of circles represent con-
stant values of the ratio RL/Z0, and will Center of R/Z = 1
0
be referred to as “R circles.” The second X/Z0 = –1 chart
X/Z0 = +1
family of circles corresponds to constant
values of XL/Z0 and will be referred to as R/Z0 = 3
the “X circles.” RL and XL are of course X/Z0 = –2 X/Z0 = +2
the resistive and reactive components
of the load impedance, ZL. The X circles
to the left of the center line are negative
FIGURE 9–24.
values of XL/Z0, representing capacitive
reactance, and those on the right are posi- Basic construction of the Smith Chart.
tive, representing inductive reactance.
The vertical center line is the XL = 0 line. The R circle that passes through the exact
center of the chart represents RL/Z0 = 1. Therefore the exact center point of the chart
corresponds to a load impedance that is a pure resistance of value equal to the charac-
teristic impedance; that is, at this point RL = Z0 and XL = 0. (This is of course the load
value that results in a unity value of VSWR.)
These two families of circles are in effect a system of coordinates, one coordinate set
representing the resistive component and the other set the reactive component of the load
impedance. Any particular point on the chart corresponds to a load impedance, ZL, whose
components are given by the two orthogonal R and X circles that intersect at that
point.
In addition to these R and X coordinates, there is another set of coordinates for the
measured quantities, VSWR and d/l. These coordinates are not printed on the chart, since
they would result in a hodgepodge of lines and make it difficult to read the chart. Instead,
they are to be plotted in by the user for the specific measured values in a particular case.
The VSWR coordinates are circles whose centers are at the center of the Smith Chart—
that is, at the XL/Z0 = 0, RL/Z0 = 1 point. This point corresponds to VSWR = 1, and the
circles of increasing size correspond to increasing values of VSWR. The largest of these
circles forms the outer boundary of the chart and represents an infinite VSWR (∞).
The d/l coordinates are radial lines emanating from the center of the chart. A circular
scale of values of d/l is provided on the outer periphery of the chart. The full circle spans
the range from d/l = 0, at the top of the chart, through d/l = 0.25 at the bottom of the
chart, to d/l = 0.5 again at the top; thus the complete circle of values corresponds to
values of d going from zero to 12 wavelength. The values increase counterclockwise,
when d is the distance from the antenna terminals to the first voltage minimum. This
direction on the scale is usually marked “wavelengths toward the load,” which refers to
the location of the null when the load is short-circuited, with respect to the voltage
minimum with the short removed. A complementary scale is also usually provided,
marked “wavelengths toward the generator.” This scale increases in the opposite direction
and corresponds to the distance from the voltage minimum (short removed) to the nearest
null (load shorted) in the direction of the generator (signal source).
406 Antenna Measurements
FIGURE 9–25.
Example of impedance and admittance calculation, using the Smith Chart.
An example of the use of the Smith Chart is shown in Fig. 9–25. In this example it
is supposed that the VSWR has been measured and found to be 5.3. Therefore a circle
has been drawn, with its center at the center of the Smith Chart, passing through the
VSWR = 5.3 point on the vertical center line. The d/l value has been found to be 0.16,
and a radial line has been drawn from the center of the chart to this point on the d/l scale
(the one labeled “wavelengths toward load”). At the intersection of this circle and radial
line, a pair of the R and X circles (shown dashed) also intersect. These are seen to be the
ones for RL/Z0 = 0.6, and XL/Z0 = −1.4. Therefore, the antenna input impedance in this
case consists of a resistive component RL = 0.6Z0, and a negative (capacitive) reactance
component XL = −1.4Z0. If (for example) the characteristic impedance of the line is
Antenna Noise Measurement 407
50 ohms, then RL = 30 ohms, and XL = −70 ohms. (Note that XL here designates “load
reactance” and not, as in some cases, “inductive reactance.”)
This example illustrates the basic use of the Smith Chart. It can also be used to solve
impedance-matching problems. For this purpose it is often convenient to work with
admittance, rather than impedance. As discussed in chapter 2 using equations (2–26)
through (2–34), admittance Y is the reciprocal of impedance. Its real and imaginary
components are called, respectively, conductance G and susceptance B. To find these
admittance components, in terms of their ratio to the characteristic admittance of the line
Y0 (= 1/Z0), is extremely simple using the Smith Chart. The radial line for the applicable
value of d/l is merely extended (as shown dashed in Fig. 9–25) until it intercepts the
opposite side of the VSWR circle. The values of GL/Y0 and BL/Y0 are then read at this
point, from the same set of orthogonal circles that represent RL/Z0 and XL/Z0 at the inter-
section of the solid radial d/l line with the VSWR circle. In this particular example it is
found that GL/Y0 = 0.26 and BL/Y0 = +0.6. Since Y0 = 1/Z0, if Z0 = 50 ohms then Y0 =
0.02 mho. Consequently GL = 0.26 × 0.02 = 0.0052 mho, and BL = +0.6 × 0.02 =
+0.012 mho, in this example. (Note that whereas capacitive reactance is negative, capaci-
tive susceptance is positive; inductive susceptance is negative.)
* Parenthetically, absolute temperatures are correctly expressed in Kelvins and not in degrees Kelvin.
Following IEEE (1992, p. 320), the Kelvin or absolute temperature scale is related to the Celsuis and
Fahrenheit scales by the formula:
an approximation that holds for most conditions at frequencies in the radio spectrum. A
more exact formula is required at extremely high frequencies and at extremely low
temperatures.)
It is well known that the maximum power that can be delivered by a generator or other
source having a given internal electromotive force (open-circuit voltage) is obtained
when the external load resistance is equal to the internal resistance of the source. (More
generally, the external impedance must be equal to the complex conjugate of the source
impedance.) From this it readily follows that the thermal noise power deliverable to this
optimum load by a thermal noise source of resistance R at temperature T is
Vn2
Pn = = kTB watts (9–18)
4R
Hence the quantity kTB is the available noise power. It does not depend on the particular
value of resistance, but only on the absolute temperature T (for a given bandwidth B).
Therefore it has become customary to rate a noise source whose available noise power
is Pn in terms of its equivalent noise temperature calculated from (9–18)—that is,
T = Pn/kB. This rating practice is followed even where the noise is not actually of thermal
origin, provided it “resembles” thermal noise in statistical and frequency spectrum
characteristics.
A thermal-noise voltage is a randomly fluctuating voltage that conforms to a particular
statistical law and contains a uniform mixture of all frequencies within the radio spec-
trum—or, in particular, within any selected band, such as the intermediate frequency (i–f)
passband of a receiver. The general appearance of a noise-voltage waveform, as viewed
on an oscilloscope connected to the output of a receiver (after detection), is shown in
Fig. 9–26. This waveform is the half-wave-rectified “envelope” of the rf noise; the rf
cycles have been filtered out. (Such noise is sometimes called “grass.”)
Because of their statistically random nature, noise voltages from two different (inde-
pendent) sources are uncorrelated. Thus, when combined or superimposed in a circuit or
as electromagnetic waves in space, they add so that the resultant rms voltage (or field
intensity) is the square root of the sum of the squares (root-sum-square) of the super-
imposed voltages (or fields). Consequently, the resultant noise power is the sum of
the individual-component noise powers; it is unnecessary to consider whether phase
relationships cause addition, cancellation,
or some intermediate result. This deduc-
tion follows from the fact that the relative
Voltage
The mathematical analysis of noise, the interrelationships of signals and noise, and
the physical origins of noise have been intensively investigated, and an extensive litera-
ture exists. See, for example, Blake (1986, ch. 4) or Davenport and Root (1987).
Sun
Galaxy
Main beam
Side and
back lobes
lo
no
sp
Beam-elevation
he
re
angle
Antenna
Tr
op
os
ph
er
e
Earth
FIGURE 9–27.
External sources of antenna noise.
Techniques q = 0°
q = 1°
1,000 q = 2°
As in antenna gain measurement, the q = 5°
q = 10°
most convenient technique of measure- q = 30° 0°
q = 90°
ment of noise temperature is comparison 100
90°
with a source of known (and variable)
noise temperature. Figure 9–29 shows the 10
measurement set-up.
The switch in this arrangement must 1
be a high-quality transmission-line 100 1,000 10,000 100,000
switch—a coaxial switch if the line is FREQUENCY (MHz)
coaxial, or a waveguide switch when the FIGURE 9–28.
line is waveguide. It is first set to connect
Noise temperature of an idealized antenna
the receiver input to the antenna, and the located on the earth’s surface, as a function of
receiver gain is adjusted to cause the frequency and beam elevation angle q (from
output meter to read a convenient value. Blake 1969).
Antenna Noise Measurement 411
Psource = kTs B ⎛⎜ ⎞⎟
1
(9–19)
⎝ L⎠
Now let Ti be the attenuator noise temperature, and thus kTiB is the attenuator noise
power at the attenuator input. Recall that the available power from separate noise sources
412 Antenna Measurements
is the sum of the noise powers. Thus, the total noise power at the attenuator input is
kTsB + kTiB, and the total noise power Pn(total) at the attenuator output is
1
Pn(total ) = [ kTs B + kTi B ] (9–20)
L
Ti = Tt ( L − 1) (9–21)
Note that (9–21) was found true regardless of the magnitudes of either Ts and Tt,
providing only that Ts = Tt. In fact, Dicke et al. (1946) report that noise power
created internally by an attenuator is always equal to (9–21). Consequently, by replacing
Ti in (9–20) with Tt(L − 1), one finds that attenuator output power versus Ts, Tt, and
L is
= kB ⎡⎢ s + Tt ⎛⎜ 1 − ⎞⎟ ⎤⎥
1 T 1
Pn(total ) = [ kTs B + kTt ( L − 1) B ] (9–22)
L ⎣L ⎝ L⎠⎦
Then, from the relationship Pn(total) = kBTn with Tn being the effective noise tempera-
ture at the attenuator output terminals, one finds
+ Tt ⎛⎜ 1 − ⎞⎟
Ts 1
Tn = (9–23)
L ⎝ L⎠
When L approaches infinity, it can be seen from (9–23) that Tn = Tt, whereas if
L = 1, Tn = Ts. Therefore, by varying L effective noise temperatures in the approximate
range Ts to Tt are available by the method of Fig. 9–29.
If the attenuator is calibrated in decibels, and the decibel attenuation expressed as a
positive number is AdB, the value of L is given by
L = anti log ⎛⎜ dB ⎞⎟
A
(9–24)
⎝ 10 ⎠
The standard reference temperature for noise measurements is 290 K (IEEE, 1992, p.
850), which is a chilly “room temperature” of 16.8° C (62.3° F). The method (Fig. 9–30)
of obtaining variable output noise temperature will yield values that range from Ts up
to or down to 290 K. This depends on whether Ts is less than or greater than 290 K.
Antenna Noise Measurement 413
Ordinarily Ts is less than 290 K, if it is presumed that the antenna noise temperature a
low-noise antenna is to be measured.
Tp + Te Tp + Te
R= =
Ta + Te Top
with
Tp = physical temperature K of ambient termination
Te = receiver effective noise temperature K at its input
Ta = antenna noise temperature K
The ratio R is determined by means of a precision attenuator located in the receiver IF
stage, which is positioned remotely enough that a change in attenuator settings does not
influence the effective receiver input temperature Te. Another readily met requirement is
that the receiver response be linear between its input and the attenuator.
The receiving operating noise temperature Top from the equation for R above is
Tp + Te
Top = (9–25)
R
Then, Top is calculated by using (9–25) with the measured a value of R, measured
ambient temperature Tp, and the effective receiver noise temperature Te that is either
measured or estimated. Notice that if Te << Tp, Te may not need to be known
accurately.
Here, as used previously, B is the noise bandwidth, G is the power gain of the device
being evaluated, and Ti is the effective noise temperature of its internally generated noise
(referred to the device input). Parenthetically, with T0 being 290 K, the product kT0 is
System Noise Calculations 415
Ti = T0 ( NF − 1) = 290 ( NF − 1) (9–27)
Thus, (9–27) gives the internally generated noise temperature Ti of a device when referred
to its input in terms of its noise figure NF.
An example is now given on determining the output noise power Pout of a transistor
amplifier having NF of 2 dB (NF = 1.58) and an input source with noise temperature Ts
of 200 K. Then, from (9–27) Ti is 290(1.58 − 1) or 168.2 K. Thus, the noise temperature,
referred to the device input, is Ti + Ts or 368.2 K. Now assume the noise bandwidth
B and gain G are one MHz and 30 dB (1,000), respectively. Thus, the average noise
power output Pout of the transistor amplifier with input source temperature Ts of 200 K
becomes
N = kTsys B (9–28)
where
k = 1.38 × 10−23 watt-sec/Kelvin (Boltzmann’s constant)
Tsys = system noise temperature in Kelvins
B = receiver noise bandwidth
Then, if the net available power gain between the antenna output and the receiver output
is G, the actual noise-power of the receiver output is G kTsysB. Thus it is important to
state what reference point has been chosen when the concept of system noise temperature
is employed.
416 Antenna Measurements
In general, the effect of a noise source at one point in the system may be referred to
another point by properly multiplying or dividing the source noise power by the net
power gain of the components between the two points. Then, when both signal and noise
powers are referred to the same point in a system, their ratio is the same as the output
signal-to-noise ratio, provided that the gain is the same for signal and noise.
Following Blake (1961), the system noise temperature Tsys of a system that includes
antenna, transmission line, and the receiver is
Tsys = Ta + Tr ( Lr − 1) + Lr Te (9–29)
where
It is important to note that in (9–29), Ta, Tr(Lr − 1), and LrTe are each equivalent noise
temperatures referenced to the antenna output terminals. The reader will recognize that
a transmission line is a fixed attenuator. Therefore, the noise temperature created by the
transmission line and referenced to its input (antenna output) may be calculated using
(9–27) in sec. 9.14.1. The terms Te and LrTe denote the internally generated receiver noise
referenced to the receiver input and to the antenna output, respectively. The factor Lr
accounts for the transmission line attenuation between the receiver input and the antenna
output. In other words, since Lr ≥ 1, the noise temperature of the receiver referenced to
the antenna output LrTe exceeds the internally generated receiver noise temperature Te
referenced to the receiver input.
Alternatively, system noise temperature Tsys can be expressed in terms of receiver noise
figure NF, by substituting (9–27) into (9–29). Then, Tsys becomes
Sometimes the term system noise figure NFsys is used to express the performance of
an overall receiving system. In that case, the system noise temperature and the system
noise figure are related as follows:
Table 9-1 below gives calculated noise temperatures when using the assumptions
300 K transmission line thermal temperature, plus low and moderate values of transmis-
sion line losses Lr and receiver noise figures NF. Note that signal-to-noise ratio (SNR)
is proportional to system noise temperature Tsys. Therefore, significant changes in signal
detection performance require substantial changes in the ratios of Tsys, system noise
temperature. Note that changes in the antenna noise temperature Ta will result in signifi-
cant improvement in signal-to-noise only if both Lr and NF are very small. From Table
9-1, it is apparent that small values of Tsys require small values of both Lr and NF. Notice
that as Lr is enlarged, the noise contributions from both the transmission line and from
the receiver are increased. It is thus doubly important to place the receiver as close to
the antenna output as practicable, for minimizing Lr and to thereby improve the system’s
signal detection performance.
References
Blake, L. V., “Antenna and Receiving System Noise-Temperature Calculation,”
Proceedings of IEEE, vol. 49, October 1961, pp. 1568–69.
Blake, L. V., “A Guide to Basic Pulse-Radar Maximum-Range Calculation,” Naval
Research Laboratory 6930, December 23, 1969; also see L. V. Blake, ch. 2 in M. I.
Skolnik (ed.), Radar Handbook, 2nd ed., McGraw-Hill, 1990, p. 2.29.
Blake, L. V., ch. 4, Radar Range-Performance Analysis, Artech House, 1986; or ch. 4,
Munro Publishing, 1991.
Bodnar, D. G., “Feed Study for USAEPG Compact Range,” Final Technical Report,
Project A-3922, Georgia Tech Research Institute, August 1986.
Cain, F. L., and K. G. Byers, Jr., “Statistical Gain Characteristics of Radar Antennas at
Very Short Fresnel Zone Distances,” IEEE Transactions on Electromagnetic
Compatibility, EMC-11, February 1969; pp. 1–9.
Cain, F. L., and K. G. Byers, Jr., “Relations of Site Effects to Statistical Gain
Characteristics of Radar Antennas,” 1968 IEEE Electromagnetic Compatibility
Symposium Record, pp. 339–48.
418 Antenna Measurements
Johnson, R. C., “Mutual Gain of Radar Search Antennas,” Proceedings of the Ninth Tri-
Service Conference on Electromagnetic Compatibility, IIT Research Institute, Chicago,
Illinois, October 1963.
Johnson, R. C., “Radar Search Antennas and RFI,” IEEE Transactions on Electromagnetic
Compatibility, EMC-6, July 1964, pp. 1–8.
Johnson, R. C., “Antenna Range for Providing a Plane Wave for Antenna Measurements,”
U. S. Patent 3,302,205, January 31, 1967.
Johnson, R. C., H. A. Ecker, and J. S. Hollis, “Determination of Far-Field Antenna Patterns
from Near-Field Measurements,” Proceedings of the IEEE, December 1973;
pp. 1668–94.
Johnson, R. C., H. A. Ecker, and R. A. Moore, “Compact Range Techniques and
Measurements,” IEEE Transactions on Antennas and Propagation, September 1969,
pp. 568–76.
Joy, E. B., and D. T. Paris, “Spatial sampling and filtering in near-field Measurements,”
IEEE Transactions on Antennas and Propagation, vol. AP-20, May 1972,
pp. 253–61.
Joy, E. B., and R. E. Wilson, “Shaped Edge Serrations for Improved Compact Range
Performance,” Proceedings of the 1987 AMTA Symposium, Seattle, Washington.
Kerns, D. M., “Plane-Wave Scattering-Matrix Theory of Antennas and Antenna-Antenna
Interactions,” NBS Monograph 162, National Bureau of Standards, Boulder, CO, June
1981.
Kraus, J. D., Antennas, 2nd Edition, McGraw-Hill, 1988.
Ko, H. C., “The Distribution of Cosmic Background Radiation,” Proceedings of the IRE,
46, No. 1, January 1958, pp. 208–15.
Leach, W. M., Jr., and D. T. Paris, “Probe Compensated Near-Field Measurements on A
Cylinder,” IEEE Transactions on Antennas and Propagation, July 1973; pp. 435–45.
Lee, T. H., and W. D. Burnside, “Performance Trade-Off Between Serrated Edge and
Blended Rolled Edge for Compact Range Reflectors,” IEEE Transactions on Antennas
and Propagation, January 1996, pp. 87–96.
Long, M. W., “Wide Angle Radiation Measurements,” Proceedings of the 6th Conference
on Radio Interference Reduction and Electronic Compatibility, IIT Research Institute,
Chicago, Illinois, October 1960.
Mahmoud, M. S. A., T. H. Lee, and W. D. Burnside, “Enhanced Compact Range Concept
Using an R-Card Fence: Two-Dimensional Case,” IEEE Transactions on Antennas and
Propagation, March 2001, pp. 419–28.
Nyquist, H., “Thermal Agitation of Electric Charge in Conductors,” Physical Review, vol.
32, July 1928, pp. 110–13.
Pearson, L. W., X. Wang, D. G. Bodnar, and B. Tian, “Spherical Near-Field Scanning
Measurements of Base Station Antennas,” 2007 Antenna Systems / Short-Range
Wireless Conference, Denver, Co, Sept 26, 2007.
Pistorius, C. W., and W. D. Burnside, “An Improved Main Reflector Design for Compact
Range Reflectors,” IEEE Transactions on Antennas and Propagation, March 1987,
pp. 342–47.
420 Antenna Measurements
Chapter 1
1. 2.59 v/m
3. (a) 1.93 × 10−8 w/m2. (b) 7.16 × 10−6 a/m
5. 15.52°
7. (a) 159 km. (b) 198 km.
9. 1.27 × 10−5 w/m2
Chapter 2
1. The plotted curve is a sinusoid of unit amplitude, with its negative maximum at
x approximately 0.75 m and its positive maximum at x approximately 2.25 m.
3. (a) Irms = −0.22 − j1.82 amp; |Irms| = 1.83 amp
(b) f = −96.9°
5. (a) 215.4 ohms. (b) L = 3.43 × 10−6 henry (3.43 µh)
7. (a) r = 0.52 + j 0.3; r = |r|ejf with |r| = 0.6 and f = 30°
(b) ZL = 1198 + j1125
9. 50 ohms
Chapter 3
1. (a) F; (b) T; (c) T; (d) F; (e) F; (f) F
4π
3. D = = 2.1 × 105 , or 53.2 dB
Ω
5. (a) 23.9 m2 (b) 5.5 m
7. (a) Rr = 75 − 8 = 67 ohms (b) kr = 67/75 = 0.89 (c) G = 1.46
9. N
Chapter 4
1. The radiation field is ten times the reactive field at distance r = 10 m.
3. (a) 12.5a. (b) 2.5a. (c) Zero
423
424 Answers to Problems
Chapter 5
1. (a) a = 120°.
(b) f 0°, 180° 30°, 150° 60°, 120° 90° 210°, 330° 240°, 300° 270°
E 0.50 0.71 0.83 0.87 0.26 0.07 0
Chapter 6
1. From Eq. 6–1, (D/2)2 = (4)(10)(6.4) = 256. (a) Therefore, Da = 32 feet. (b)
f/D = 10/32 = 0.3125.
3. cot b = 0.6 − 1/2.4 = 0.1833; b = 79.43°. Therefore (a) 2b = 158.9°. (b) t2 =
−12(79.43/100)2 = −7.58 dB. (c) t1 = 40 log cos b/2 = −4.56 dB. Therefore t1 + t2 =
−12.14 dB. Hence (from tabulation given on p. 238) sidelobe level is about −26 dB.
5. l = 2d = 1 meter = 3.28 feet = 39.36 inches. Da = 1.414 meters = 4.64 feet = 55.68
inches.
7. The reflector antenna is: (i) lighter in weight; (ii) less complicated; (iii) less
expensive.
9. φ = +/−5.74°
Chapter 7
1. (a) 535 MHz and 825 MHz ( f1 = 1.07f0 and f2 = 1.65f0).
(b) 500 MHz and 875 MHz ( f1 = 1.00f0 and f2 = 1.75f0).
(Answers within ± 5 MHz of these figures are acceptable.)
3. 31, 38, 48, 75, 94, 117, 146, 183, 229, and 286 Mz.
Answers to Problems 425
Chapter 8
1. 1.67 ns
3. Zero degrees
5. 2.6 ns
7. 8 degrees
9. (a) 5, (b) 11.25 degrees and 180 degrees
Chapter 9
1. 743 meters or 2438 feet
3. G = 54 or 17.3 dB
5. G = 2512 or 34 dB
7. 1.1 watt
9. 1088.9 K
11. 37.3 dB
APPENDIX A
Maxwell’s Equations
427
428 Maxwell’s Equations
div D = ∇⋅ D = ρ (A–1)
div B = ∇⋅ B = 0 (A–2)
−∂B
curl E = ∇ × E = (A–3)
∂t
∂D
curl H = ∇ × H = J + (A–4)
∂t
Equations (A–5) and (A–6) that follow include the definitions of divergence and curl in
rectangular coordinates.
curl H = J, (A–7)
that describes previous experimental observations of the magnetic field H resulting from
a steady current density J. However, from vector calculus it is known that div curl A is
Maxwell’s Equations 429
zero. Therefore div J = 0, which means that “the current is always closed and there are
no sources or sinks” (Slater and Frank 1947, p. 84). Plates of a capacitor are examples
of both a source and a sink. In this regard, consider discharging a capacitor. Current starts
at the positive plate, its charge decreases as the current flows to the negative plate, and
there the charge is diminished and removed. Thus, one plate serves as a source and the
other a sink.
To mathematically include sinks and sources in electrical phenomena, Maxwell
invented displacement current density, which is ∂D/∂t. It is entirely different than conduc-
tion current density J that flows, for example, because of the conductivity of wire. Beliefs
pronounced by Maxwell include (a) displacement current only exists in the presence of
a varying field, and (b) displacement and conduction currents will produce, separately
or collectively, a magnetic field.
Thus, because matter has both conducting and dielectric properties, Maxwell
added ∂D/∂t to J in (A–7) to provide his revolutionary equation (A–4). After this
change, the divergence of curl H from (A–4) is zero and the contradiction between
mathematics and physical phenomena, as conjectured by Maxwell, is removed. Then,
the combination of (A–3) and (A–4) provide the theoretical mechanism whereby E-fields
create H-fields, and conversely; thereby creating electromagnetic waves that propagate
indefinitely.
A basis for displacement current density having amplitude ∂D/∂t follows. Let
v = voltage across two parallel capacitor plates
d = spacing of the plates
A= area of each plate
e = dielectric constant
C= eA/d, where C is capacitance
I = current input to a fixed capacitance
From basic circuit analysis, the current I that flows into a fixed capacitance is related to
the change in voltage across its capacitor plates as follows:
I = C ⋅ ( ∂v ∂t ) = ( ε A d ) ⋅ ( ∂v ∂t ) = ε A ⋅ ( ∂E ∂t ) = A ⋅ ( ∂D ∂t )
Then, the current density along the capacitance plates is I/A, that equals ∂D/∂t. Thus,
since a variable voltage applied to a capacitor forms a closed circuit, the displacement
“current” density between the plates equals the current density at the capacitor plates.
Therefore, total current density (J + ∂D/∂t) is continuous around the circuit.
Maxwell’s equations are simplified when considering plane waves in a uniform mate-
rial having permittivity e, magnetic permeability m, zero conductivity s, and zero electric
charge r. Even further simplification occurs if we assume the propagation medium is
lossless, so that s = 0. Then Maxwell’s equations, in differential form, become
∂H
curl E = − µ (A–10)
∂t
∂E
curl H = ε (A–11)
∂t
Equations (A–8) through (A–11) are much simpler than (A–1) through (A–4). Equa-
tions (A–8) and (A–9) differ from (A–1) and (A–2) only because r is now zero. Equation
(A–10) says a changing magnetic field will produce an electric field, and (A–11) says
that a changing electric field will produce a magnetic field. Since conductivity is assumed
zero, (A–11) does not include the conduction current term J = s E of (A–4), that other-
wise also contributes to producing magnetic field. Equations (A–8) through (A–11) can
be solved to obtain the plane wave equations (1–2) and (1–3) of chapter 1 (see, e.g.,
Slater and Frank 1947, pp. 90–93). However, the necessary mathematical detail is not
included here. Instead, the effects of (A–10) and (A–11) on producing E-M waves are
now discussed.
As noted above, a changing electric field produces a magnetic field and that in turn
produces an electric field, and so on. Thus, there is a series of energy transfers started
when either an electric or magnetic disturbance occurs. Energy is transferred from the
electric to the magnetic field, and back to the electric field, and repeated indefinitely.
Therefore, in this process energy is being transferred from one form to the other as it
propagates through space, the result within a uniform material is that electromagnetic
waves propagates at constant velocity. Then, the effect of an electric or a magnetic dis-
turbance at one location creates an E-M wave that later reaches another location. The
resulting time delay depends on separation distance between the two locations, and on e
and m that control the propagation velocity.
Maxwell was able to calculate the velocity of electromagnetic waves, if they were to
exist, in terms of e and m. Then, based on his careful electrical measurements on e and
m in about 1865, he expected that E-M waves travel at the rate of about 3 × 108 m/s.
Furthermore, he collected all the best measurements on the speed of light, and he found
that the average of those available to him was about the same that he had predicted from
his measurements and wave calculations (Skilling 1942, p. 112). Thus, Maxwell hypoth-
esized that light is an electric wave and it appeared to him that he had substantiated his
hypothesis. However, many scientists did not accept Maxwell’s hypothesis until 1887.
This was when Heinrich R. Hertz proved that radio waves and light have similar proper-
ties, including the same velocity in a vacuum.
As a wave passes through an imaginary surface in space, there is a flow of power
through the surface area. The power density at the surface can be expressed as a vector
P, called the Poynting vector after a mathematician of the nineteenth century. In addition
to the magnitude of power density, the vector P gives the direction of power flow, that
is, direction of propagation. In terms of vector analysis,
P = E×H (A–12)
References 431
where E and H are the electric and magnetic field vectors. The magnitude of the vector
cross product A × B is A⋅B sin g, where g is the angle between A and B. Thus, since E
and H are perpendicular, the magnitude of P is simply E⋅H. With E and H denoting
effective (rms) amplitudes, the average of the Poynting vector, that is, power density, is
E2/377. The direction of the Poynting vector, as with other cross products, is obtained
from a right-hand rule. Specifically, if the fingers of the right hand curve from E to H,
the thumb shows the direction of propagation. Thus, from the relative pointing directions
of the E- and H-fields, one can determine the direction of propagation.
In conclusion, experimental optics and ray analyses for light were known before
Maxwell conceived and developed his equations, but without them a theoretical explana-
tion for the propagation of light did not exist. After Maxwell developed the equations in
1864, he used them to calculate the propagation velocity of light, and the calculations
agreed with previously measured values. Still later, in 1887, Heinrich Hertz demonstrated
that the velocities (and other properties) of radio waves and light were the same. Thus,
from the use of Maxwell’s equations and experimental observations, the theory of E-M
waves was established. Today it is known that the phenomenon of E-M radiation covers
the entire energy spectrum from very high frequency cosmic rays to low-frequency, long-
wavelength radio waves.
References
Herrera, J. C., “Electromagnetic Radiation,” in Encyclopedia of Physics, 2nd ed., by R. G.
Lerner, R. G. and G. L. Trigg, VCH Publishers, 1991, pp. 285–88.
Skilling, H. H., Fundamentals of Electric Waves, John Wiley and Sons, 1942, p. 112.
Slater, J. C., and N. H. Frank, Electromagnetism, McGraw-Hill, 1947.
APPENDIX B
Polarization Theory
Wave polarization is introduced in sec. 1.1.3 and further discussed in sec. 3.10. Polariza-
tion is a property of an electromagnetic wave, which describes the time-varying direction
and amplitude of its E-vector. Specifically, polarization describes the figure traced as
a function of time of the extremity of the E-vector at a fixed location in space, when
observed looking in the direction of propagation. Polarization of an antenna is defined
as the polarization of the far-field wave that it radiates. As discussed in secs. 1.1.3 and
3.10, linearly polarized waves are the most commonly used. When the amplitude of
E-vector varies and the E-vector points only up and down, the wave is vertically or V
polarized. Similarly, if the amplitude and direction of the E-vector varies and its direction
is only horizontal, the wave is horizontally or H polarized. Circular polarization (CP) is
the next most frequently used polarization. Its E-field has constant magnitude, but its
direction rotates (within a plane) continuously through 360° and it completes a cycle in
one wave period.
The most general form of a wave polarization is elliptical. The E-vector of an ellipti-
cally polarized wave can be regarded from three viewpoints. Namely, it can be considered
to be (1) a rotating vector, the end point of which traces out an elliptical helix whose
axis lies in the direction of propagation; (2) the resultant of the E-vectors of two linearly
polarized waves of the same frequency; or (3) the resultant of the oppositely rotating E-
vectors of two circularly polarized waves of the same frequency. Figure B–1 illustrates
the E-vector of an elliptically polarized wave at various positions in space, at a fixed
instant of time. An ellipse within an x-y plane is shown at the left end of this spiraling
wave. The ellipse is created by tracing the end-point (terminus) of the E-vector onto the
x-y plane, during the duration of one rf cycle.
E = 1x E x + 1y E y (B–1)
433
434 Polarization Theory
Transmitting where
Antenna
E x = E xm sin (ω t − γ z ) (B–2)
and
Y E y = E ym sin (ω t − γ z + δ ) (B–3)
Left Screw-Sense in Space
X In (B–2) and (B–3), Exm and Eym are the
Z Counterclockwise Rotation in Plane peak amplitudes of the x- and y-polarized
FIGURE B–1. electric fields.
Equations (B–1) through (B–3) describe
A left circularly polarized wave exhibits
counterclockwise rotation of the E-vector in a an elliptically polarized wave of radial
plane, when viewed in propagation direction Z. frequency w and with propagation con-
(Adapted from Huynen 1970, p. 10.) stant g traveling in the positive z direc-
tion. There is a phase difference d between
its x- and y-directed components, which
Eym
are in space quadrature and which have
maximum values Exm and Eym. If viewed
in the direction of propagation, the end-
E2 point of the E-vector traces an ellipse on
a fixed plane in space, as shown in
Fig. B–2. This ellipse is known as a
polarization ellipse.
Exm Readers will recognize the similarity
of a polarization ellipse and a Lissajous
E1
figure. A Lissajous figure is used to deter-
mine relative amplitude and relative phase
of two equal-frequency sinusoids with an
oscilloscope. In the polarization ellipse of
Fig. B–2, the E-field is obviously largest
Note: when pointed along the major axis and
+z is into page smallest when pointed along the minor
+x is to the left
+y is up axis. Also, notice that the field strengths
in Fig. B–2 along the x- and y-axes are
FIGURE B–2.
smaller than Exm and Eym, respectively.
Resolution of an elliptically polarized wave These properties, as well as major-axis tilt
into linearly polarized components along the +x
and +y axes. and E-vector rotation direction, depend
on the relative phase d and the ratio
Exm/Eym.
Now examine (B–1) through (B–3). If Eym = 0, the E-vector varies in magnitude and
direction, but it is always along the x-axis. Thus, the wave is linearly polarized in the x
direction. Similarly, if Exm = 0, the E-vector is always along the y-axis and the wave is
linearly polarized in the y direction. Now suppose Exm ≠ 0, Eym ≠ 0, and d = 0 or
Elliptically Polarized Waves Resolved into Circularly Polarized Components 435
tilt angle of the linearly polarized E-field –180° –135° –90° –45° –0° +45° +90° +135° +180°
then depends on the ratio Exm/Eym. With δ
Exm = Eym and the special cases d = 0 or FIGURE B–3.
d = 180°, the linear polarization is ori- Polarization ellipses with waves propagating
ented at 45° to the X- and the Y-axes. into the page. The polarization is right circular
Other special cases are when d = ±90°. if d = −90° and left circular if d = +90°.
Then, the major axis will lie along the x-
axis if Exm > Eym and along the y-axis if
Eym > Exm. After horizontal and vertical
polarizations, circular polarization (CP) EL E
is the most frequently used (sec. 7.3). CP
is the special case of d = ±90° and Exm =
Eym. Using (B–2) and (B–3), let d = 90°
and t = z = 0; then Ex = 0 and Ey = Eym.
ER
Now consider a later time so that wt =
90°; then Ex = Exm and Ey = 0. Thus, E
rotates from the +y direction to the +x
direction during the time that causes wt to
change from zero to 90°. In other words,
E rotates counter-clockwise when looking
outward (the direction of propagation);
that is, the polarization is left circular. Note:
Correspondingly, if d = −90° the polariza- +z is into page
+x is to the left
tion is right circular. +y is up
It should now be apparent that knowl-
FIGURE B–4.
edge of the amplitudes of the two orthog-
onal (perpendicular) linear components Resolution of a polarization ellipse into a left-
and the phase difference between them is circular EL and a right-circular ER polarized
wave. Note that the relative phase D of EL and
sufficient to define a polarization ellipse. ER is determined by the orientation of the
Figure B–3 includes the polarization major axis.
ellipses for selected ratios Eym/Exm and
relative phases d.
polarized wave. Similarly, the E-vector of a left circular, or LC, polarized wave rotates
counterclockwise if the observer looks in the direction of propagation.
An elliptically polarized wave can be created by the vector sum of an RC and an LC
wave. This is done below by using phasor notation to indicate the E-field direction versus
both time and position. In other words, the phasor serves the role of a unit vector having
variable direction. In (B–4) through (B–6) below, EL and ER are the constant amplitudes
of the left- and right-circular polarized waves, and bold type indicates a vector having
both amplitude and direction.
E = EL + ER (B–4)
where
EL = EL e j (w t −g z) (B–5)
and
E R = E R e - j (w t - g z + D ) (B–6)
E R + EL
AR = (B–7)
E R − EL
From Fig. B–4 it also can be determined that the rotation direction of the E-vector is
the same as that of the larger of ER and EL. Thus the ratio of the magnitudes of the RC
and LC waves determines the direction of rotation, in addition to the axial ratio AR.
More detailed discussion of the theory of polarized waves can be found in Kraus
(1988, pp. 70–80), and in Stutzman (1993, pp. 1–65). Stutzman is also recommended
reading on multipolarization antennas, systems, and phenomena. Long (2001) is recom-
mended for reading on the polarization and depolarization properties of radar echoes
from land and sea.
The magnitudes of the four polarizations (H, V, RC, LC) may be received simultane-
ously with four separate antennas. Alternately, the separate signals from two orthogonal
linear or two orthogonal circular polarization antennas, when combined appropriately,
can be used to obtain simultaneously the magnitudes of the four separately polarized
signals. Further, a dual linear or a dual circular polarization feed (coupler) can be
employed, thereby allowing the amplitudes of each of the four polarizations to be simul-
taneously obtained with an appropriately designed antenna.
References
Huynen, J. R., Phenomenological Theory of Radar Targets, Drukkerij Bronder-Offset N.V.,
1970, p. 10.
Kraus, J, D., Antennas, 2nd ed., McGraw-Hill, 1988.
Long, M. W., Radar Reflectivity of Land and Sea, 3rd ed., Artech House, 2001.
Stutzman, W. L., Polarization in Electromagnetic Systems, Artech House, 1993.
APPENDIX C
Review of Complex-
Variable Algebra
Complex variables are (variable) numbers containing the factor j, which is a symbol
denoting −1. The notation −1 means “the number that, when squared (multiplied by
itself), equals −1.” But none of the “real numbers” fits this specification. There is no real
number that, when multiplied by itself, equals −1. When any real number, positive or
negative, is squared, the result is a positive number. In particular, (−1)2 = +1.
Therefore, the “number” that, when squared, equals −1, represents an extension of the
system of real numbers. It is called an “imaginary” number—possibly unfortunate ter-
minology, since it implies something that does not “really” exist. Imaginary numbers
exist just as much as do the so-called real numbers. Although j is “imaginary,” it need
not be mysterious. It is just another way of writing −1. It is a different kind of number
from the familiar real numbers, which are 0, ±1, ±2, ±3, etc. Yet obviously it has some-
thing in common with them, since the symbol “1,” which is part of the quantity −1, is
a real number.
When j (or as it is denoted in physics, i) is multiplied by any real number (for example,
j5 or −j6), the resulting product is an imaginary number. Thus the totality of imaginary
numbers is obtained by multiplying j by all the real numbers. Another way of writing j5
is 5 −1 or −25 . Another way of writing −j6 is −6 −1 or − −36 . (Note that − −36
is not the same thing as +36 , which is of course a real number equal to either +6 or
−6.)
If two imaginary numbers are multiplied together (for example j3 × j5), the ordi-
nary rules of algebraic multiplication apply. In other words, j3 × j5 = j215. But
j 2 = j × j = −1 × −1 = −1 . Therefore, j3 × j5 = −15, which is a real number. Thus the
product of two imaginary numbers is a real number (further evidence that imaginary
numbers are not wholly “unreal”).
By the same process of reasoning, all even powers of j are real numbers. Thus, j 2 = −1
(as was just shown); j 4 = j 2 × j 2 = −1 × −1 = +1; j 6 = −1; j 8 = +1; and so on. On the
other hand, odd powers of j are imaginary: j 3 = j 2 × j = −j; j 5 = j 4 × j = +j; and so forth.
The reciprocal of j is equal to −j, as can readily be seen by the following argument:
1/j = j/j 2 = j/(−1) = −j. Reciprocals of odd powers of j may thus be shown equal to minus
the value of the odd power; that is: 1/j 3 = −j 3 = +j, etc. Reciprocals of even powers are
equal to the even power: 1/j 2 = j 2; 1/j 4 = j 4; etc. Reciprocal quantities are of course also
439
440 Review of Complex-Variable Algebra
expressible with negative exponents; 1/j = j −1; 1/j 2 = j −2; and so on. In short, j may be
manipulated algebraically in exactly the same way as any symbol.
A complex number is the sum of a real number and an imaginary number. Thus if
z = x + jy, where both x and y are any real numbers, then z is a complex number. (Actu-
ally, since either x or y may have the value zero, in which case z will be either purely
imaginary or purely real, the purely real and purely imaginary numbers are both special
cases of complex numbers.) The real number x is called the real part of z, and the real
number y is called the imaginary part of z. This statement is sometimes written in the
abbreviated form: Re (z) = x; Im (z) = y. Notice that y is a real number, though it is called
the imaginary part of z. The imaginary number is jy.
Manipulation of complex numbers follows all the rules of ordinary algebra. The
symbol j is treated like any other algebraic symbol, although it may at times be conve-
nient to convert expressions like j2 and j3 into their equivalent forms, −1 and −j, and
so on.
As an example, if the two complex numbers z = x + jy and w = u + jv are to be added,
the result is
z + w = ( x + u) + j( y + v ) (C–1)
That is, the real part of the sum is the sum of the real parts of the added complex numbers,
and the imaginary part of the sum is the sum of the imaginary parts.
Multiplication of imaginary numbers is performed as follows:
(since j2 = −1).
The real numbers are represented as points on a line, with one point designated zero;
numbers to the left are “mirror images” of numbers to the right except that those to the
left are negative and those on the right are positive. The complex numbers are represented
as points in a plane by a cartesian coordinate system. The horizontal coordinate axis
is the real-number line just described, and the vertical coordinate axis is an identical
imaginary-number line, its zero coinciding with the real-axis zero. This complex-plane
representation is illustrated in Fig. C–1 for the particular complex numbers x = 4, y = 3
(z = 4 + j3), and x = −2, y = 5 (z = −2 + j5).
In this representation the direct or straight-line distance from the origin to the
complex-number point is called its modulus or absolute value. The modulus is (by
definition) a positive real number, regardless of which quadrant of the coordinate plane
contains the number point. It is the length of a line. The modulus of z1 in Fig. C–1 is
designated r1. From the Pythagorean theorem for right triangles it is apparent that
r1 = x12 + y12 = 42 + 32 = 25 = 5 . That is, the modulus of a complex number is the
(positive) square root of the sum of the squares of its real and imaginary parts. The
modulus is denoted algebraically by the “absolute value” brackets ||; thus:
z = x + jy = x 2 + y2 (C–3)
Review of Complex-Variable Algebra 441
z = z e jθ = re jθ (C–4)
The equivalence of this representation and the original one is based on deMoivre’s
theorem, which defines the meaning of a complex exponent as follows:
(The proof of this theorem is based on concepts of advanced calculus.) Rewriting (C–4)
in terms of deMoivre’s theorem gives
By inspection of the right triangle of Fig. C–1 (in which z = z1, r = r1, etc.) it is apparent
that
x = r cos θ
(C–7)
y = r sin θ
442 Review of Complex-Variable Algebra
z = x + jy (C–8)
( x + jy ) ( x − jy ) = x 2 − j 2 y 2 = x 2 + y 2 (C–9)
In the complex exponential notation the conjugate of rejq is re−jq. This is shown from de
Moivre’s theorem in the following way:
e − jθ = e j( −θ ) = cos ( −θ ) + j sin ( −θ )
(C–10)
= cos θ − j sin θ
because cos(−q) = cos(+q) and sin(−q) = −sin(+q). That the product of the complex
conjugates yields the square of the modulus is also readily seen from this notation, as
follows:
re jθ ⋅ re − jθ = r 2 e( jθ − jθ )
(C–11)
= r 2 e0 = r 2
a + jb ( a + jb ) ( c − jd ) ac + jbc − jad − j 2 bd ac + bd bc − ad
= = = 2 + j 2 (C–12)
c + jd ( c + jd ) ( c − jd ) c −j d
2 2 2
c +d 2 c + d 2
An important theorem of complex algebra is that if two complex numbers are equal
to each other, their real parts and their imaginary parts are separately equal. This means
that every complex equation is equivalent to two real equations. Thus, if
Review of Complex-Variable Algebra 443
x + jy = u + jv (C–13)
x=u
y=v (C–14)
(two equations). This theorem has many applications. For example, the relationship
between impedance and admittance is Z = 1/Y, where in general Z = R + jX and Y =
G + jB. Therefore
1 R − jX R X
G + jB = = 2 = 2 − j 2 (C–15)
R + jX R + X 2
R +X 2 R + X 2
R
G= (C–16)
R + X2
2
−X
B= (C–17)
R + X2
2
as was stated without detailed proof in equations (2–30) and (2–31), chapter 2.
APPENDIX D
445
446 Complex Reflection Coefficients and Multipath Effects
sin ψ − (ε r − cos2 ψ )
12
ΓH = = ρ H exp ( − jφ H ) (D–1)
sin ψ + (ε r − cos2 ψ )
12
ε r sin ψ − (ε r − cos2 ψ )
12
The reader will notice that (D–1) and (D–2) contain −fH and −fv terms, meaning that
positive fH and fv denote phase delays. This use of phase delay follows Kerr (1947, pp.
396–97) and is consistent with the fact that the phase of reflected waves for land and sea
lags that of the incidence wave. This practice of defining fH and fv as phase lags differs
from sec. 1.2, where positive f denotes leading phase. As also discussed in sec. D.2,
there can be uncertainties regarding the correct phases when using (D–1) and (D–2),
because the square root of a complex (real and imagery parts) number has two answers
and they are at different angles. Consequently, because of this ambiguity in phase, it is
useful to remember that, for land and sea, reflected phase always lags.
As shown in Fig. D–1, the angle y is the angle between the incident wave (and the
reflected wave) and the tangent to the surface at the point of reflection. The term er is
the complex permittivity, relative to permittivity of vacuum, expressed as
where l is the free-space wavelength in meters and s is conductivity in mhos per meter.
Example values of e′r and e″r for soil and water are given in Table D-1. By definition,
er = 1 for a vacuum and it is essentially the same for the earth’s atmosphere. In addition,
Appendix H includes the electrical properties of some low-loss dielectric materials used
for radomes.
For normal incidence (y = 90°), where EH and EV are parallel to the reflecting plane,
the H and V reflection mechanisms are necessarily identical. However, from (D–1) and
(D–2) GH and GV become
Path of Path of
Incident Wave Reflected Wave 1 − ε r1 2
EV V EV
ΓH = (D–4)
1 + ε r1 2
EH H EH
ε r − ε 1r 2 ⎛ ε −1 2 ⎞ ⎛ ε r − ε 1r 2 ⎞
ΓV = = = −ΓH
ε r + ε 1r 2 ⎜⎝ ε −1 2 ⎟⎠ ⎜⎝ ε r + ε 1r 2 ⎟⎠
y y Horizontal (D–5)
TABLE D-1 Approximate Electrical Properties of Soil and Water Sources: Kerr (1947,
p. 398) and Dicaudo (1970, p. 14–30)
0
0 10 20 30 40 50 60 70 80 90
Grazing Angle in Degrees
D.2. Reflection Coefficients for V-POL, 10 MHz
Smooth Land and Sea V POL, 10 GHz
H-POL, 10 MHz
Figures D–2 and D–3 show the general H-POL, 10 GHz
trends in reflection coefficient versus FIGURE D–2.
grazing angle y and frequency for a flat, Magnitude r of reflection coefficient for
smooth seawater surface. Figure D–2 smooth sea surface versus grazing angle Y.
includes reflection coefficient amplitude 10 MHz & 10 GHz; horizontal & vertical
r for both H- and V-polarizations at two polarizations.
frequencies, 10 MHz and 10 GHz, with
e′r = 80 and s = 4.3 mhos/m assumed.
Note that r is essentially unity at zero grazing angle for each polarization and frequency.
The magnitude rH is usually close to unity at all y angles, but it drops to as low as 0.8
at the higher frequencies and larger y values. Figure D–3 shows examples of phase delay
on reflection for H- and V-polarizations at 10 MHz and 10 GHz. For H-polarization, the
448 Complex Reflection Coefficients and Multipath Effects
200
Phase Delay of Reflection Coefficient
grazing angles approach zero, the complex reflection coefficients for horizontal and verti-
cal polarizations approach −1 for both land and sea.
σ h sin ψ ≥ λ 8 (D–6)
Experience indicates that the “significant wave height,” the height of a sea wave that
most observers guess from a quick observation, exceeds sh of sea surfaces by a factor
of about four. Thus, the standard deviation is actually substantially smaller than peak-to-
trough surface height.
A further refinement in calculating the effects of roughness provides an estimate of
the effective value of the magnitude of G, as follows:
2πσ h sin ψ 2
Γ = ρo exp −2 (D–7)
λ
Here ro is the magnitude of the reflection coefficient for a smooth, flat surface. From
(D–7), |G | is approximately ro if sh sin y << l and it equals 0.29ro if sh sin y = l/8.
Thus, as is already known, surface reflections become more specular-like with decreased
450 Complex Reflection Coefficients and Multipath Effects
t t Gr λ
2
PG
Pr = F2 (D–8)
( 4π ) R
2 2
where Pt is transmit power, Gt and Gr are transmit and receive antenna gains, and R is
distance between transmit and receive antennas.
Figure D–4 shows the basic multipath process that consists of the phasor addition of
waves at point P. The shortest path is the direct path and it has length designated as Rd,
Rd
A R2
d
h2
R1
r
h1
y y
B
–h1
FIGURE D–4.
Geometry for a flat earth, with image of source A located at −h2.
Flat Earth Geometry 451
and the other path is the indirect path and has length R1 + R2. The phase difference of
the two waves is affected by three factors: (l) the path-length difference d which equals
(R1 + R2 − Rd), (2) the phase change of the reflected wave that occurs upon reflection,
and (3) the phase difference, if any, of the fields radiated by the antenna in the direct-ray
and reflected-ray directions (qd and qr, respectively).
The two waves may also have amplitude differences caused by the following factors:
(l) each wave is subject to the inverse-square-law power-density reduction, and they
travel different distances, (2) the reflected wave undergoes at least some loss of intensity
in the process of reflection, and (3) antenna patterns are, in general, not constant in
the vertical plane and may therefore have different field intensities in the qd and qr
directions.
In most multipath situations, the path difference d is small compared to the path lengths
of the direct and reflected paths. Therefore, the usual small difference in amplitudes
caused by these path length differences d can be neglected, and the primary effect of d
is on the phase difference between the waves that arrive from the direct and reflected
paths.
As already noted, the elevation pattern of the antenna affects the relative strength of
the rays in directions qd and qr, and therefore possible effects of the pattern must be
addressed. Often in multipath situations, however, the angles qd and qr are small enough
that the pattern’s relative strengths at those angles are insignificant. Then, if both the
direct and indirect paths are energized equally, the resultant electric field at point P is
where Ed is the field due to the direct wave, d is the difference in lengths between the
indirect and direct paths, and f is the phase delay created by the reflecting properties of
a surface. Therefore, in accordance with (D–9), the propagation factor F is
F = E Ed = 1 + ρe − j[(2π λ )δ +φ ] (D–10)
Note that F consists of two components, of amplitude one and r, separated by angle
(2p/l)d + f. Thus, from use of the law of cosines, it is seen that (D–10) becomes
F 2 = (1 + ρ 2 + 2 ρ cos ψ ) (D–11)
where Y = (2p/l)d + f.
1 h − h 2
Rd ≈ G 1 + 2 1 (D–14)
2 G
and
1 h + h 2
R1 + R2 ≈ G 1 + 2 1
2 G
By using the equations (D–14), it can be seen that the difference d, in the reflected
wave path-length R1 + R2 and the direct wave-path length Rd, may be expressed as
2h1h2
δ = ( R1 + R2 ) − Rd ≈ (D–15)
G
h2 − h1
sin θ d = (D–16)
Rd
Multipath Dependencies on Frequency, Polarization, and Surface Roughness 453
cos θ d = G Rd (D–17)
h1 + h2
sin ψ = (D–18)
G
θr = ψ (D–19)
Recall that the solution of multipath problems in exact form requires knowledge of y
versus qd. However, the relationship between y and qd is not available, mathematically,
in a simple closed-form. However, accurate multipath solutions can be acquired by com-
puter with step-by-step procedures. For example, to calculate an antenna pattern at dis-
tance Rd and with antenna height h1, let h1 and Rd be constants. Then, the following are
attained: qd versus h2 from (D–16), G versus qd from (D–17), y versus G from (D–18),
and qr from (D–19).
f = 1 GHZ
θ = 0.8° θ = 4° θ = 8°
6 6 6
5 5 5
HEIGHT ABOVE SEA (ft)
4 4 4
3 3 3
2 2 2
1 1 1
0 0 0
0 1 2 0 1 2 0 1 2
f = 1 GHZ RELATIVE FIELD STRENGTHS
θ = 0.8° θ = 4° θ = 8°
6 6 6
5 5 5
HEIGHT ABOVE SEA (ft)
4 4 4
3 3 3
2 2 2
1 1 1
0 0 0
0 1 2 0 1 2 0 1 2
RELATIVE FIELD STRENGTHS
SOLID LINES INDICATE HORIZONTAL POLARIZATION
DOTTED LINES INDICATE VERTICAL POLARIZATION
FIGURE D–5.
Electric field amplitude patterns above a smooth sea relative to free-space amplitudes. 1 and
4 GHz, with 0.8°, 4°, and 8° grazing angles. From Long, Wetherington, Edwards, and Abeling
(1965).
Figures D–7 and D–8 show the propagation factor F for H-POL with and without
surface roughness. The curves were prepared under the following assumptions: sh = 0
(flat, smooth earth) and sh = 0.3 m (representing a recently plowed agricultural field),
l = 3 m (100 MHz) and 0.3 m (1 GHz), and land with er = 16 − j60ls and s = 10−2 mho/
m. In Fig. D–7 for 100 MHz, the surface roughness causes only a slight filling of the
minimum at 30°, and there are little effects of surface roughness at smaller grazing angles.
Also in Fig. D–7, the effect of surface roughness at 100 MHz is not strong, even for
Spherical Earth Geometry 455
Propagation Factor F in dB
0
not apparent for angles less than 3°, and –10
for the rough surface essentially all effects
–20
of multipath have disappeared for angles
above 10°. Thus, in some applications, –30
the use of higher frequencies may be a –40
useful method of minimizing the effects –50
of multipath interference. There are prac-
–60
tical limits, of course, on the use of the
higher frequencies (shorter wavelengths) –70
0 10 20 30 40 50 60 70 80 90
imposed by the general increase in atmo- Grazing Angle in Degrees
spheric attenuation. In summary, the elec- V-POL, 1 MHz
H-POL, 1 MHz
tromagnetic effects of roughness depend V-POL, 100 MHz
on sh, q, and l through their relationships H-POL, 100 MHz
with the roughness parameter sh sin y/l FIGURE D–6.
and on its effect on the magnitude of Propagation factor F in dB versus grazing
reflection co-efficient in accordance with angle. V- and H-polarizations, 1 and 100 MHz,
(D–7) of sec. D–3. antenna 3 m above flat smooth land.
Similarly, the distance from the tangent point to point P of height h2 is 2ae h2 + h22 .
Therefore, the maximum practical range for detecting an electromagnetic wave emitted
from point A, given that the heights at points A and P are h1 and h2, is
The equations that follow are from Blake (1986, pp. 249–58). They allow the param-
eters of Fig. D–9 to be computed if the heights h1 and h2, the total ground range G, and
the equivalent earth radius ae are given. The first step is to obtain the ground ranges
G1 and G2, where G is (G1 + G2). This requires solving for two parameters, p and x,
that follow:
Spherical Earth Geometry 457
P
Horizontal
at point P
Rd
Horizontal
at Point A R2
h2
d
Antenna
Effective
A R1 Earth
r
y Surface
y
h1 B G2
G1
ae ae
Earth Center
C
FIGURE D–9.
Geometry for spherical-earth model. (Effective earth has radius ae, compatible with assumption
of straight ray paths A-P and A-B-P). Adapted from Blake (1991, p. 255).
458 Complex Reflection Coefficients and Multipath Effects
2
h −h
ae( h1 + h2 ) +
2 G
p= ξ = sin −12aeG 1 3 2 (D–22)
3 2 p
G sin ξ
G1 = −p G2 = G − G1 (D–23)
2 3
Finally, one can attain R1, R2, and Rd by using the following:
G G
R1 = h12 + 4ae( ae + h1 ) sin 2 1 R2 = h22 + 4ae( ae + h2 ) sin 2 2
2ae 2ae
(D–24)
2 G1 + G2
Rd = ( h2 − h1 ) + 4 ( ae + h1 ) ( ae + h2 ) sin
2
2ae
In (D–24), the angles within the sin2 terms are each expressed in radians.
Now that equations for all linear dimensions of Fig. D–9 are available, the angles qd,
qr and y can be obtained as follow:
If h1 and h2 are both much smaller than ae, which is true for most applications,
h −h R
θ d ≅ sin −1 2 1 − d (D–26)
Rd 2ae
The exact equations for qr and y, along with approximations for the usual scenarios for
which h1 << ae, follow
and
qr is the depression angle, relative to the horizontal at point A of the indirect path ray
between point A and the earth (point B); and
y is the angle measured above the horizontal at the earth (point B) of the indirect path
ray between point A and the earth.
For many applications where multpath interference is significant, the elevation (verti-
cal) antenna pattern is sufficiently broad that the pattern strengths at the beam center and
at angles qd and qr are approximately equal. Then, the approximation of (D–9) in sec.
D.4 is valid. Otherwise, multipath analyses must include effects of the shape of the eleva-
tion pattern.
The pathlength difference d is, of course, equal to (R1 + R2) − Rd. Consequently,
because this equation involves differences in numbers that are large compared to d, sig-
nificant computing errors can occur. Blake (1986, p. 257) uses a more accurate method
for calculating d, which follows:
4 R1R2 sin 2 ψ
δ= (D–29)
R1 + R2 + Rd
Thus, by starting with only the heights h1 and h2, the ground range G and an assumed
equivalent earth radius ae, one can make calculations for the various path lengths and
angles of Fig. D–9. The accompanying SM includes spherical earth calculations using
Mathcad, where the user can choose the initial assumptions in accordance to his or her
needs.
2πσ h sin ψ
2
Γ = ρo exp −2 D (D–30)
λ
460 Complex Reflection Coefficients and Multipath Effects
where ro is the reflection coefficient magnitude if the surface is smooth and the expo-
nential function is the roughness factor that was introduced in Sec. D.3.
From detailed geometrical analysis, Kerr (1951, p. 406) in his equation (16) provides
a general equation for D. This equation, expressed in terms of the parameters of Fig. D–9,
was obtained by the present author by using Kerr’s equations (11) and (12). That equa-
tion that follows Kerr and various approximations for D are included in SM 3.6 of the
accompanying SM. To obtain another and yet simpler equation, Kerr uses assumptions
that are valid if the effects of divergence are appreciable. These assumptions include
small grazing angles, and terminal heights and path lengths small compared to the earth’s
radius. Kerr’s approximation follows:
−1 2
2G1G2
D = 1 + (D–31)
aeG sin ψ
A number of other approximations for D appear in the literature (Balanis 1984, 2005;
Beckmann and Spizzachino 1963). These approximations and (D–31) are compared in
SM 3.6 with both terminal heights (h1 and h2) being 20 km, and each gives almost identi-
cal calculated values. Consequently, each of the equations will provide acceptably cal-
culated geometrical values for D if neither platform height exceeds 20 km. However, an
accurate geometrical calculation of D does not necessarily provide a valid electromag-
netic result.
The risk of using the calculated divergence factor for very small grazing angles (about
one degree or less) is now illustrated. Recall that a small D value implies a small reflected
field, and then the total field would approximate the free space field. Contrarily, measure-
ments indicate that the field strength actually becomes smaller as y becomes smaller and
the horizon is approached. However, because of diffraction from the earth, the pattern
propagation factor F is very small, but not actually zero at ranges that exceed the calcu-
lated horizon Rmax (Blake 1991, pp. 271–74). Accordingly, incautious use of a D value
can cause the calculated pattern propagation factor F versus range to incorrectly approach
the free-space value of unity.
References
Balanis, C. A., R. Hartenstein, and D. DeCarlo, “Multipath Interference for In-Flight
Antenna Measurements,” IEEE Trans. on Antennas and Propagation, January 1984,
pp. 100–104.
Balanis, C. A., Antenna Theory, Analysis and Design, 3rd ed., Wiley-Interscience, 2005,
pp. 208–11.
Beckmann, P., and A. Spizzachino, The Scattering of Electromagnetic Waves from Rough
Surfaces,” The MacMillan Company, 1963, p. 224.
Blake, L.V., Radar Range-Performance Analysis, Munro Publishing, 1991; Artech House,
1986.
Dicaudo, V. J., “Radomes,” ch. 14 in M. I. Skolnik, Radar Handbook, McGraw-Hill, 1970.
Institute of Electrical and Electronics Engineers, IEEE Standard 686-1997, 1997.
References 461
Radomes
A radome is a dielectric structure that is used to protect an antenna from its environment
(wind, rain, ice, salt spray, dust, insects), and it may be used to reduce aerodynamic drag.
Radomes serve as feed covers, covers attached to the antenna, or covers within which
an antenna moves. Therefore, they have a wide variety of shapes and wall structures
(Huddleston and Bassett 1993).
A radome can cause unwanted changes in beam-pointing direction, create energy loss
by reflection and absorption, cause gain loss and side-lobe degradation by defocusing,
create internal reflections that produce impedance mismatches, increase side-lobe levels,
and may also depolarize the radiation. Rulf (1985) discusses electrical design issues for
an airborne radome used with a large, very-low side-lobe antenna. The reader should be
aware that radome design is a highly specialized subject. For example, the design for a
radome with pointed nose for high-speed aircraft may be more complex than the antenna
design.
Schrank, Evans, and Davis (1990) discuss a variety of radome types, and descriptions
of some of the more common radome wall cross sections are given here and illustrated
in Fig. E–1.
Thin wall. A solid dielectric, with thickness usually less than l/10 measured in the
dielectric. The reflections from the air-radome and radome-air interfaces have oppo-
site phases. Thus, if the radome wall thickness is thin compared to the wavelength
within the radome material, the net effect is near-cancellation of the reflections. A
thin wall radome has good electrical properties, but it may be weak structurally
when designed for frequencies above 1 GHz.
Half wavelength. This is a solid dielectric surface having an electrical thickness near
one-half wavelength. A half-wavelength-thick surface is nonreflecting at its design
frequency. Because of the half-wavelength requirement, its bandwidth and its useful
range of incidence angles are limited.
A-sandwich. A commonly used three-layer wall configuration consisting of two thin
dielectric skins separated by a thicker but low-dielectric constant core. The core
might be a honeycomb or a foam material. Since the reflections from the two
463
464 Radomes
Solid, thin wall or l/2 thickness skins will be roughly equal in ampli-
A sandwich, low dielectric constant foam tude and phase, a quarter-wavelength
or honeycomb core, skins spaced l/4 spacing of the skins will minimize their
C sandwich, two back-to-back A sandwiches, combined reflections
for increased strength C-sandwich. Two back-to-back A-sand-
Multiple-layer sandwich, thin layers of fiberglass wiches. It can be used when the ordi-
with low density cores, for good electrical nary A-sandwich does not provide
properties and strength
sufficient strength.
FIGURE E–1. Multiple-layer sandwich. These include
Commonly used radome wall cross sections. many thin layers of fiberglass with low-
density cores for providing great strength
and good electrical performance.
Dicaudo (1970) includes numerous useful graphs that show electrical performance
versus incidence angle for single-layer, A-sandwich, and C-sandwich walls; and for
polarizations parallel and perpendicular to the plane of incidence formed by the direction
of propagation and the normal to the radome surface. These curves show transmission
efficiency and insertion phase delay versus the radome-wall thickness-to-wavelength
ratio. The data are included for several dielectric materials.
Relative permittivity is the permittivity of a medium relative to its free space value eo
of 8.85 × 10−12 farads/meter. To account for the ohmic loss of a dielectric material, a
complex relative permittivty er is used that follows:
(εr′ )1 2 = n = c v (E–2)
where v is wave velocity in the medium and c is wave velocity in free space. Then from
l = v/f and the wavelength l in an open, unbounded (not a waveguide) medium is
λ = v f = c (ε r′ )1 2 f = λ0 (ε r′ )1 2 (E–3)
where l0 is the free space wavelength. Thus, for a dielectric constant of 4 (typical of
epoxy resin fiberglass radomes), the internal wavelength would be one-half l0. Then, for
this example, a layer of l0/10 thickness corresponds to l/20 internal thickness and would
be electrically acceptable as a thin radome or as a thin wall for a layered radome.
The term e″r accounts for a dielectric’s ohmic loss, and the ratio e″r/e′r is defined as the
loss tangent or tan d. Although not relevant to antennas and transmission lines, another
electrical term is dielectric power factor PF, where PF = sin d. Loss tangents for radome
Radomes 465
dielectrics are in the range of 10−4 to 10−2. Thus, for these materials loss tangent and
power factor are essentially equal.
Table E-1 includes typical electrical properties of some radome materials at 10 GHz.
For further details, see Bodnar (2007) and Burks (2007). Although e′r varies with fre-
quency, the values given are generally typical for kiloHertz and gigaHertz frequencies.
On the other hand, e″r = j 60 l s and thus loss tangent e″r is a strong function of fre-
quency. Electrical properties also depend on fabrication method, material density, and
temperature.
The loss per unit length of a dielectric is proportional to loss tangent and is derived
theoretically by Ragan (1948, pp. 28–29). The attenuation, expressed in nepers per meter,
within the dielectric is
or
By using the relationship between attenuation in nepers ANp and in decibels AdB, namely
ANp = 0.115 AdB, we obtain
or by using the index of refraction n, that is, (e′r )1/2, of the dielectric medium,
and
Thus it is seen that ohmic loss, when using a relatively high loss material, is insignificant
for a thin wall radome. However, the loss can be appreciable if the thickness were one
wavelength. Another major concern is the likely negative effect of the radome on antenna
focusing. Generally, thin wall radomes do not create significant degradation to antenna
gain or side lobes. However, much caution is needed when using the thicker radomes,
often needed for structural strength.
References
Bodnar, D. G., “Materials and Design Data,” ch. 55, pp. 55–3 through 55–5 in J. L.
Volakis, Antenna Engineering Handbook, 4th ed., 2007.
Burks, D. G., “Radomes,” ch. 53, pp. 53–16 and 53–17 in J. L. Volakis, Antenna
Engineering Handbook, 4th ed., 2007.
Dicaudo, V. J., “Radomes,” ch. 14 in M. I. Skolnik, Radar Handbook, McGraw-Hill, 1970.
Huddleston, G. K., and H. L. Bassett, “Radomes,” ch. 44 in R. C. Johnson (ed.), Antenna
Engineering Handbook, 3rd ed., McGraw-Hill, 1993.
Ragan, G. L., Microwave Transmission Circuits, vol. 9, M.I.T. Laboratory Series,
McGraw-Hill, pp. 28–29, 1948.
Rulf, B., “Problems of Radome Design for Modern Airborne Radar, Part 1,” Microwave
Journal, Jan. 1985, pp. 145–48, 152–53; “Problems of Radome Design for Modern
Airborne Radar, Part 2,” Microwave Journal, May 1985, pp. 265–67, 271.
Schrank, H. E., G. E. Evans, and D. Davis, in ch. 6, “Reflector Antennas,” pp. 6.44–6.52
in M. I. Skolnik, Radar Handbook, 2nd ed., McGraw-Hill, 1990.
APPENDIX F
Far-Zone Range-Approximation
and Phase Error
The far-field region, that is, the far zone, is the region of an antenna field where the
antenna radiation pattern is essentially independent of the distance from the antenna.
In free space, with the antenna maximum dimension D being large compared to a
wavelength, the far-field region is commonly taken to be greater than rmin, where
rmin = 2 D 2 λ (F–1)
R 2 + ( D 2 ) = (δ + R ) = δ 2 + 2δ R + R 2
2 2
or
( D 2 )2 = δ 2 + 2δ R (F–2)
( D 2 )2 ≈ 2δ R
*While reviewing the manuscript, E. B. Joy added two additional criteria: (1) rmin = 20 l, that pertains to
electrically small antennas and assures that the reactive near-field is negligible at rmin and (2) rmin = 20 D,
that limits the relative down range far-field decrease across the antenna under test to an acceptable value.
As with the factor 2 in (F–1), choice of the factors 20 depends on the accuracy of measurement required.
According to J. S. Hollis, T. J. Lyon, and L. Clayton, Jr., (1970, sec. 14.2), the relevant effects are usually
considered negligible if rmin ≥ 10l and rmin ≥ 10D.
467
468 Far-Zone Range-Approximation and Phase Error
or
δ ≈ ( D 2 )2(1 2 R ) = D 2 8 R (F–3)
d
We can now find the phase error caused
by the path-length difference d if rmin =
R 2D2/l, with d << 2R in accordance with
(F–1). Then from (F–3),
R
D P δ = D 2 8 R = D 2 8 (2 D 2 λ ) = λ 16 = 22.5°
(F–4)
R
Therefore, when the separation distance
is 2D2/l between an aperture (of maximum
dimension D) and the source, the
d maximum phase error is 22.5° (l/16). In
addition, it is to be noted that the phase
error caused by the pathlength difference
d varies in proportion to D2, from zero at
FIGURE F–1.
the aperture center to a maximum at an
Path-length difference d for an aperture aperture edge.
dimension D and range R. A question now to be addressed is the
required minimum far-zone distance
2D2/l, when d is not much less than 2R, the assumption used to get the l/16 far-zone
criterion of (F–4). Returning to the exact equation (F–2) and letting l/16, it may be seen
that D = l/8 if R = 0. Thus, the pathlength difference d of Fig. F–1 is less than l/16 if
D < l/8. In other words, according to the l/16 far-zone criterion, the far zone always
exists if the antenna maximum aperture dimension D is less than l/8. However, it is to
be noted that both reactive and radiative fields surround an antenna (sec. 3.2). Thus,
although the d < l/16 far-field criterion may be satisfied at very short distances from
very small antenna aperture, the reactive near-field may exceed the far-field at separation
distances less than 20l (sec. 3.2.5).
References
Hollis, J. S., T. J. Lyon, and L. Clayton, Jr., sec. 14.2 of Microwave Antenna
Measurements, Scientific-Atlanta, 1970.
The New IEEE Standard Dictionary of Electrical and Electronic Terms, 1992.
APPENDIX G
Appendix G discusses the radiating E- and H-fields that surround an energized antenna,
including radiation patterns versus separation distance from an antenna aperture. This
material is supplemented by the accompanying website that contains example calcula-
tions using Mathcad for radiating near- and far-field patterns, where the Huygens-Kirch-
hoff formulation is used for the near-aperture radiating field calculations (Good 1990).
469
470 Radiating Near and Far Fields, and the Obliquity Factor
2π
−j r1
g(θ , φ ) =
a2 b2 e λ
cos η ( x, y ) + 1 dxdy
∫− a 2 ∫− b 2 A ( x, y ) r1 2 (G–1)
where A(x, y) is the amplitude and phase distribution over the aperture, h(x, y) is the
angle between the outward direction of r1 and the Z axis at each x,y aperture position.
The assumptions include: (1) the r1 and R lengths are at least several wavelengths and
(2) the phase fluctuations across the aperture are small enough so that the phase distribu-
tion is essentially uniform.
Persons familiar with physical optics will recognize (G–1) as being the phasor sum-
mation of Huygens’ wavelets (Fig. 1–9, chapter 1) that emanate from the aperture x,y
positions, with their amplitudes modified based on propagation direction, in accordance
with the factor (1/2)[cos h(x, y) + 1]. This amplitude multiplier is the Fresnel obliquity
factor, which has been verified with mathematical rigor (see, e.g., Slater and Frank 1947).
Accordingly, forward propagating wavelets have unity amplitude if h(x, y) = 0°; and the
The Radiating Fields 471
wavelets have zero amplitude in the backward direction if h(x, y) = 180°. In general,
h(x, y) is a complicated function of the aperture coordinates and q. Fortunately, however,
the obliquity is not needed for calculating array patterns, and often it is not needed when
calculating radiation from continuous aperture antennas.
Another approximation, valid for the far-field and much of the near radiating field,
assumes the lengths r1 and R are large compared to the linear aperture dimensions a and
b. Now, with this, we neglect the variation of cos h over the aperture and we replace h
with q and r1 with R. Then, (G–1) is simplified and becomes (G–2) that follows
2π
( cos θ + 1) a2 b2 −j r1
g(θ , φ ) = ∫− a 2 ∫− b 2 A ( x, y ) e λ dxdy (G–2)
2R
2π 2π 2π 2π
− r1 − (δ + R ) − R − δ
e λ = e λ =e λ e λ
To obtain an equation for use only within the radiating far-field, distance r1 in (G–2) is
replaced by using (H–6) of Appendix H. Then, the integrand of (G–2) is simplified,
R
R q
y z
y f
f
x
(a) (b)
FIGURE G–2.
Spherical coordinates with aperture in x-y plane: (a) shows observation point P at distance R
from origin and (b) shows point P at distance R from the origin and distance r1 from a specific
position on the x-y plane.
472 Radiating Near and Far Fields, and the Obliquity Factor
and the equation for the pattern amplitude in the radiating far-field becomes (G–3) that
follows:
2π
( cos θ + 1) a2 b2 j sin θ ( x cos φ + y sin φ )
g(θ , φ ) = ∫− a 2 ∫− b 2 A ( x, y )e λ dxdy (G–3)
2R
Note that, consistent with the definition of far zone, the integral of (G–3) is not a
function of R.
Recall that q is the observation direction and is measured from the normal to the
aperture (the z-axis). For q of 30°, 60°, and 90°, the obliquity factor reduces the calculated
E-field amplitude by 0.6, 2.5, and 6 dB, respectively. On the other hand, the beam point-
ing direction is determined by the phase term within the aperture excitation A(x, y). Often
the main beam (major lobe) is pointed along the z-axis (q = 0). Then, especially for
narrow-beam antennas, the q values of greatest interest are small enough that the obliq-
uity factor is assumed unity. Therefore, the term (1 + cos q)/2 in (G–3) is commonly
neglected. Furthermore, the equation for g(q, f) is usually simplified by neglecting the
multiplicity factor 1/R. Then the electric field pattern in the radiating far-field may be
expressed as
2π
a2 b2 j sin θ ( x cos φ + y sin φ )
g(θ , φ ) = ∫ ∫ A ( x, y )e λ dxdy (G–4)
−a 2 −b 2
If the limits of integration of (G–4) are replaced by infinity and minus infinity, the
integral becomes a Fourier integral of two variables. Then, for practical purposes (G–4)
is a Fourier integral, because the aperture illumination A(x, y) is zero outside the aperture.
Thus, the integral in (G–4) vanishes beyond the aperture dimensions ±a/2 and ±b/2. With
this in mind, the reader will recognize that the far-zone pattern of an aperture antenna is
the Fourier transform of its aperture illumination. This Fourier-transform relationship
was especially important for antenna engineers before the general availability of digital
computers. Then, calculations for patterns based on aperture distributions were made
through use of published Fourier transform tables. Nowadays, the fast Fourier transform
(FFT) is useful for pattern calculations that include a large number of aperture data
points.
The Huygens’ concept of summing wavelets is applicable to arrays, but the obliquity
factor is replaced by unity. For a continuous aperture, the summation is of the wavelets
within a plane wave that propagates in the +z direction, and the obliquity factor provides
for no propagation in the −z direction. For array factors (patterns without inclusion of
element patterns), the summation is of a finite number of wavelets that propagate isotro-
pically from each element, and thus the obliquity factor does not apply to arrays.
Therefore, to get array factor far-field patterns, one may use (G–1) through (G–4) by
replacing (a) the integrals with discrete summations and (b) the obliquity factors with
the numeral one.
Principal-Plane Patterns Versus Range 473
2π
−j r1
λ
cos η ( n ) + 1
N
e
g(θ ) = ∑ a (n) r1 2 (G–5)
n =1
Ordinarily, r1 is large compared to the dimensions of the antenna, and then the bracketed
term in (G–5) becomes (cos q + 1)/2. However as noted in sec. G–1, for array elements
the (cos q + 1)/2 term is inapplicable and is replaced by numeral one.
In general, pattern calculations are highly sensitive to r1 and the pathlength difference
d, where d is r1 − R. For r1, equation (H–4), Appendix H, is exact. Then, with radiating
elements located only at positions x(n) on the x-axis, r1 becomes
0.5
x (n) x (n)
2
r1 = R 1 − 2 sin θ +
R
(G–6)
R
δ = r1 − R ≈ − x ( n ) sin θ (G–7)
Now, using (G–6) and (G–7) when R is much larger than r1, one obtains (G–8) that
follows.
2π
( cos θ + 1) N j x ( n ) sin θ
g(θ ) =
2R
∑ a( n ) e λ (G–8)
n =1
474 Radiating Near and Far Fields, and the Obliquity Factor
Equation (G–8) may be used for far-field calculations, provided R is 2D2/l or greater.
As in the previous equations, the obliquity factor (cos q + 1)/2 for arrays is replaced by
numeral one. In addition, this factor is often assumed to be unity for aperture antennas;
because major interest is usually near small q values, where the obliquity factor is essen-
tially one.
Example pattern calculations versus range are included in the accompanying
website.
References
Good, Myron L., “Diffraction,” in Encyclopedia of Physics, 2nd ed., R. Lerner and G. L.
Trigg (eds.), 1990, VCH Publishers, p. 252.
Hansen, R. C., “Aperture Theory,” in R. C. Hansen (ed.), Microwave Scanning Antennas,
Academic Press, Inc, New York, 1964.
IEEE Standard 100-1992, The New IEEE Standard Dictionary of Electrical and Electronic
Terms, 1992.
Johnson, R. C. (ed.), Antenna Handbook, 3rd ed., p. 1–10, McGraw-Hill, 1993.
Johnson, R. C., H. A. Ecker, and J. S. Hollis, “Determination of Far-Field Antenna Patterns
from Near-Field Measurements,” Proceedings of the IEEE, December 1973, pp.
1668–94.
Silver, S., Microwave Antenna Theory and Design, McGraw-Hill, 1949.
Slater, J. C., and N. H. Frank, Electromagnetism, 1947, Mc-Graw-Hill, ch. 13.
APPENDIX H
x p = R sin θ cos φ ≡ Rα
y p = R sin θ sin φ ≡ Rβ (H–1)
z p = R cos θ
In (H–1) above, the abbreviations Ra and Rb are used to denote xp and yp to simplify
equations that follow. Notice that
12 12
R = ( x p ) + ( y p ) + ( z p ) = ( Rα ) + ( Rβ ) + ( R cos θ )
2 2 2 2 2 2
(H–2)
Next, referring to Fig. H–1(b), we find the distance r1 between points at xa and ya on the
aperture (za = 0) and point P is as follows
12 12
r1 = ( xa − x p ) + ( ya − y p ) + ( z p ) = ( Rα − xa ) + ( Rβ − ya ) + ( R cos θ )
2 2 2 2 2 2
(H–3)
After expanding (H–3) and noting from (H–2) that [(Ra)2 + (Ra)2 + (R cos q)2] = R2, we
find
12
2 x 2 + y2
r1 = R 1 − (α xa + β ya ) + a 2 a (H–4)
R R
Now assume that the aperture dimensions are small compared to R, that is, xa << R
and ya << R. Then, r1 of (H–4) is simplified and becomes (H–5) that follows
475
476 Path Length Differences from a Planar Aperture
R
R q
y z
y f
f
x
(a) (b)
FIGURE H–1.
Spherical coordinates with array elements in x-y plane. In (a), point P is at distance R from
origin; in (b), point P is at distance R from the origin and distance r1 from a position xa,ya on
the x-y plane.
12
r1 ≈ R ⎡⎢1 − (α xa + β ya )⎤⎥ ≈ R ⎡⎢1 − (α xa + β ya )⎤⎥
2 1
⎣ R ⎦ ⎣ R ⎦
= R − xa sin θ cos φ − ya sin θ sin φ (H–5)
Now consider effects on r1 and d with R fixed, and with P only on one of the principal
planes, x-z or y-z. Then, for the x-z and y-z planes: f = 0 and f = p/2 (90°), respectively.
Thus, from (H–6), with R fixed and P in the x-z plane
Now assume the x and y axes are vertical and horizontal, respectively, as in Fig.
H–1(b). Then from (H–7) and (H–8), which are applicable when xa << R and ya << R,
we see that to close approximations the following are true:
(1) for P on the vertical plane (x-z), r1 and d are functions of the xa (vertical)
location on the aperture—not on the ya (horizon) location; and similarly
(2) for P on the horizontal plane (y-z), r1 and d are functions of the ya (horizon-
tal) location on the aperture—not on the xa (vertical) location.
Path Length Differences from a Planar Aperture 477
In arrays, aperture phase errors are caused by element misalignment and improper phase
excitation, and surface errors are a major source of aperture error for reflector antennas.
Section 6.5 includes a general discussion on reflector errors, and the present appendix
addresses relatively simple calculations on the effects of random phase errors.
In Fig. I–1, it may be seen that the difference DR in the pathlengths aa′ and bb′ is
∆ R = 2h sin α (I–1)
Thus the phase difference in radians DY caused by the path length difference DR between
paths aa′ and bb′ is
∆ψ = ( 2π λ ) ∆ R (I–2)
Then, if the height distribution for a random rough surface is designated by Dh, the
distribution of phase differences can be expressed as
∆ψ = ( 2π λ ) 2 ∆hsin α (I–3)
Experience shows that random surface roughness is well described by surface heights,
which can be expressed statistically as a normal (i.e., Gaussian) distribution. Then, in
Mathcad, the surface height distribution Dh becomes
∆h = rnorm ( n, µ, σ ) (I–4)
where
n = number of data points
m = average value
s = standard deviation
If m = 0, the standard deviation s equals the rms of the distribution in height.
479
480 Effects of Random Aperture Phase Errors
α = (180° − β ) 2 (I–5)
xi = (i − 200 )10 −2
Effects of Random Aperture Phase Errors 481
Then, using the general relationship from array theory for the amplitude of the electric
field (sec. 5.5.1), for the far zone we have
{ 2π
}
401
E (θ ) = ∑ ai exp j α i + xi sin θ (I–6)
i=0 λ
where
ai = the amplitude of the ith element
ai = phase of the ith element
We now make the following additional assumptions, namely, that (1) amplitude is con-
stant over the aperture with each ai = 1 and (2) the phase is uniformly and normally dis-
tributed over the xi.
Figure I–3 shows results of assuming rms phase errors of l /16 and l /8 (22.5° and
45°), and comparing those results with zero phase error (for the calculations, see the
accompanying SM). Note that the pattern peak levels are reduced with increases in
–5
–10
Field Strength in dB
–15
–20
–25
–30
–9 0 9
Off-Broadside Angle in Degrees
the rms errors. Additionally, the loss in gain caused by l /16 rms phase error is only a
fraction of a dB, but it is discernible. On the other hand, because of the randomness in
phase error and the resulting randomness in side-lobe levels, the patterns are not sym-
metrical between the negative and positive angles from the major lobe.
Reference
Long, M. W., Radar Reflectivity of Land and Sea, 3rd ed., Artech House, 2001, p. 52.
Index
483
484 Index
antennas receiving, 30
array, 169 resonant, 139
broad-band, 285 standard gain, 397
current and voltage in longer, 132–136, structures of, 89–91
133 fig.–135 fig. surface-wave, 159
defined, 1, 87 synthetic-aperture, 285, 324–325
direction-finding, 310–314, 312 fig., transmitting vs. receiving, 30–31
313 fig. antennas, long-wire
electrically small, 308–310 classification of, 138–139
with extremely low side lobes, directivity of, 141
325–327 effect of ground on, 143
fractals, 328–330, 330 fig., 331 fig. polarization of, 144–145
frequency-independent, 293–301, 295 radiation patterns of, 138–141, 140 fig.,
fig., 296 fig., 298 fig.–301 fig. 142, 142 fig.
frequency-scan, 338, 341, 339 fig. radiation resistance of, 141, 141 fig.
helical, 148–149, 148 fig. uses of, 143
ideal, 87 antenna-under-test (AUT), 385
leaky-wave, 159 aperture, of antenna, 269
loop, 144–148, 144 fig., 147 fig. circular, 279, 279 fig.
low-noise receiving, 320–324, 323 fig. of parabolic aperture, 240
mechanical scan, 314–321, 316 fig., shape of, 233
318 fig.–320 fig. aperture analysis, 329
microstrip, 157–159 continuous line source in, 273–276,
for multiple polarization 273 fig., 277 fig., 278 table
for circular polarization, 302–305, general, 274
303 fig. historical note on, 270
corrugated horn feeds, 307, 307 fig. pattern calculations for continuous,
dual-mode transducers, 306, 306 fig., 269, 271, 273, 275, 277, 279
307 fig. pattern vs. Q and F in, 279, 279 fig.
polarizers, 305–307, 307 fig. principal plane patterns in, 279, 278 table
nonresonant, 142, 142 fig. separable x and y distributions in, 275,
omnidirectional, 307–308 277, 277 fig.
parameters for, 87 aperture dimensions
aperture dimensions, 113–115 and beamwidth, 119
bandwidth, 111–113 interdependent with gain, 113–114, 116
beamwidth, 105–107 aperture distribution, 151
directivity and gain, 98–103 calculating, 209
effective area, 103–105 defined, 236
input impedance, 110–111 Dolph-Chebyschev (D-C) distribution,
minor lobes, 107–108, 108 fig. 203
polarization, 112 planar and volume arrays with uniform,
radiation resistance and efficiency, 190–191, 193, 191 fig.
108–109 for reflector antenna, 251, 255, 255 fig.
patch, 157–159 Taylor distribution, 239
Index 485
Cassegrain feed system, 245–246, 244 fig. decibels, space attenuation in, 13
and noise reduction, 323–324 delta-match feed, 159, 159 fig.
Index 487
feed methods for, 159, 162, 161 fig. input impedance, 110, 111
radiation patterns for, 151, 152, 154, of basic folded dipole, 291
155, 157 calculation of, 414, 415
horns, corrugated, 307, 307 fig. for electrically small antennas,
Howells-Applebaum method, of pattern 308–311
optimization, 360–362 equations for, 59, 62
Huygens-Kirchhoff formulation, 469, 470 and feed methods, 159, 162
Huygens’ principle, 2, 25–26, 26 fig., measurement of, 389–391
151, 227 of short dipoles, 126–129, 127 fig.
Huygens’ wavelets, 470, 472 Institute of Electrical and Electronics
hybrid couplers, 80–82, 82 fig. Engineers (IEEE), 43, 450, 460
Instrument Low-Approach System (ILS),
IEEE. see Institute of Electrical and 314, 319
Electronics Engineers, Inc. insulators
illumination protection of, 90, 91
defined, 236 uses for, 83
reflectors, 227, 229–252, 228 fig., 230 intercontinental ballistic missile (ICBM)
fig., 235 fig., 243 fig., 244 fig. detection and analysis, 366
ILS. see Instrument Low-Approach interference, 15, 394
System and antenna applications, 20, 23, 25
images, principle of, 21, 21 fig., 22, effects of, 29
129 problem of, 53
impedance charts, 404 interferometer system, 314
impedance match International Telecommunications Union
in broad-band design, 285–290, 292, (ITU), 3–4
293, 296 inverse-square law, 10
between transmitter and free space, and attenuation, 12–14
221, 223 and power density, 8, 10–12
impedance measurements, 401–404 and radiation process, 28
network analyzers for, 391, 400–404, spherical waves and, 8
403 fig. ionosphere
standing-wave method, 401, 404, and long-distance radio propagation,
401 fig. 33, 34
impedances and radio waves, 2, 40–41
in array theory, 169, 171 ionospheric scatter, 34
characteristic, 51, 53, 55, 71, 72, 76 irises, reactance introduced into
mutual, 177 waveguides by, 77, 77 fig.
impedance transformation, analysis of, Isbell log-periodic dipole array, 299, 230,
65, 67, 69 305
impedance transformer isotrope, defined, 8
single-stub, 66, 66 fig. isotropic point source, 169–171,
transmission line as, 54, 55, 64, 65 171 fig.
induction field, 28 ITU. see International
infinitesimal dipole, 120 Telecommunications Union
492 Index
Maurice Long’s interest in antennas and propagation began as a teenager with his amateur
radio station W4GPR. Shortly after World War II, he began research work at Georgia
Institute of Technology on microwave propagation, and that was followed by antenna
developments for microwave and millimeter radar. Activities in multipolarization and
rapid scan antennas lead to his being General Chairman of the Department of Defense
sponsored 1956 Georgia Tech/SCEL Symposium on Scanning Antennas. Subsequently,
much of his work at Georgia Tech focused on antennas, electromagnetic scattering, and
microwave and millimeter radar. Later, as a consultant to Lockheed Martin and other
companies, he worked in radar design and analysis, including the use of large airborne
mechanical and electronic scanning antennas.
At Georgia Tech he held research and academic positions, including Principal Research
Engineer, Professor of Electrical Engineering, Associate Graduate Dean for Research,
and Director of the Engineering Experiment Station (now Georgia Tech Research Insti-
tute). Presently he teaches graduate courses part time at Southern Polytechnic State
University in antenna design and radar systems, works as a radar consultant, and main-
tains an affiliation as a retiree with the Georgia Tech Research Institute.
Dr. Long’s previous books include Radar Reflectivity of Land and Sea, 3rd ed., Artech
House and Airborne Early Warning System Concepts, SciTech Publishing. He is a Life
Fellow of the Institute of Electrical and Electronics Engineers; and member of Academy
of Electromagnetics and Commission F of International Union of Radio Science.
Additional personal information is available in Who’s Who in America, Who’s Who in
Engineering, American Men and Women in Science, and McGraw-Hill Leaders in
Electronics.
503