Electromagnetics
Electromagnetics
Antennas
Mario Boella Series on Electromagnetism in Information
and Communication
This publication is copyright under the Berne Convention and the Universal Copyright
Convention. All rights reserved. Apart from any fair dealing for the purposes of research
or private study, or criticism or review, as permitted under the Copyright, Designs and
Patents Act 1988, this publication may be reproduced, stored or transmitted, in any
form or by any means, only with the prior permission in writing of the publishers, or in
the case of reprographic reproduction in accordance with the terms of licences issued
by the Copyright Licensing Agency. Enquiries concerning reproduction outside those
terms should be sent to the publisher at the undermentioned address:
The Institution of Engineering and Technology
Michael Faraday House
Six Hills Way, Stevenage
Herts, SG1 2AY, United Kingdom
www.theiet.org
While the authors and publisher believe that the information and guidance given in this
work are correct, all parties must rely upon their own skill and judgement when making
use of them. Neither the authors nor publisher assumes any liability to anyone for any
loss or damage caused by any error or omission in the work, whether such an error or
omission is the result of negligence or any other cause. Any and all such liability is
disclaimed.
The moral rights of the authors to be identified as authors of this work have been
asserted by them in accordance with the Copyright, Designs and Patents Act 1988.
1 Introduction 1
References 3
3 History 17
3.1 The early years 17
3.2 The golden years 17
3.3 Waveguide slot antennas 18
3.4 The many shapes of slotted waveguide array antennas 18
References 23
5 Slot models 59
5.1 Modelling principles 59
5.1.1 Using experimental data 59
5.1.2 Equivalent circuit approach 60
5.1.3 Electromagnetic models 62
5.1.4 Finite wall thickness 66
5.2 Integral equation solution 68
5.2.1 The internal field 68
5.2.2 The external field 70
5.2.3 Matrix elements 71
5.3 Longitudinal slot characteristics 72
5.3.1 Slot electric field distribution 72
5.3.2 Slot magnetic field distribution 76
5.3.3 Experimental verification 79
5.3.4 Weakly excited slots 82
5.4 Transverse slots 83
5.4.1 Introduction 83
5.4.2 Theoretical analysis 84
5.4.3 Results 86
References 88
Index 341
About the authors
Lars Josefsson and Sembiam Rengarajan met at the University of California, Los
Angeles (UCLA), USA, where both worked on slotted waveguide array antennas
together with Professor R.S. Elliott, a well-known scientific leader and teacher in
this field. Since then they have kept contact and collaborated on research projects
and courses several times.
Lars Josefsson was born in Norrköping, Sweden. He graduated from The Royal
Institute of Technology in Stockholm (KTH) and got his PhD from Chalmers
University of Technology in Göteborg. A grant from the Marcus Wallenberg
Foundation and a Fulbright Scholarship enabled his stay in 1982–83 as a Visiting
Scientist at UCLA, working on waveguide slotted array antennas. He became
Adjunct Professor in Antenna Technology (1983–86) at Chalmers and obtained the
equivalent position at KTH (1996–2003). He returned to Chalmers again as
Adjunct Professor (2004–07).
For more than 40 years Lars was with Ericsson Microwave Systems AB in
Mölndal, Sweden, where he became responsible for new antenna technology and
systems, internal R&D projects and internal education relating to antennas. In 2004
he received the Thulin Silver Medal from the Swedish Aeronautical Society for his
contributions to several generations of airborne radar antenna developments. He is
fellow of the IEEE since 1999.
In 2003 he formed his own consulting company: Lars Microwave (technical
consulting, research and education). Lars Josefsson is the author and co-author
of more than 100 scientific publications. He is the author (together with Patrik
Persson) of the book Conformal Array Antenna Theory and Design, published by
IEEE/Wiley in 2006.
Sembiam R. Rengarajan received the PhD degree in Electrical Engineering
from the University of New Brunswick, Canada, in 1980. Since then he has been a
faculty member at California State University, Northridge, USA. He has held vis-
iting appointments at Chalmers University of Technology, Sweden; University of
Santiago de Compostela, Spain; University of Pretoria, South Africa; and Technical
University of Denmark. In addition to teaching electromagnetics and antenna
courses, he has performed significant research in topics on applications of elec-
tromagnetics to antennas, scattering, and microwave components. He has also
consulted for many companies and government agencies. He has published more
than 240 journal articles and conference presentations.
Much of Sembiam’s research has dealt with slotted waveguide array antennas.
He collaborated with Prof. Elliott in the 1980s and 1990s for more than a decade.
xii Slotted waveguide array antennas: theory, analysis and design
His research in this area was supported by the University of California, Hughes
Aircraft Company, Lockheed Martin, Saab Ericsson Space, European Space
Agency, Rantec Corporation, DSO Singapore, NASA/Jet Propulsion Laboratory
(JPL), among others. He served as a Distinguished Lecturer for IEEE Antennas and
Propagation Society, lecturing on slotted waveguide array antennas all over the
world. Presently he serves as the Chair of the United States National Committee for
the International Union of Radio Science (USNC-URSI). He has received more
than 20 awards from NASA for his research contributions to JPL. He is a life fellow
of IEEE.
Abbreviations and acronyms
AF array factor
CEM computational electromagnetics code
CFRP carbon fibre-reinforced plastic
dB decibel
DBF digital beam forming
DOA direction of arrival
EM electromagnetic
FDTD finite difference time domain
FEM finite element method
FSS frequency selective structure
GA genetic algorithm
GHz giga hertz
GO geometrical optics
GSM generalised scattering matrix
GTD geometric theory of diffraction
HFSS high-frequency structural simulator
HTCC high temperature co-fired ceramics
IE integral equation
LEO low earth orbit
LMS least mean square
LTCC low temperature co-fired ceramics
MC mutual coupling
MIMO multiple input multiple output
MIT Massachusetts Institute of Technology
MMIC monolithic microwave integrated circuit
MM mode matching
MoM method of moments
MTI moving target indication
NDT non-destructive testing
xiv Slotted waveguide array antennas: theory, analysis and design
Symbol Unit
Symbol Unit
The Mario Boella series is devoted to textbooks and monographs in all areas of
Radio Science, with a particular emphasis on the applications of electromagnetism
to information and communication technologies. The series is financially spon-
sored by the Istituto Superiore Mario Boella (ISMB), a research entity affiliated
with the Politecnico di Torino, Italy, and is scientifically co-sponsored by the
International Union of Radio Science (URSI). It is named to honour the memory of
Professor Mario Boella of the Politecnico di Torino, who was a pioneer in intro-
ducing electronics and telecommunications into higher education in Italy for half a
century and was a vice-president of URSI from 1966 to 1969.
This research monograph on slotted waveguide antennas is the eighth volume
in the ISMB series. It contains a very comprehensive coverage on a subject that is
of practical importance in antenna and radar systems for a variety of civilian and
military applications. Both theory and design, as well as manufacturing aspects and
future applications, are discussed in depth. The authors are well-known authorities
in the field. Dr Josefsson has decades of experience as a design and project engi-
neer, and has been honoured for his technical achievements. Professor Rengarajan
has taught and done research on antennas and radars for many years; he is highly
regarded in the electromagnetics community for his original contributions. The two
authors have combined their complementary knowledge to produce a monograph
that will be very useful to students, instructors and practitioners.
Piergiorgio L. E. Uslenghi
ISMB Series Editor
Chicago, April 2018
Preface
1500
1000
500
0
1980 1990 2000 2010 2020
Year
Number of hits in IEEE Xplore from 1946 to ‘year’ for the key word ‘slotted
waveguide array antenna’
xx Slotted waveguide array antennas: theory, analysis and design
How do other antenna types such as microstrip patch array antennas and
reflector antennas compare with slotted waveguide array antennas? Without going
into details we try a simple comparison:
Antenna type Characteristics
Microstrip patch arrays Transmission losses, surface waves
Reflector antennas Fixed beam, require depth
Slotted waveguide arrays þ Low losses, high precision
There are of course both pros and cons for all types, depending on the appli-
cation. Several books discuss microstrip antennas and/or reflector antennas. The
present book is the first totally devoted to slotted waveguide array antennas.
With specific examples of waveguide array designs, accompanied by detailed
illustrations and antenna characteristics, the book is a must-have reference for
engineers involved in the design and development of slotted waveguide array
antennas. An additional value is provided by a thorough overview of the related
literature in the field and the history of these antennas given in several chapters.
With this book system engineers and design and development engineers in industry
and government will have a single source with a coherent treatment instead of
turning to numerous papers in the literature.
The book goes well beyond some of the commonly discussed topics on slotted
waveguide array design, into areas that include
● higher-order mode coupling and edge effects
● performance optimisation in terms of bandwidth and pattern performance
● special slot excitation methods
● applications such as monopulse and phase steering
● manufacturing techniques and tolerances, etc.
The book is organised as follows:
The first few pages contain a list of abbreviations and acronyms that are
common when dealing with slotted waveguide array antennas and related appli-
cations. There is also a list of symbols with definitions, representing mathematical
and physical quantities that appear in the text. Finally, a table providing definitions
of waveguide bands and designations, with corresponding rectangular waveguide
inner dimensions is included. The following chapters all start with some intro-
ductory material such as the basic concepts that are essential to get an under-
standing of the more advanced concepts to follow.
Chapter 1, ‘Introduction’, is a short overview of the radiating mechanism of
slots in waveguides carrying the fundamental TE10 mode. There is also a short
discussion of equivalent circuits for various types of resonant slots: longitudinal,
transverse and inclined slots.
Chapter 2, ‘Review of electromagnetic theory, starts from Maxwell’s equa-
tions. Important concepts such as boundary conditions, reciprocity theorems, field
equivalence and Green’s functions and other definitions are included. The purpose
is to serve as a reference to be consulted if the need arises when reading the
theoretical chapters that follow.
Preface xxi
Much pioneering work in the field of slotted waveguide array antennas was done
at the Radiation Laboratory at the Massachusetts Institute of Technology (MIT)
during the years 1940–45. The Radiation Laboratory Series of publications
(28 volumes) remain a valuable source of reference even today.
The book draws on material from several short courses on ‘Slotted Waveguide
Array Antenna Technology’ by the authors. One such course was presented toge-
ther with Dr M Ando and Dr J Hirokawa from the Tokyo Institute of Technology.
Numerous colleagues and friends have supported our work with their advice,
encouragement and contributions. In particular we would like to mention Dr Hans
Steyskal and Dr Bob Mailloux for their valuable comments.
We also remember with gratitude Professor Bob Elliott of UCLA, a mentor
and friend, who introduced us to many exciting aspects of slotted waveguide arrays.
We are grateful for the permission to use results from several studies on slotted
waveguide arrays. This includes Ericsson AB, Saab AB, RUAG Space AB, in
Sweden and many more.
One of us (SR) would like to acknowledge his research sponsors including
Hughes Aircraft, University of California, Saab Ericsson Space, JPL, Rantec,
Lockheed Martin, DSO Singapore and CSUN Foundation, and his many students
and colleagues who have contributed to his work.
Last but not least, we express our appreciation and gratitude to our families for
their encouragement, understanding and patience during the writing of this book.
Chapter 1
Introduction
The first successful slotted waveguide array antennas were developed in Canada
during the Second World War. The immediate applications were in military ground
and airborne radar systems for target detection and tracking. Later applications
include remote sensing from aircraft and space vehicles and microwave commu-
nication links. Spaceborne synthetic aperture radar (SAR) with slotted waveguide
arrays are used for weather forecasting, environmental monitoring, climate change
studies, etc. The slotted array antenna is also considered in automobile collision
avoidance systems.
For the interested reader not too familiar with slotted waveguide array anten-
nas a few simple concepts are introduced in the following text.
Slots in rectangular waveguides are the most commonly used. The wave-
guide can be standard dimension, or half height or even quarter height, in order to
save space and weight. Many other realisations are possible, however, as will
be discussed in later chapters. The fields inside the waveguide are related to the
currents in the waveguide walls. For the fundamental TE10 mode we have the
current distribution in the waveguide wall as shown in Figure 1.1. Typical slot
positions are illustrated in Figure 1.2. By comparing the two figures we realise
that slot positions c and e are not very useful since these slots do not interrupt any
wall current and hence do not couple to the waveguide mode. Slots a (transverse
Figure 1.1 Surface currents on the waveguide wall for the TE10 mode in a
rectangular waveguide
2 Slotted waveguide array antennas: theory, analysis and design
d
b
q
c a
f
g e
q
Figure 1.2 Possible slot positions in a rectangular waveguide.
Adapted from [1]
broad wall slot) and g, on the other hand, couple strongly, and slot f (inclined
sidewall slot) couples to a degree depending on the inclination angle q. Slot b
(longitudinal broad wall slot) also couples strongly when displaced far from the
waveguide centreline.
The rectangular slot in a large conducting ground plane radiates approximately
as a dipole of same length and width. Thus, the slot in a waveguide has a broad
antenna pattern approximately equal to the corresponding dipole, only that E- and
H-planes are interchanged. The corresponding antenna impedances differ con-
siderably, however (Babinet’s principle) [2, p. 336].
A. F. Stevenson at the National Research Council in Canada [3] found
suitable equivalent networks for the most common slots (at resonance) as shown in
Figure 1.3.
A more detailed analysis of waveguide modes and slot characteristics will be
presented in subsequent chapters. Also array design principles and applications will
be discussed.
New fabrication techniques for slotted arrays such as electroplating on
dielectric materials have demonstrated good performance at frequencies as high
as 100 GHz. Thus, highly directive antennas with small dimensions can be
made, and a very large signal bandwidth becomes possible. Applications at such
high frequencies are mainly for short ranges: covert communication, imaging
radar, collision avoidance, planetary landing, etc. An important step forward is
the possibility to integrate active microelectronics and circuits with the antenna
structure.
Important progress over the recent years has also been made in the design and
optimisation of slotted waveguide array antennas. It is now possible to fabricate
planar arrays in a cost-effective and simplified way in several layers using new
techniques; there are new ways to match and optimise the electrical design,
increase bandwidth, etc.
Introduction 3
x1
q
Inclined slot in the narrow wall
b g + jb
Equivalent shunt element
References
[1] Silver S. ed. Microwave Antenna Theory and Design, MIT Rad. Lab. Series,
Vol. 12, McGraw-Hill, New York, 1949.
[2] Elliott R. S. Antenna Theory and Design. Prentice-Hall, Englewood Cliffs,
NJ, 1981.
[3] Stevenson A. F. Series of Slots in Rectangular Waveguides, National
Research Council of Canada, Radio Reports 12 and 13, 1944.
Chapter 2
Review of electromagnetic theory
In this chapter, we review the electromagnetic theory and concepts used in later
chapters. We will start with Maxwell’s equations in time harmonic form, followed
by boundary conditions. Expressions for energy and power are derived. We then
discuss the reciprocity theorem. Vector and scalar potentials are derived. Image
principle and the field equivalence principle are presented. Green’s functions are
discussed with a presentation of dyadic Green’s functions for the magnetic current
in a waveguide.
r~ ~
E ¼ jwmH (2.1)
~ ¼ jwe~
rH E þ~
J (2.2)
r~
D ¼ rv (2.3)
r~
B¼0 (2.4)
In these equations ~ ~, D
E, H ~ and ~B are the electric field, magnetic field, electric
displacement density and magnetic flux density, respectively. e and m are the per-
mittivity and permeability, respectively. In lossy media e ¼ e0 je00 and m ¼ m0 jm00 .
The sources ~ J and rv are the electric current density and volume charge density,
respectively. The current density term ~ J may consist of both induced currents and
impressed currents.
The constitutive relations are given next, where s is the conductivity of the
medium.
~ ¼ e~
D E (2.5)
~ ~
B ¼ mH (2.6)
~
J ¼ s~
E (2.7)
6 Slotted waveguide array antennas: theory, analysis and design
^ ~1 ¼ ~
nH Js (2.14)
1 n^
2
Figure 2.1 Two regions with a boundary surface separating them
Review of electromagnetic theory 7
and
^ ~ 1 ¼ rs
nD (2.15)
If medium 2 is a perfect magnetic conductor (PMC), introduced in some
equivalent field problems,
n ~
^ ~s
E1 ¼ M (2.16)
and
n ~
^ B 1 ¼ rms (2.17)
where rms is the equivalent surface magnetic charge.
r~ ~ M
E ¼ jwmH ~i (2.18)
~ ¼ jwe~
rH E þ~
Jc þ~
Ji (2.19)
Let us consider the expression
r ~ E H~ ¼ H ~ r ~
E ~ ~
E rH (2.20)
where the superscript * denotes the complex conjugate.
Both sides of (2.20) are integrated over a volume v, and the divergence theo-
rem is applied to obtain
h i
∯ ~ EH ~ d~s¼∭ H ~ r~ E ~ E rH ~ dv (2.21)
s v
dissipated power and Ps is the complex power delivered by the sources. If m and e
are complex, additional dissipation terms will be present.
r~ ~a M
E a ¼ jwmH ~a (2.24)
~ a ¼ jwe~
rH Ea þ ~
Ja (2.25)
r~ ~b M
E b ¼ jwmH ~b (2.26)
~ b ¼ jwe~
rH Eb þ ~
Jb (2.27)
Using vector identities, (2.28) and (2.29) are obtained.
r ~ ~b ¼ ~
Ea H ~b H
Ea r H ~b r ~
Ea
¼ jwe~
Ea ~
Eb þ ~
Ea ~ ~a H
J b þ jwmH ~b þ H
~b M
~a (2.28)
r ~ ~a ¼ ~
Eb H ~a H
Eb r H ~a r ~
Eb
¼ jwe~
Ea ~
Eb þ ~
Eb ~ ~a H
J a þ jwmH ~b þ H
~a M
~b (2.29)
Integrating (2.30) over a volume containing all sources and by using the divergence
theorem we obtain
∯ ~ ~b ~
Ea H ~ a d~
Eb H s¼∭ ~Ea ~ ~a M
Jb H ~b ~
Eb ~ ~b M
Ja þH ~ a dv
s v
(2.31)
Using the reaction concept [2] the above equation may be stated as
ha; bi ¼ hb; ai (2.35)
The term ha,bi is called the reaction of ‘a’ fields with ‘b’ sources. Reciprocity
theorem has been widely applied in microwave circuits, antennas and scattering
problems.
r~ ~
E ¼ jwmH (2.36)
~ ¼ jwe~
rH E þ~
J (2.37)
Since r ~
B ¼ 0 everywhere ~
B can be expressed as a curl of a vector.
Let ~ ~ ¼ r~
B ¼ mH A (2.38)
where ~
A is the magnetic vector potential [3]. From (2.36) and (2.38) we obtain
r ~ E þ jw~
A ¼0 (2.39)
Since the curl of a gradient of any scalar function is zero, we express the electric
field in terms of ~
A and an electric scalar potential f, as given next.
E þ jw~
~ A ¼ rf E ¼ jw~
or ~ A rf (2.40)
1 h i
r r~
A ¼ jwe jw~
A rf þ J (2.41)
m
10 Slotted waveguide array antennas: theory, analysis and design
Assuming that the medium is homogeneous the above equation may be rewritten as
A ¼ k 2~
rr~ A jwmerf þ mJ (2.42)
r 2~
A þ k 2~
A ¼ m~
J (2.43)
By performing the divergence operation on (2.40) one can derive the scalar wave
equation
r
r2 f þ k 2 f ¼ (2.44)
e
~ from
One can solve (2.43) for the given source distribution and then determine H
(2.38) and E from
r~A
E ¼ jw~
~ Aþr (2.45)
jwme
r~ ~ M
E ¼ jwmH ~ (2.46)
~ ¼ jwe~
rH E (2.47)
H ¼ jw~
F rjm (2.48)
where jm is the scalar magnetic potential
r2 ~
F þ k 2~ ~
F ¼ eM (2.49)
and
~ ¼ jw~ r~F
H F þr (2.50)
jwme
Review of electromagnetic theory 11
J M
M
J
PEC
(a)
J M
J M
J
M
J
M
(b)
Figure 2.2 Application of the image principle: (a) electric and magnetic
current sources above an infinite planar PEC; (b) equivalent
problem for the upper half space
12 Slotted waveguide array antennas: theory, analysis and design
II e0, m0
E, H M1
J1
I e, m S
(a)
II e0, m0
E, H
E =0=H n^
I e0, m0 S Js1 = n^ × H
Ms1= –n^ × E
(b)
Figure 2.3 (a) Geometry of the original problem. (b) Love’s equivalence for
region II. (a, b) IEEE 2000, reprinted with permission from [6]
Review of electromagnetic theory 13
It is noted that the vector sign in the Green’s function is left out since the direction
of the vector potential is the same as that of the current density vector. The volume
integral is carried out over the entire region of the current distribution. Alter-
natively we can express the Green’s function as a dyad, noting that only the
diagonal terms in the dyad will contribute to the free space dyadic Green’s func-
tion, that is,
0
~ ~
0 ejkjRR j
G0 ~R; ~
R ¼ 0 I (2.55)
4pj~
R ~ Rj
satisfies
0 0 0
r2 G 0 ~R; ~
R þ k2G0 ~ R ¼ I d ~
R; ~ R ~
R (2.56)
The dyadic Green’s function is readily available [8] and is given next for the time
harmonic field with time dependence in the form exp( jwt).
0 we X X ð2 d0 Þ
G1 ~R; ~
R ¼ exp½jbmn ðz z0 Þ
ab m n bmn k 2 b2mn
n o
1 mp np
2 jbmn sx cy ^x jbmn cx sy ^y þ k 2 b2mn cx cy ^z
k a b
n mp 0 0 np o
jbmn sx cy ^x jbmn cx0 sy0 ^y þ k 2 b2mn cx0 cy0 ^z
a b
nnp mp onnp mp 0 0 oi
þ sx cy ^x cx sy ^y sx0 cy0 ^x cx sy ^y
b a b a
jwe 0
þ 2 ^z^z d ~ R ~ R
k
(2.62)
where m and n vary from 0 to ? in the double summation except for m ¼ 0 and
n ¼ 0, sx ¼ sin(mp x/a), sx0 ¼ sin(mp x0 /a), cx ¼ cos(mp x/a), cx0 ¼ cos(mp x0 /a),
sy ¼ sin(np y/b), sy0 ¼ sin(np y0 /b), cy ¼ cos(np y/b), cy0 ¼ cos(np y0 /b), in and
the upper sign is used if z > z0 while the lower sign is appropriate for z < z0 . d0 ¼ 1
Review of electromagnetic theory 15
For the case of a slot in the broad wall (y ¼ 0), cy ¼ 1 and sy ¼ 0, whereas if y ¼ b,
cy ¼ (1)n and sy ¼ 0. If y ¼ y0 ¼ 0 or b, the Green’s function reduces to the fol-
lowing simple form:
0
we X X
G1 ~ R; ~
R ¼ 2 ð2 d0 Þexp½jbmn ðz z0 Þ½gxx ^x ^x þ gxz^x^z þ gzx^z^x þ gzz^z^z
k ab m n
jwe ~ ~0
þ ^z^z d R R
k2
(2.63)
where
h i
k 2 ðmp=aÞ2
gxx ¼ sinðmpx=aÞsinðmpx0 =aÞ;
bmn
gxz ¼ jðmp=aÞsinðmpx=aÞcosðmpx0 =aÞ;
(2.64)
gzx ¼ jðmp=aÞcosðmpx=aÞsinðmpx0 =aÞ and
2
k b2mn
gzz ¼ cosðmpx=aÞcosðmpx0 =aÞ:
bmn
We have listed the tangential components of interest only here. These results have
also been derived using Green’s theorem by Elliott [9]. For sidewall slot problems
the required Green’s function components can be obtained easily from (2.62).
References
[1] Harrington R. F. Time Harmonic Electromagnetic Fields. McGraw-Hill,
New York, 1961.
[2] Rumsey V. H. ‘The reaction concept in electromagnetic theory’. Physical
Review. 1954;94(6):1483–1491.
[3] Stutzman W., Thiele GA. Antenna Theory and Design. John Wiley and Sons,
Hoboken, NJ, 2013.
[4] Love A. E. H. ‘The integration of equations of propagation of electric wave’.
Philosophical Transactions of the Royal Society London, Ser A.1901;197:1–45.
16 Slotted waveguide array antennas: theory, analysis and design
The most important technical advance in the 1930s was, in the opinion of the
writer, the invention of the resonant slot. This was a device which was both an
aperture radiator and a resonant structure. Its novelty was major. Nothing as
important had appeared since Hertz invented the dipole and the loop and Lodge and
Bose experimented with open-ended waveguide radiators.
The words are from Ramsay [1]. He mentions in particular Alan D. Blumlein
at EMI Central Research Laboratories in the UK as the inventor of the resonant
slot antenna (patent no. GB 515684, 1938). Blumlein also proposed linear arrays
of slots.
The period 1935–45 has been named the golden years for advances in microwave
technology. The reason was an urgent need to develop a radar for the detection of
military targets, and both the United States and Great Britain were heavily involved
18 Slotted waveguide array antennas: theory, analysis and design
in this effort. Radar development was independently carried out also in France and
Germany in the 1930s, while Japan and Russia entered the field later.
In 1940 the United States and Great Britain started to exchange information
about radar developments. The British brought with them an extremely important
invention, the resonant-cavity magnetron, which could produce power in kilowatts
at centimetre wavelengths [4]. At this time the Radiation Laboratory at MIT (1940–45)
was established for the development of microwave radar and navigational
equipment for military purposes. A major legacy even today is the Radiation
Laboratory Series of 28 volumes published after the war by McGraw-Hill,
describing the results of the research at Rad Lab in areas related to microwaves and
radar. Worth mentioning in the present context are in particular the Volumes 10 and
12 edited by Nathan Marcuvitz (Waveguide Handbook) and Samuel Silver
(Microwave Antenna Theory and Design), respectively [5–8].
It was in Canada that the first successful waveguide slot antennas were developed.
Watson and his research group at McGill University in Montreal demonstrated the
usefulness of several slot configurations in rectangular waveguides: broad wall
longitudinal and transverse slots and inclined edge wall slots [9]. An example of
experimental data is shown in Figure 3.1, illustrating the equivalent slot resistance
versus slot offset for longitudinal slots [6].
Stevenson [10] studied the electromagnetic boundary problem resulting from
matching the (external) slot field, assuming resonance and the field in the wave-
guide. From power conservation he could predict the longitudinal slot conductance
and show how it varied with the slot offset from the waveguide centreline. See also
Chapter 4 in this book. Stegen [11] provided a set of carefully measured slot
characteristics that could be normalised and used in array designs.
Further progress on slot modelling includes the variational solution by
Oliner [12] and the moment method solution by Vu Khac [13] using pulse
expansion functions, and many more. The goal was to predict the frequency
dependence and the variations due to slot offset and other physical parameters
such as waveguide dimensions and wall thickness. This subject will be discussed
in detail in Chapter 5.
For the efficient design of slotted waveguide array antennas one has to com-
bine slot models, mutual coupling among slots and models for the waveguide
feeding system. Such an array design model was developed by Elliott [14] in 1983.
Design methods are treated in Chapters 6–8. An overview of the progress in slotted
waveguide array antennas up to 1999 has been presented by Rengarajan et al. [15].
100
80
60
40
20
10
R 8
Z0 6
2 x1
1
0.8
0.6
0.5
0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
Slot displacement from centre of guide to centre of slot, in.
Figure 3.1 Measured longitudinal slot resistance versus offset from centreline
[6]. From J. W. Dodds, E. W. Guptill and W. H. Watson, National
Research Council of Canada
Stub support
Radiating slots
Bearing
series of eccentric disks, displaced from the centre of the outer slotted conductor.
As the inner conductor rotates, the wavelength in the line changes, causing the
beam to scan.
Other scanners were also developed, for example, the Eagle Scanner, of
which more than 2,000 units were produced before the end of the war. In the
Eagle Scanner radiating dipoles were probe-fed from a rectangular waveguide.
By mechanically changing the width of the waveguide the relative phases
of the dipoles were changed, resembling frequency scanning. 30 of scan
could be realised [17]. Slots in circular waveguides are further discussed in
Section 9.10.
A typical layout of a planar slotted waveguide array antenna, also called a flat
plate antenna, is shown in Figure 3.3. The waveguides act both as a feeding
structure and as structural support. The radiating elements are the openings (slots)
in the front face. The antenna has a flat radiating surface and low thickness. It is
suitable for applications in the range of about 1–100 GHz.
In Figure 3.3 we see eight waveguides with radiating slots forming a rectan-
gular antenna aperture. The cut-out reveals the coupling from a feeding waveguide
with inclined slots in the common wall. With slot dimensions and positions care-
fully optimised excellent radiation performance and very low losses are realised;
more about this in the following chapters. Although the slotted waveguide array
antenna was first used in military radar systems, it is also found in many non-
military applications such as remote sensing, weather radar, navigation, radio links
and several other communication systems. Figure 3.4(a) and (b) shows two slotted
waveguide arrays of very different dimensions.
The SAR antenna shown in Figure 3.4(b) was used on the Earth Resources
Satellite ERS-1 that was launched already in 1991, followed by ERS-2 in 1995.
Several more similar systems have been launched since providing high-resolution
information for land, ocean and atmospheric monitoring [20,21].
There are obviously numerous applications of slotted waveguide array anten-
nas. A single slotted waveguide can act as a linear feed for a cylindrical reflector
antenna. With a non-resonant (series) feed the antenna beam can be scanned by
means of frequency variations. An example of an advanced phase/frequency
History 21
(a)
(b)
Figure 3.4 (a) Slotted array antenna. 2017 Rantec Microwave Systems Inc.
Reproduced from [19]. (b) The 10-m-long Satellite SAR antenna
(C-band) for remote sensing undergoing near-field tests in an
anechoic chamber. Courtesy of RUAG Space AB
scanned planar array is shown in Figure 3.5(a). It is used in a C-band radar system
(Figure 3.5(b)) for pinpointing the location of hostile artillery. As seen in Figure 3.5(a)
numerous vertical slotted waveguides make up the aperture. The antenna beam is
scanned in azimuth by phase steering and in elevation by frequency variation. This is
22 Slotted waveguide array antennas: theory, analysis and design
(a) (b)
Figure 3.5 (a) Phase/frequency scanned slotted waveguide array; (b) radar
unit. Courtesy of Saab AB
(a)
(b)
Figure 3.6 Radio link antennas: (a) antenna with narrow beam for point-to-point
systems; (b) two antennas (E- and H-plane) with sector beam for
point-to-multipoint systems. Courtesy of Ericsson AB
References
[1] Ramsay J. ‘Highlights of antenna history’. IEEE Antennas and Propagation
Newsletter. December 1981. pp. 8–20.
[2] Van Atta L. C. ‘A history of early microwave antenna development’. IEEE
Microwave Theory and Techniques Newsletter. October 1981. pp. 10–14.
[3] Sobol H., Tomiyasu K. ‘Milestones of microwaves’. IEEE Transactions on
Microwave Theory and Techniques. 2002;50(3):594–611.
[4] Skolnik M. I. Introduction to Radar Systems. McGraw-Hill, New York, 1962.
[5] Burns R. W., ed. ‘The background to the development of the cavity
magnetron. In Radar Development to 1945. Peter Peregrinus, London, 1988.
24 Slotted waveguide array antennas: theory, analysis and design
[6] Silver S., ed. Microwave Antenna Theory and Design. MIT Rad. Lab. Series,
Vol. 12, McGraw-Hill, New York, 1949.
[7] Marcuvitz N., ed. Waveguide Handbook. MIT Rad. Lab. Series, Vol. 10,
McGraw-Hill, New York, 1951.
[8] Flock W. L. ‘The radiation laboratory, fifty years later’. IEEE Antennas and
Propagation Magazine. 1991;33(5):43–48.
[9] Watson W. H. ‘Resonant slots’. Journal of the Institution of Electrical
Engineers. 1946;93(4):747–777.
[10] Stevenson A. F. ‘Theory of slots in rectangular waveguides’. Journal of
Applied Physics. 1948;19(1):24–38.
[11] Stegen R. J. ‘Slot radiators and arrays at X-band’. IRE Transactions on
Antennas and Propagation. 1952;1(1):62–84.
[12] Oliner A. A. ‘The impedance properties of narrow radiating slots in the
broad face of rectangular waveguide’. IRE Transactions on Antennas and
Propagation. 1957;5(1):4–20.
[13] Vu Khac T. A Study of Some Slot Discontinuities in Rectangular Wave-
guides. Ph.D. Dissertation, Monash University, Australia. November 1974.
[14] Elliott R. S. ‘An improved design procedure for small arrays of shunt slots’.
IEEE Transactions on Antennas and Propagation. 1983;31(1):48–53.
[15] Rengarajan S. R., Josefsson L. G., Elliott R. S. ‘Waveguide-fed slot antennas
and arrays: a review’. Electromagnetics. 1999;19(1):3–22.
[16] Ridenour L. N. Radar System Engineering. MIT Rad. Lab. Series, Vol. 1,
McGraw-Hill, New York, 1947.
[17] Cady W. M., Karelitz M. B., Turner L. A. Radar Scanners and Radomes.
MIT Rad. Lab. Series, Vol. 26, McGraw-Hill, New York, 1948.
[18] Petersson R., Ingvarson P. ‘The planar array antennas for the European
remote sensing satellite ERS-1’. Proceedings of the European Microwave
Conference. 1988. pp. 289–294.
[19] http://www.rantecantennas.com, retrieved 2017-04-28.
[20] van’t Klooster K. ‘ERS-1, European remote satellite was launched 20 years
ago’. 21st International Crimean Conference on Microwave and Tele-
communication Technology (CriMiCo’2011), September 2011, Sevastopol,
Crimea, Ukraine. pp. 117–118.
[21] Moreira A. ‘A golden age for spaceborne SAR systems’. 20th International
Conference on Microwaves, Radar and Wireless Communication (MIKON).
2014. pp. 1–4.
[22] Svensson B. ‘Dual use of slotted waveguide array antennas’. IEEE Conference
on Antennas and Propagation for Wireless Communication, November 2000.
pp. 149–152.
[23] Svensson B., Manholm L., Wikgren E. ‘A waveguide sector antenna for
point-to-multipoint systems’. IEEE Antennas and Propagation Society
Symposium. 2003. pp. 1185–1188.
Chapter 4
The slot antenna
In this chapter, we will consider the single-slot antenna. We will first discuss an
arbitrarily shaped aperture in an infinite conducting ground plane. Then we spe-
cialise to a rectangular resonant slot antenna and derive its radiation conductance.
We continue with the special but important case of a longitudinal slot in the broad
wall of a rectangular waveguide. The normalised modal functions for describing
fields in waveguides are introduced. The equivalent slot conductance as seen from
the feeding waveguide is then calculated. Mutual coupling to other slots and ground
plane edge effects is also discussed.
z, z'
Θ r
Ea(x, y)
y, y'
r'
x, x' Φ
(a) (b)
Figure 4.1 (a) A general aperture in a ground plane; (b) cross section
Equation (4.5) resembles a Fourier transform, except for the integration limits.
However, let us extend the integration limits to infinity (makes no difference, since
1
We have chosen the most common definition of the potential F from the relation eE ¼ r F .
However, in [3,4] the relation E ¼ r F is used.
The slot antenna 27
the aperture field is zero outside the aperture). Hence, we introduce the following
Fourier transform:
ð1 ð1
f t k x ; ky ¼ E a ðx; yÞejðkx xþky yÞ dxdy (4.6)
1 1
where f t is a vector with the components ( fx, fy). Thus, we can write
Fx e ejkr fy
¼ (4.7)
Fy 2p r fx
From the x- and y-components of F we get the q- and f-components:
FQ ¼ Fx cos f þ Fy sin f cos q
(4.8)
FF ¼ Fx sin f þ Fy cos f
ðð
1
E ðx; y; zÞ ¼ 2 ~ kx ; ky ; z ejðkx xþky yÞ dkx dky
E (4.13)
4p
28 Slotted waveguide array antennas: theory, analysis and design
k 2 ¼ kx 2 þ ky 2 þ kz 2 (4.17)
That is,
8 qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
< kz ¼ þ k 2 kx2 ky2 when kx2 þ ky2 < k 2
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi (4.18)
: k ¼ j k 2 þ k 2 k 2 when k 2 þ k 2 > k 2
z x y x y
ky
Invisible
Visible
kx
k
since f k ¼ 0, that is, the plane wave fields are perpendicular to the direction of
propagation, see also [2, p. 624].
Knowing the aperture fields we can now formulate an expression for the
complex power flow out of the aperture:
ðð ð þ1 ð þ1 2
1 1 k ky2
fx kx ; ky j2 dkx dky
P ¼ Pr þ jPi ¼ Ex Hy dS ¼ 2
S2 8p kh 1 1 kz
(4.22)
Inserting Ex and Hy from (4.20) and (4.21) into the expression for the complex
power integral given earlier leads to a sextuple integral. However, it can be reduced
to this form (4.22, right member) using the relation
ð þ1
ejðk1 k2 Þx dx ¼ 2pdðk1 k2 Þ (4.23)
1
An excellent illustration of the complex power from a narrow slot aperture has
been given by Rhodes [6] (see Figure 4.3). The volume within the visible region
kx2 þ ky2 < k 2 corresponds to the radiated power. The reactive power is outside this
region. Note that the reactive, non-radiating (stored) power goes both negative and
positive (electric and magnetic energy, respectively). When they balance each other
the aperture is resonant.
V0 I0 jV 2 j
P ¼ Pr þ jPi ¼ ¼ 0 Y (4.24)
2 2
We have so far only discussed the characteristics of a general aperture. In
practice an aperture is backed by some feeding arrangement, cavity or waveguide,
30 Slotted waveguide array antennas: theory, analysis and design
ky
(k2 – ky2)|Fx(kx,ky,k)|2 k 2k 3k 4k
4k
2kZ
8 0{(k
2 – kx2 – ky2)½}*
3k
2k
V02k
22Z0
k
0 k kx
Figure 4.3 One quadrant of the complex power density from a thin half wave
slot antenna with applied voltage V0 at the centre. Reproduced from
[6, Fig. 3.1, p. 50]. By permission of Oxford University Press
etc., and the total admittance is modified by these structures. Before going into
more detail on this we will in the next section analyse the conductance of a narrow
rectangular slot in an infinite ground plane.
Consider a rectangular slot in a large ground plane coincident with the x/z-plane as
shown in Figure 4.4. The feeding voltage is V0 in the slot centre. The slot is narrow,
z-directed, with width w and length 2L.
As in (4.24) we can relate the slot conductance Gslot to the radiated real power Pr:
1
ReðPÞ ¼ Pr ¼ V02 Gslot (4.25)
2
We will find the radiated power by integrating the power pattern from the slot.
Since a narrow slot and a thin dipole have identical radiation patterns (if E- and
H-fields are interchanged) we can use known results for the electric dipole, the dual
problem [3].
In the dipole case with a sinusoidal current Iz fed in the dipole centre
Iz ðzÞ ¼ I0 sin½k ðL jzjÞ
the pattern is [4]
θ
r
y
2L
Feed line
w
x Φ
Ground plane
V0
Esx ðzÞ ¼ cos kz (4.28)
w
The equivalent magnetic current source is
M ¼ 2E ^y ; (4.29)
The last expression is obtained directly from the dipole case (4.27), by repla-
cing the electric current I0 by the magnetic current 2V0/w and have the impedance
of free space h replaced by 1/h.
32 Slotted waveguide array antennas: theory, analysis and design
1 1
jSj ¼ jE H j ¼ hjHq j2 (4.32)
2 2
thus
ðp ðp
1
Pr ¼ h dF r2 sin qjHq j2 dq (4.33)
2 0 0
We get
ð p=2 cos2 p cos q
V02 2
Pr ¼ I; where I ¼ dq (4.34)
hp 0 sin q
This last integral is a bit complicated to evaluate. The solution can be for-
mulated in terms of the cosine integral function Ci(x), see, for example, [7, p. 224],
or be solved numerically.
Changing variables in (4.34) we obtain for the integral I
ð 2p
1 1 cos v
I¼ dv (4.35)
4 0 v
1
I ¼ ½0:5772 þ lnð2pÞ Cið2pÞ (4.37)
4
Ci(x) is tabulated in Standard Mathematical Tables [8, p. 532] and we obtain
finally
V02
Pr ¼ 0:609 (4.38)
hp
We have already in (4.25)
1
Pr ¼ V02 Gslot (4.39) = (4.25)
2
0:609 1
From the last two expressions we obtain Gslot ¼ 2
siemens, thus
60p 1;000
about 1,000 W (or 500 W if radiating on both sides of the ground plane).
The slot antenna 33
4.3.1 Definitions
A general expression for the waveguide fields can be written as [1]
X
E ðx; y; zÞ ¼ Ai e ti ezi^z e
gi z
X t
i
(4.41)
H ðx; y; zÞ ¼ Ai h i þ hzi^z e
gi z
i
y
z
4.3.1.4 Normalisation
We choose
ðð
0; i 6¼ j
e i e i dS ¼ dij ¼
t t
(4.43)
S 1; i¼j
where the integration is over S ¼ the waveguide cross section.
Other criteria are sometimes used, for example, in [1,10]:
ðð
t 0; i 6¼ j
e i h j ^z dS ¼
t
S 1; i ¼ j
which differs from our choice by a factor ¼ the modal admittance. See also [11].
Our choice according to (4.43) results in
ðð
t
e ti h j ^z dS ¼ Yi dij (4.44)
S
2
Note, however, that Collin [1] uses nm.
The slot antenna 35
b
x
a
TM case:
2 hmp mpx npy np mpx npy i
e tmn ¼ pffiffiffiffiffi cos sin ^x þ sin cos ^y (4.51)
kc;mn ab a a b b a b
2kc;mn mpx npy
ezmn ¼ pffiffiffiffiffi sin sin (4.52)
gmn ab a b
t 2jwe0 hnp mpx npy mp mpx npy i
h mn ¼ pffiffiffiffiffi sin cos ^x cos sin ^y (4.53)
gmn kc;mn ab b a b a a b
The special case with m ¼ n ¼ 0 is often neglected since it does not contribute
anything to the power flow in the waveguide [12]. It contributes, however, to the
36 Slotted waveguide array antennas: theory, analysis and design
stored (reactive) power and is sometimes essential in the analysis of slots in the
wall of a waveguide [13].
n̂
Waveguide Es
S2
S1
n̂ n̂
z
S3
We assume
E 1 =H 1 ¼ the field produced by the excited aperture
E 2 =H 2 ¼ a testing field ða normalised waveguide modeÞ
and rewrite (4.63) in the following form:
ð
^
n E 1 H 2 ^n E 2 H 1 dS ¼ 0 (4.64)
S
Now
^
n E 1 ¼ 0 on S3 except in the aperture
^
n E 2 ¼ 0 on all of S3 :
Thus
ð ð
^
n E s H 2 dS þ E 1 H 2 E 2 H 1 ð^z ÞdS
aperture S1
ð
þ E 1 H 2 E 2 H 1 ^z dS ¼ 0 (4.65)
S2
Take two modes, i and j with amplitudes Ai and Aj, respectively. Then
ð
E i H j ^z dS ¼ Ai Aj Yi dij (4.66)
cross
section
The result is
ð
1
A
i ¼ ð^n E s Þ h i
e gi z dS (4.68)
2Yi aperture
L L
w
x0
z
3
How narrow is ‘narrow enough’? From experience a possible criterion is w/2L < 0.1 to 0.2. For wider
slots the longitudinal E-field may become significant.
The slot antenna 39
where bi is the propagation constant for mode i in the waveguide and x0 is the slot
position from the side wall. Note that Aþ
i ¼ Ai since the slot field was assumed to
be symmetric.
This result is valid for longitudinal, narrow slots in arbitrary waveguides. As
the next step we will study the special case with rectangular waveguides and cal-
culate the TE10 scattering, see Figure 4.9.
Carrying out the integration in (4.72), neglecting variations across the narrow
slot width, we obtain
pffiffiffi p
V0 jp 2 cos b1;0 L px0
Aþ
1;0 ¼ A1;0 ¼ pffiffiffiffiffi 2L2 cos (4.74)
b1;0 a ab p a
b1;0 2
2L
X0
b
a
x
Inserting the resonant length 2L ¼ l/2 and b1,0 ¼ 2p/lg, where lg is the guide
wavelength, we get finally:
pffiffiffi rffiffiffi
j 2lg a pl px0
A1;0 ¼ V0 cos cos (4.75)
pl b 2lg a
To summarise we have now a relation between the slot voltage V0, the scat-
tered wave amplitude A 1;0 and the geometrical parameters. The simplifying
assumptions leading to (4.75) were
● narrow rectangular longitudinal slot
● slot voltage V0 given, cosinusoidal field distribution
● slot length 2L ¼ l/ 2
● fundamental TE10 mode
● zero wall thickness
For non-zero wall thickness, V0 applies to the inner slot field. The inner and
outer fields are about the same for a practical wall thickness. Nevertheless, the
thickness has an impact on the slot admittance and the slot resonant length as we
will see later.
We still need to find the relation between the incident waveguide mode
amplitude and the slot excitation.
Slot
Ainc
i Ainc
i
Gs Yi
– +
Ai Ai
z=0
Figure 4.10 Circuit model with a shunt conductance Gs representing the slot
loading of the waveguide. Ai represent the scattered wave
amplitudes (voltages) referenced to the slot position z ¼ 0. Yi is the
characteristic admittance of the (waveguide) transmission line
The slot antenna 41
The power radiated by the slot is represented by the power dissipated in the
shunt conductance in our circuit model:
1
Pdis ¼ jVn j2 Gs (4.76)
2
where the mode voltage Vn ¼ Ainc
i þ Ai :
This power must be equal to the radiated power already derived:
V02
Prad ¼ 0:609 (4.77) = (4.38)
hp
The normalised shunt conductance Gs/Yi is related to the reflection coefficient
G ¼ A
i =Ai and we get (assuming resonance)
inc
This important result, originally derived by Stevenson [15], shows how the
conductance (i.e. the excitation) of a slot depends on the slot position. We repro-
duce the Stevenson result in Figure 4.11.
So far we have discussed the slot conductance for resonant slots. The
assumption has been that the slot length l/2 automatically results in slot resonance
42 Slotted waveguide array antennas: theory, analysis and design
0.70
0.60
0.50
0.40
0.30
Normalised resonant conductance, Gr /G0
0.06
0.04
0.03
0.02
0.01
0 0.050 0.100 0.150 0.200 0.250
Slot offset in inches
Figure 4.11 Resonant conductance versus slot offset from the waveguide
centreline for a longitudinal slot in standard X-band waveguide,
calculated by Stevenson [15]. Points are experimental results
according to Stegen [16]. 1981 J. H. Elliott. Reprinted from [4],
with permission
and that the slot admittance is real. However, resonance also depends on the slot
shape, the waveguide dimensions and the effects of mutual coupling among the
slots in an array. We are also interested in the complete slot admittance and how it
varies over a band of frequencies. A more detailed analysis is therefore necessary in
order to control these parameters and optimise the slot (array) antenna for best
performance including bandwidth.
Consequently it is not necessarily true that a slot length of l/2 results in
resonance. From the more detailed slot modelling examples discussed in Chapter 5,
we will find that both shorter and longer slots (than l/2) can be resonant, depending
on several parameters such as slot offset, waveguide height and dielectric loading.
In order to accurately calculate the slot admittance we need to include the reactive
fields in the slot vicinity. For the external region we already have found one field
representation that includes this: the plane wave spectrum [(4.20) and (4.21)]. For
the internal waveguide region we have similarly the ‘spectrum’ of waveguide
The slot antenna 43
modes (4.68). Before going into detailed discussions on how these and other
representations can be used in order to find the true slot field distribution and the
slot admittance we will take a look at the mutual coupling effects.
dB
0
Isolated
–5
Embedded
–10
–100° –50° 0° 50° 100°
θc
(a) (b)
Figure 4.12 (a) A slotted stripline array; (b) measured isolated and embedded
E-plane patterns for an element in the centre of the array.
IEEE 1974. Adapted from [20], with permission
44 Slotted waveguide array antennas: theory, analysis and design
[21]. See also Section 9.1. From the slot spacing the critical ‘blind angle’ qc is
predicted to be at 37 according to the grating lobe angle equation.
sin qc ¼ l=d 1 (4.84)
The H-plane coupling for slots is much weaker than the E-plane coupling, and
can sometimes be disregarded in non-scanned waveguide slotted array designs.
However, our example demonstrates that the array excitation can be dramatically
changed by mutual coupling and that we need to accommodate this effect in a
general design procedure.
A final word on coupling definitions used here: isolated coupling stands for the
case with only two elements present. Array coupling is again coupling between two
elements, but now all other elements are also present and terminated in matched
loads. Active impedance/admittance stands for the situation where all elements are
fed. Isolated (or self) impedance/admittance is for one single element only. Typical
isolated and embedded patterns are compared in Figure 4.12.
Ex
#m #n x
z .
Figure 4.13 Cross section of a slotted array antenna illustrating the E-plane
coupling between apertures m and n
The slot antenna 45
J1
V1 +
– Slot
array
N ports
JN
VN +
–
and obtain
þ
Vn Vn Y0 ¼ Vnþ þ Vn hHnext ðen Þ; en i (4.88)
The part inside the brackets h i is the aperture self-admittance: Ynn ¼
hHnext ðen Þ; en i. The reflection coefficient G becomes
Y0 Ynn
G ¼ Vn =Vnþ ¼ (4.89)
Y0 þ Ynn
An accurate evaluation of the self-admittance is rather complicated
(cf. Section 6.4), partly because we need to find and integrate the field inside the
source region. Furthermore, in order to improve the accuracy of the solution; it is
often necessary to include a number of higher-order modes besides the fundamental
mode in the aperture; more about this in Chapter 5.
appear. Note that Hnext ðem Þ is the external field at aperture n due to the source at
aperture m.
Matching internal and external magnetic fields at aperture n gives
Vn Hnext ðen Þ þ Vmþ þ Vm Hnext ðem Þ ¼ Y0 Vn en (4.91)
Let us also match the magnetic fields at the source aperture m, giving
þ
Vm þ Vm Hmext ðem Þ þ Vn Hmext ðen Þ ¼ Y0 Vmþ Vm em (4.92)
Taking inner products with the modal functions em and en as before yields the
two equations:
)
Vn Ynn þ Vmþ þ Vm Ynm ¼ Y0 Vn
þ (4.93)
Vm þ Vm Ymm þ Vn Ymn ¼ Y0 Vmþ Vm
Note that Ynm ¼ Ymn ¼ hHnext ðem Þ; en i is the mutual admittance between the
two apertures. Eliminating Vm gives the mutual coupling in terms of the scattering
parameter Smn
The coupling between just two elements (no other element present) is called
the isolated coupling. For slot apertures it decays with distance approximately as
1/R in the E-plane. In the H-plane the decay is roughly 1/R2.
M s ¼ 2Ea^z (4.100)
1 1
H ¼ rE ¼ rrF (4.101)
jwm jwme
That is,
ðð
1 ejkR
H ¼ rr Ms dS (4.102)
4pjwm S R
We can expand
r r F ¼ r rF r2 F (4.103)
The slot antenna 49
a
x
z Ground plane
r2 F þ k 2 F ¼ 0 (4.104)
we obtain,
r r F ¼ r rF þ k 2 F (4.105)
@ 2
F has only a z-component, so rðrF Þ ¼ 2 Fz . Therefore,
@z
2
1 @
Hz ¼ þ k 2
Fz (4.106)
jwme @z2
For this case it can be shown that derivation with respect to field coordinates is
the same as derivation with respect to source coordinates. Thus,
ðð
1 @ 2 ejkR 0 0
Hz ðx; zÞ ¼ Ea ðx0 ; z0 Þ k 2 þ 0 2 dx dz (4.107)
2pjwm0 slot @z R
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
where R ¼ ðx x0 Þ2 þ ðz z0 Þ2 .
Neglecting variations in the aperture electric field across the slot (x-direction)
we have
ð a=2
b 0 @ 2 ejkR 0
Hz ðx; zÞ ¼ Ea ðz Þ k þ 0 2
2
dz (4.108)
2pjwm0 a=2 @z R
The electric field variation along the slot is Ea(z0 ) ¼ Ve10(z0 ), where e10(z0 ) is
the normalised modal function:
rffiffiffiffiffi
0 2 pz0
e10 ðz Þ ¼ cos (4.109)
ab a
and V is the slot voltage.
50 Slotted waveguide array antennas: theory, analysis and design
z'
R1 z
x'
R
(x', z')
Field point
#m R2
Source (x, z)
#n
Figure 4.16 The two-slot geometry for calculating the magnetic field in slot #n
due to coupling from the source slot #m. See also Section 6.4
The slot antenna 51
and finally
jb
ð pz
p
ejkR1 ejkR2
Ynm ¼ cos þ
pawm0 #n a a R1 R2
(4.114)
p2 ð 0 jkR
pz e
þ k2 cos dz0 dz
a #m a R
4.5.2.7 Examples
The mutual coupling between two elements, that is, with no other elements present,
is defined as isolated coupling, just two slots and a common ground plane
(Figure 4.17). In the full array, coupling between two slots will be influenced by the
presence of the other slots, array coupling. One effect is that the other slots will
absorb some power, thus reducing the coupling values. See the following examples
showing isolated and array coupling for a linear array. The calculated results have
been obtained with one aperture mode only, which appears to be sufficient in many
cases. The variations across the slot width have not been included. However, a
correction factor sinc2(x) has been employed for the E-plane results [25]. This is
not needed for more narrow slots (Figure 4.18).
52 Slotted waveguide array antennas: theory, analysis and design
–15
Computed
–20 Measured
Coupling dB
–25
–30
–35
1 2 3 4 5 6 7 8 9 10
Element number
–120
–125
Coupling degrees
–130
–135
–140
Computed
–145 Measured
–150
1 2 3 4 5 6 7 8 9 10
Element number
Figure 4.17 The magnitude and phase of the isolated coupling (two elements) for
a 10 element E-plane array. The free space delay (phase) has been
subtracted. Element no. 1 is fed. Aperture size: 0.75l 0.20l
–15
Computed
–20
Measured
Coupling dB
–25
–30
–35
–40
1 2 3 4 5 6 7 8 9 10
Element number
–140
–145 Computed
Coupling degrees
Measured
–150
–155
–160
–165
–170
1 2 3 4 5 6 7 8 9 10
Element number
Figure 4.18 The magnitude and phase of the array coupling (all elements) for the
same array as in Figure 4.17. The free space phase delay has been
subtracted. Element no. 1 is fed
The slot antenna 53
–7.8
–7.85
–7.9
–7.95
–8
S11/dB/
–8.05
–8.1
–8.15
–8.2
–8.25
–8.3
1 2 3 4 5 6 7 8 9 10
Second slot position no.
Figure 4.19 Reflection |S11| in dB of slot no. 1 with a single parasitic slot
at increasing positions 2–10. Dashed line shows also
intermediate positions. Dash-dotted (straight) line: the isolated
reflection. E-plane
We find here that the scattering parameters, that is, elements of the scattering
matrix, do not only depend on the apertures m and n, but also on the surrounding
apertures. Even the self-reflection, that is, the S11 element in the scattering matrix,
is shown in Figure 4.19 to depend on the distance to a second, parasitic slot (only
two apertures present in this case). See also [26]. We have in this situation
Y02 Y11
2
þ Y12
2
S11 ¼ (4.117)
ðY0 þ Y11 Þ2 Y12
2
which becomes the conventional isolated (one element only) scattering if Y12 ? 0.
In terms of scattering parameters we can also write
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
0
S11 1 1 þ S11 0
S11 ¼ þ þ S12 2 (4.118)
2 2
0
where S11 stands for the scattering of the single (isolated) element.
Coupling in an H-plane linear slot array is shown in Figures 4.20 and 4.21.
In the H-plane example the phase agreement (measured vs. calculated) for
array coupling is not as good as in the previous examples (E-plane). It has been
found that higher-order modes are excited in the apertures and should be included if
better agreement is sought. However, the coupled levels are very low so this effort
is probably not needed in practice. As we can see from these results, the H-plane
54 Slotted waveguide array antennas: theory, analysis and design
–20
–30 Computed
Measured
Coupling dB
–40
–50
–60
–70
–80
0 2 4 6 8 10 12 14 16
Element number
50
Computed
Coupling degrees
Measured
–50
0 2 4 6 8 10 12 14 16
Element number
Figure 4.20 The magnitude and phase of the isolated coupling (two elements)
for a 16-element H-plane array. Element no. 1 is fed. Aperture
size: 0.73l 0.30l
–20
–30 Computed
Measured
Coupling dB
–40
–50
–60
–70
–80
0 2 4 6 8 10 12 14 16
Element number
50
Computed
Coupling degrees
Measured
–50
0 2 4 6 8 10 12 14 16
Element number
Figure 4.21 The H-plane array coupling for the same array as in Figure 4.20
The slot antenna 55
Image
Edge
Ground plane
Figure 4.23 Edge reflection effect seen as a contribution from an image slot
coupling is very low for rectangular slots. This also leads to the interesting fact that
isolated and array H-plane coupling (for linear arrays) are at about the same level.
However, in two-dimensional arrays, the ‘H-plane coupling’ can seek other ways
for the interaction. Figure 4.22 shows how elements m and n can interact via several
of the elements in their surroundings.
Edge
Ground plane
m n
Figure 4.24 Set-up for measuring the edge effect on mutual coupling between two
aperture elements m and n. The other apertures were covered with
conducting tape. From [29]. Reproduced by permission of the
Institution of Engineering & Technology
–20
–30
–40
Coupling dB
–50
–60
–70
–80
–90
–100
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
(a) Element number
360
320
280
Coupling degrees
240
200
160
120
80
40
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
(b) Element number
Figure 4.25 Measured (dashed line) and computed (solid line) mutual coupling
including the effect of edge diffraction. The no edge case (ground plane
edges covered with absorbing material) is shown as a dash-dotted line.
(a) is magnitude and (b) is phase. From [29]. Reproduced by
permission of the Institution of Engineering & Technology
The slot antenna 57
References
[1] Collin R. E. Field Theory of Guided Waves. IEEE Press, New York, 2nd ed.,
1991.
[2] Balanis C. A. Antenna Theory, Analysis and Design. Wiley, New York,
1997.
[3] Harrington R. F. Time-Harmonic Electromagnetic Fields. IEEE Press, New
York, 2001.
[4] Elliott R. S. Antenna Theory and Design. Prentice-Hall, Englewood Cliffs,
NJ, 1981.
[5] Clemmow P. C. The Plane Wave Spectrum Representation of Electro-
magnetic Fields. Pergamon Press, Oxford, 1966.
[6] Rhodes D. R. Synthesis of Planar Antenna Sources. Clarendon Press,
Oxford, 1974.
[7] Kraus J. D. Antennas. McGraw-Hill, New York, 2nd ed., 1988.
[8] Selby S. M. (ed.). Standard Mathematical Tables. The Chemical Rubber
Co., Cleveland, OH, 1973.
[9] Booker H. G. ‘Slot aerials and their relation to complementary wire aerials
(Babinet’s principle)’. Journal of the Institution of Electrical Engineers.
Part. IIIA. 1946;93(4):620–626.
[10] Collin R. E. Foundations for Microwave Engineering. McGraw-Hill, New
York, 1966.
[11] Pathak P. H. ‘On the Eigenfunction expansion of electromagnetic dyadic
Green’s functions’. IEEE Transactions on Antennas and Propagation.
1983;31(6):837–846.
[12] Van Bladel J. ‘Contribution of the Y¼constant mode to the modal expansion in
a waveguide’. IEE Proceedings H – Microwaves, Antennas and Propagation.
1981;128(5):247–251.
[13] Khac T. V., Carson C. T. ‘m¼0 n¼0 mode and rectangular-waveguide slot
discontinuity’. Electronics Letters. 1973;9(18):431–432.
[14] Collin R. E., Zucker F. J. Antenna Theory, Part I. McGraw-Hill, New York,
1969.
[15] Stevenson A. F. ‘Theory of slots in rectangular waveguides’. Journal of
Applied Physics. 1948;19(1):24–38.
[16] Stegen R. J. ‘Slot radiators and arrays at X-band’. IRE Transactions on
Antennas and Propagation. 1952;1(1):62–84.
58 Slotted waveguide array antennas: theory, analysis and design
Stegen used the following waveguide and slot dimensions (standard X-band
waveguide):
Slot width 0.0625 inch (1.59 mm)
Waveguide width 0.9 inch (22.86 mm)
Waveguide height 0.4 inch (10.16 mm)
Waveguide wall thickness 0.05 inch (1.27 mm)
Frequency 9.375 GHz
The normalised resonant conductance G/G0 ¼ g(x) as a function of offset (x)
was already introduced in Chapter 4 (Figure 4.11). Additionally, the resonant
length as a function of offset is also needed. The data was presented in curves, and
examples are shown in Figure 5.1(a) and (b).
Figure 5.1(a) h1(L/Lr) and h2(L/Lr) conductance and susceptance vs. slot length
Figure 5.1(b) Lr/l0 resonant length versus offset
Figure 4.11 g(x) normalised conductance versus offset
The curves appear to be quite universal for the given waveguide dimensions.
The data can easily be fitted to polynomials to be used in the array design iterations.
However, new dimensions (half height waveguide, etc.) would require new
measurements.
–0.2 0.4
0.490
0.480
0.050 0.100 0.150 0.200 0.250
(b) Slot displacement off waveguide centreline, x inches
Figure 5.1 (a) Normalised admittance versus normalised slot length. (b) Resonant
slot length versus slot offset from centreline. (a, b) 1981
J. H. Elliott. Reprinted from [7] with permission
The outer slot admittance Ysext could be determined theoretically [9]. The other
parameters (Bsint, nd and B0) were determined from a set of experimental data on
slot length and offset for several frequencies. Polynomials could then be derived
from the experimental values and subsequently be used in array designs (within the
range of the experimental data).
62 Slotted waveguide array antennas: theory, analysis and design
Y G0
Γ
Z Y
jB0 0 j jBsint Ysext
Z=
j 0
1 : nd t
Figure 5.3 Equivalent circuit for a shunt slot in a rectangular waveguide. 1990
Microwave Exhibitions and Publishers Ltd. Reproduced from [8] by
permission of Microwave Exhibitions and Publishers Ltd
Figure 5.4 Test equipment for slot measurements. A wedge absorber and a
waveguide short in the foreground
A test fixture for measuring slot characteristics is shown in Figure 5.4. The
fixture is calibrated using a matched load (wedge absorber) and a moveable short.
Various slot plates with different slot dimensions could be inserted and the offset
carefully monitored with the micrometre shown. The example is from an evaluation
of data for slots in a ridge loaded waveguide where theoretical models are some-
what complex [10].
1 2
3 4
H ext
z Region e
t Ex Hz
H inc
z Region i
H zint
(a) (b)
Figure 5.6 (a) The radiating slot problem. IEEE 1987. Adapted from [16], with
permission. (b) Major slot field components
w
t
X" Z
X'
2L
b
a
Figure 5.7 The slot geometry. 2L is the slot length and w the slot width. URSI
1986. Reprinted from [33], with permission
The four problems in Figure 5.5 are all about a slot interface between two
regions which we call the external (superscript e or ext) and the internal (super-
script i or int); cf. Figure 5.6(a).
where Ex is the unknown transverse electric field in the slot, expanded in basis
functions:
X
N
Ex ðx0 ; z0 Þ ¼ Ep ep ðz0 Þ (5.3)
p¼1
ep(z) and es(z) are the basis and testing functions, respectively. The solution to (5.2)
is in the form of (5.3) with coefficients found in the vector
½ E ¼ ðE1 ; E 2 ; . . . ; E N Þ (5.8)
obtained as
1
½E ¼ ½Y e Y i ½h (5.9)
The results presented in this chapter have been obtained with entire basis and
testing functions:
pp 9
basis function ep ðzÞ ¼ sin ðL þ zÞ >=
2L
(5.10)
sp >
testing function es ðzÞ ¼ sin ðL þ zÞ ;
2L
That is, the Galerkin case [17].
½ E ¼ ½ Re ½ E þ (5.14)
Slot models 67
I Halfspace
Y IIa E+
i E–i
Stub
waveguide
(a)
Y
Stub
waveguide
IIb E+
i E–i
III
Main waveguide
(b)
Y
Stub
waveguide
IIb E+
i
III
Hzinc Main waveguide
(c)
Figure 5.8 Slot in thick waveguide wall. The three partial problems (a)–(c).
URSI 1986. Reprinted from [33], with permission
[Y e] was defined in (5.6), Y0 is the waveguide modal admittance and dps is the
Kronecker delta function. The vector ½hewi has elements
The excitation by a unit incident mode in the main waveguide results in the
formal solution
1 inc
½Eþ ¼ Y iw h (5.19)
where the matrix [Y iw] has elements
[Y i] was defined in (5.6) and [hinc] in (5.7). To this solution, (5.19), we will of course
have to add those waves resulting from the E waves incident on the lower interface.
It remains to relate the waveguide modes at the upper interface to those at the
lower interface:
and similarly for the downward propagating modes. [B] is a diagonal matrix with
the elements
Bii ¼ expðjbi tÞ (5.22)
where bi is the propagation constant of the TEi0 mode and t is the wall thickness.
Combining everything together we obtain the total tangential electric field at the
lower interface
1 iw 1 inc
½E ¼ f1 þ ½B½Re ½Bg Ri ½B½Re ½B 1 : Y h (5.23)
This is the final solution for the field in the slot aperture (lower interface).
1
Green’s function ¼ ‘field of a point source’ [32]. Georg Green was a British mathematician who in
1828 published ‘An Essay on the Application of Mathematical Analysis to the Theories of Electricity
and Magnetism’ [18].
Slot models 69
In our case we need the Hz field (TE mode) when there is an Ex field exciting the
slot. Stevenson’s expression for this case is
ð 2
1 0 @
Hz ðPÞ ¼
int
E x ðP Þ þ k G2 ðP; P 0 ÞdS 0
2
(5.24)
jwm0 S @z0 2
Complete expressions including other field components are found in, for
example, [19]. See also Section 2.8.
In (5.24) S 0 denotes the slot region, P and P 0 refer to points in the slot, prime
indicates source points, unprimed field points. The chosen Green’s function is
1 X
X 1 0
egmn jzz j
G2 ðP; P0 Þ ¼ ymn ðx0 ; y0 Þymn ðx; yÞ (5.25)
m¼0 n¼0
2gmn
02
egjz zj ¼ g2 2gdðz0 zÞ egjz zj (5.29)
@z
With this we get
2 ð
1 X 1 X 1
k þ gmn2 0 gmn jzz0 j 0
Hzint ðPÞ ¼ em en Fm ðxÞ E2 ðz Þe dz 2E2 ðzÞ
4jwm0 m¼0 n¼0 gmn S
(5.30)
70 Slotted waveguide array antennas: theory, analysis and design
where
ð w=2
mpx mpðx0 þ x0 Þ 0
Fm ðxÞ ¼ cos E1 ðx0 Þcos dx (5.31)
a w=2 a
Here x0 is the slot offset from the sidewall. With no variations across the slot
we have
mpx mpx0 mpw
Fm ðxÞ ¼ w cos cos sinc (5.32)
a a 2a
Finally, with the expansion of the electric field variation along the slot; cf. (5.10):
X
N
pp
E2 ðzÞ ¼ Ep sin ðL þ zÞ (5.33)
p¼1
2L
1 X 1 X 1
Hzint ðx; zÞ ¼ em en Fm ðxÞ ...
jwm0 ab m¼0 n¼0
8
X >
< pp
N
gmn 2 þ k 2
Ep pp2 egmn ðLþzÞ egmn ðLzÞ
p¼1
>
:4Lgmn g 2 þ
mn
2L
pp2 9
k2 pp >
=
þ 2L ðL þ zÞ (5.34)
pp2 sin
2L >
;
gmn 2 þ
2L
The upper sign in (5.34) is valid for p odd, the lower for p even.
We insert this in (5.35), interchange the order of integration and evaluate the
derivative with the following result:
ð ð
1 1 1 k 2 kz2
Hzext ðx; zÞ ¼ gx ðkx ; kz Þejðkx xþkz zÞ dkx dkz (5.37)
2pwm0 1 1 ky
where gx (representing the radiation pattern):
ðð
1 0 0
gx ðkx ; kz Þ ¼ Ex ðx0 ; z0 Þejðkx x þkz z Þ dx0 dz0 (5.38)
2p slot
With the expansion of the electric slot field as before (5.33) with constant field
across the slot we get finally
X ð ð
p odd
NP
Ep wp 1 1 kz2 k 2 sincðkx w=2Þ cos kz L
Hz ðx; zÞ ¼
ext pp2 j sin kz L
wm0 4pL 1 1
p¼1 ky kz2 p even
2L
ejðkx xþkz zÞ dkx dkz (5.39)
lw2 X 1 X 1
em en
Ypsi ¼ Wm . . .
pab m¼0 n¼0 4
8
> jpp ps
< gmn2 þ k 2 2
2L 2L
2 ps2 ðe2gmn L 1Þ
>
:gmn g 2 þ pp
mn gmn þ
2
2L 2L
pp2 9
k2 >
=
2L
þ 2j pp2 L dps (5.41)
>
;
gmn2 þ
2L
The function Wm is related to the weighting across the slot. With constant field
we have
mpx mpw
0
Wm ¼ cos2 sinc2 (5.42)
a 2a
where x0 is the slot offset from the side wall as before.
72 Slotted waveguide array antennas: theory, analysis and design
In (5.41) the upper sign is valid for both p and s odd, the lower sign for both p and s
even. When p is odd and s even or vice versa Ypsi ¼ 0. dps ¼ 1 for p ¼ s, 0 otherwise.
0.70
Forward scattering
Back scattering
0.65
a = 22.86 mm
w = 1.59 mm
0.60 t =0
2*L/Lambda
f = 9.375 GHz
Half height
0.55
Full height
0.50
0.45
0 1 2 3 4 5 6 7
Offset from centreline (mm)
Figure 5.9 Computed resonant length based on the backscattered and forward
scattered waves. Two waveguide heights are compared. Zero wall
thickness. IEEE 1987. Reprinted from [16], with permission
Based on the forward scattering (S12) instead, still assuming a shunt repre-
sentation, we would get
Y =G0 ¼ 2ð1 S12 Þ=S12 (5.48)
The two expressions give the same admittance for a pure shunt element since
S12 ¼ 1 þ S11 in that case. Once we have found the electric field distribution in the
slot we can compute the backward and forward scattered wave amplitudes, for
example, from (4.70), and from them calculate the assumed ‘shunt’ admittance
values. Resonance is then defined from the admittance being pure real. An example
is shown in Figure 5.9.
From this example it is clear that the scattering off the slot is not perfectly
symmetrical. It has been found [11] that the electric field variation along the slot
has an asymmetrical (odd) component as well as a symmetrical (even) component.
The odd component gives rise to asymmetrical scattering. Examples of this have
been given by Stern and Elliott [15] and others.
In Figure 5.9 the resonant length based on backscattering and on forward
scattering is shown versus the slot displacement from the waveguide centreline. We
see that the two results differ significantly when the offset is increased. The dif-
ference is even more pronounced for slots in reduced height waveguide. For slotted
arrays with very stringent requirements it may be necessary to represent the slot by
a more complex network (Tee or Pi network) replacing the simple shunt model. See
also Compound Slots in Chapter 8.
74 Slotted waveguide array antennas: theory, analysis and design
1.0
b = 10.16 mm
Ex (z) b = 2.54 mm
0.5
a = 22.86 mm
w = 1.59 mm
t=0
offset = 5 mm
f = 9.375 GHz
0.0
0.0 0.5 1.0
Normalised position along slot
Figure 5.10 Slot electric field distribution for two waveguide heights. IEEE
1987. Reprinted from [16], with permission
0.50
a = 22.86 mm offset = 3 mm
b = 10.16 mm f = 9.375 GHz
w = 1.59 mm
t = 1.27 mm
2*L/Lambda
0.48
t=0
0.46
0 2 4 6 8 10
Number of basis functions
The computed electric field distribution along a slot, Ex(z0 ), is shown in Figure 5.10
for two cases. The solid line is for a slot in full-height (standard) X-band guide, the
broken line is for a slot in quarter-height guide. The latter case is characterised by a
large asymmetrical component, indicating that the simple shunt model is a poor
representation of the scattering from the slot.
For most of the results presented here nine or ten expansion functions for the
slot field were used. The same number was also used in the stub waveguide to
represent cases with finite wall thickness. Of course fewer terms can be used when
less accurate results are acceptable to save computer time. Figure 5.11 shows for
two wall thicknesses how the computed resonant length depends on the number of
expansion functions chosen.
Slot models 75
We can see that the zero wall thickness case (t ¼ 0) converges more slowly
than the case with finite wall thickness. It is also observed that the higher-order
even terms in the expansion contribute very little to the computed result. The even
terms represent the asymmetrical part of the slot field (which, however, becomes
more important in reduced height waveguide; Figure 5.10). In general, the electric
field distribution is slightly bowed out compared to the first sinusoidal term,
pffiffiffiffiffiffiffiffiffiffi
resembling the shape of sin q.
The difference between the cases with t ¼ 0 and t ¼ 1.27 mm with respect to
convergence can be understood when the edge condition at the slot ends is con-
sidered [26]. With zero wall thickness the electric field near the slot ends is
expected to behave as r1/2, r being the distance from the edge. This results in an
expected decay rate for the coefficients Ep / p3/2. With finite thickness the right-
angle edge gives the dependence r2/3 and an expected decay rate Ep / p5/3. Thus,
in the latter case, a slightly faster convergence is expected. See Figure 5.12, where
the expansion coefficients are plotted on a logarithmic scale for two wall thick-
nesses. Slot data are the same as in Figure 5.11. pffiffiffiffiffiffiffiffiffiffi
As discussed earlier one might consider adding a term with sin q shape in
the expansion of the slot field (5.10). This term would have the proper variation at
the edges r1/2 and might reduce the number of terms needed for a good field
solution. A similar term would serve in the thick walled case [27, p. 99].
The entire basis results agree quite well with calculations using pulse expan-
sion and point matching as demonstrated in Figure 5.13. It should be noted, how-
ever, that in the pulse expansion case one extra pulse of length 0.02 l0 was added
somewhat arbitrarily to allow for the expected electric field variation down to zero
amplitude at the slot ends. Including this, the agreement between the two solutions
is very good, with a maximum difference of about 0.003 l0.
1
t=0
t = 1.27 mm
0
Log (Ep)
–1
Odd
–2
Even
–3
0.0 0.5 1.0
Log ( p)
0.51 [11]
This theory
a = 22.86 mm
0.47
b = 10.16 mm
t=0 w = 1.59 mm
f = 9.375 GHz
0.45
0 5 10
Offset from centreline (mm)
Figure 5.13 Computed resonant length versus slot offset. Entire basis (solid line)
versus pulse basis (triangles). URSI 1986. Modified from [33],
with permission
In our expansion of the slot field we have assumed the electric field to be
constant across the slot width. In fact, the expected physical behaviour [26] is a
singularity such that the field Ex varies as 1/r1/2, where r is the distance to the edge
(zero thickness case). A reasonable function modelling this behaviour is the elec-
trostatic field variation in a narrow gap:
w=p
E1 ðx0 Þ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi; jx0 j < w=2 (5.49)
ðw=2Þ2 x0 2
One could also add an odd component (and possibly even higher terms) with
the same edge behaviour [28]:
2x0 =p
E1 ðx0 Þ ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi; jx0 j < w=2 (5.50)
ðw=2Þ2 x0 2
Similar edge mode functions have been proposed for the dual case: scattering
from a strip conductor [29]. For slotted waveguide array antennas with narrow slots
it does not seem to improve the results very much to include these edge functions.
However, for transverse slot problems using the Moment Method with pulse
expansion it has been found that an edge mode may be needed in order to obtain
convergence [30].
Re(Hz)
Im(Hz)
1.5
Im(Hzext – Hzint )
Re(Hzext – Hzint )
1.0
Im(Hzinc )
0.5 Re(Hzinc )
z
–L –5 0 5 L (mm)
–1.0
–1.5
|Hzext – Hzint|
10
Relative field strength
0 5 10 15
L
Longitudinal position along and beyond slot, mm
Figure 5.15 Magnetic field along and beyond the slot [14]
Slot models 79
Now it might seem that we have not verified the boundary equation, from
which we started:
0.51 Measured
Computed
0.49
2*L/Lambda
0.47 a = 22.86 mm
b = 10.16 mm
w = 1.59 mm
t = 0.38 mm
0.45
0 5 10
Offset from centreline (mm)
Figure 5.16 Computed and measured resonant lengths for X-band rectangular
slots in standard waveguide. IEEE 1987. Reprinted from [16],
with permission
80 Slotted waveguide array antennas: theory, analysis and design
0.03
0.01
0.00
–0.01
0 1 2 3 4 5
Offset from centreline (mm)
Figure 5.17 Difference in resonant length for two wall thicknesses, t ¼ 0.3 mm
and t ¼ 1.0 mm, versus slot offset from waveguide centreline.
Rounded slot ends. IEEE 1987. Reprinted from [16],
with permission
0.4
0.3
Delta factor
0.2
0.1
0.0
0.0 0.2 0.4 0.6
Normalised slot offset
Figure 5.18 Rounded end correction factor from measured data versus slot offset
normalised to half waveguide width. URSI 1986. Reprinted from
[33], with permission
Here L0 and L& are half the resonant slot lengths for rounded and rectangular slot
ends, respectively, w is the slot width and D is a correction factor. It has been
suggested that slots with equal area, alternatively slots with equal perimeter, should
have equal resonant frequency. Yee [5] proposed an average between the two,
which leads to D ¼ 0.32. More recently the correction factor was determined
empirically from measured data for several cases with rounded slot ends and the-
oretical results for the corresponding rectangular slots, see Figure 5.18. The mea-
surements were made at C- and X-band, and for wall thicknesses from 0.3 to
1.8 mm both in normal and half-height waveguide. Theoretical calculations using
Slot models 81
2*L/Lambda
0.49
a = 22.86 mm
b = 10.16 mm
w = 1.59 mm
f = 9.375 GHz
0.45
0 5 10
Offset from centreline (mm)
Figure 5.19 Computed resonant length versus slot offset for standard X-band
guide compared with measured data from [35]. IEEE 1987.
Reprinted from [16], with permission
0.52 Measured
Computed
0.50
2*L/Lambda
0.48
0.46
0 3 6
Offset from centreline (mm)
Figure 5.20 Computed and measured resonant length versus slot offset for
C-band slot with rounded ends in half-height waveguide. a ¼ 39 mm,
b ¼ 10 mm, w ¼ 2.8 mm, t ¼ 0.6 mm, f ¼ 5.3 GHz. URSI 1986.
Reprinted from [33], with permission
MoM/FEM analysis have also been done [34]. Both studies support the assumption
that slots with equal area have approximately equal resonant frequency, that is,
D ¼ 1 p/4 ¼ 0.215. Assuming equal perimeter would lead to twice this correction.
Using this rounded end correction we obtain the results shown in Figure 5.19,
where a comparison is made with Stegen’s measured data [35]. Another example is
given in Figure 5.20.
82 Slotted waveguide array antennas: theory, analysis and design
1.0
Conductance G/Y0
0.5
Computed
Measured
0.0
0.45 0.50 0.55 0.60
(a) Slot length in wavelengths
0.5
Susceptance B/Y0
0.0
Computed
Measured
–0.5
0.45 0.50 0.55 0.60
(b) Slot length in wavelengths
Figure 5.21 (a) Computed and measured slot conductance versus slot length for
X-band slot in half-height waveguide. IEEE 1987. Reprinted from
[16], with permission. (b) Computed and measured slot susceptance
versus slot length for X-band slot in half-height waveguide. URSI
1986. Reprinted from [33], with permission
Figure 5.21 shows the computed slot conductance and susceptance versus slot
length compared with measurements. The waveguide wall thickness was 0.3 mm.
The slot had rounded ends and we applied the correction factor to account for this.
b
x
a
X
N
pp
Ez ¼ Ep sin ðx x0 þ LÞ (5.55)
p¼1
2L
with no variation across the slot width, w (see Figure 5.23). The variable x0 is the
off-centre displacement, and L is the half length of the slot. Nine terms (N ¼ 9)
were used for the numerical results presented here. The Ex field along the slot is
neglected as well as the magnetic field Hz across the slot.
The external magnetic field and the associated Moment Method matrix ele-
ments will be the same as in the longitudinal slot case (Section 5.2). The internal
matrix elements, however, are different for the transverse slot.
L L
x0
The appropriate interior Green function for the waveguide region is well
known [2,19]. Limiting ourselves to the Hx component from the slot electric Ez
field, we obtain both TE and TM contributions. They are summed to yield
pp mp2 mpx
XN XXem en L k sin
2
a a
Hx ðx; zÞ ¼ Ep ...
pp 2 mp 2
p¼1 m n jwm abg2
0 nm
2L a
0 1p odd
mpx0 mpL
sin cos
B a a C
B C 1 e gmn w=2
cosh g z (5.56)
@ mpx0 mpL A mn
cos sin
a a p even
Here
mp2 np2
g2mn ¼ þ k2 (5.57)
a b
and em ¼ 1 for m ¼ 0, while em ¼ 2 for m > 0.
For the Galerkin case we form inner products with the testing functions, ws
ps
ws ¼ sin ðx x0 þ LÞ (5.58)
2L
to yield the matrix element contributions Ypsint :
pp ps mp2 1 egmn w
XX em en k 2
w
L L a gmn
Ypsint ¼ mp2 . . .
pp 2 mp 2 sp 2
m n jwm abg2
0 nm
2L a 2L a
0 1p odd 0 1s odd
mpx0 mpL mpx0 mpL
sin cos sin cos
B a a C B a a C
B
@ mpx0 mpL
C
A B
@ mpx0
C
mpL A (5.59)
cos sin cos sin
a a p even a a s even
equivalent slot impedance. The reflection coefficient inside the waveguide in terms
of the propagating TE10 mode becomes
0 1p odd
pp px0 pL
sin cos
Hxscat sinðb10 w=2Þ X
N Ep B a a C
pp2 p2 B C
G¼ ¼ pffiffiffiffiffi 2L (5.62)
Hxinc wm0 ab p¼1 @ px 0 pL A
cos sin
2L a a a p even
5.4.3 Results
Figure 5.24 shows the computed normalised resistance and reactance versus slot
length compared with previous results by Khac [11], who used a pulse expansion
with point matching. These results are for zero wall thickness.
Figure 5.25 shows computed results compared with measured results for a
transverse slot in a reduced-height waveguide. The measured slot had rounded
ends. The length was therefore reduced by D ¼ 0.215w to represent an equivalent
rectangular slot; cf. Figure 5.18.
The effect of the waveguide height is demonstrated in Figure 5.26. The resis-
tance is approximately inversely proportional to the waveguide height. Figure 5.27
1.6 0.8
Resistance R/Z0
Reactance X/Z0
0.8 0.0
0.0 –0.8
0.40 0.46 0.52
Slot length in wavelengths
Figure 5.24 Transverse slot impedance in zero wall standard X-band waveguide
versus normalised slot length 2L/l. Rings and crosses denote
computed results (present method). Lines are from [11]. IEEE
1993. Reprinted from [43], with permission
Slot models 87
1.25
4.8 GHz
Reactance X/Z0
0.00
5.8 GHz
–1.25
0.0 2.5
Resistance R/Z0
Figure 5.25 Transverse slot impedance for a 28 mm long slot with zero offset from
4.8 to 5.8 GHz. a ¼ 39 mm, b ¼ 10 mm, w ¼ 3 mm, t ¼ 1.9 mm. Solid
curve with crosses is computed, rings denote measured values.
IEEE 1993. Modified from [43], with permission
2.5
10 mm
2.0
Resistance R/Z0
1.5
20 mm
1.0
0.5
0.0
4.8 5.3 5.8
Frequency (GHz)
Figure 5.26 Computed slot resistance for two waveguide heights b ¼ 10 mm and
b ¼ 20 mm, other parameters as in Figure 5.25. Curve with rings is
measured results for b ¼10 mm. IEEE 1993. Reprinted from [43],
with permission
88 Slotted waveguide array antennas: theory, analysis and design
1.5
0 mm
1.0
Resistance R/Z0
5 mm
0.5
0.0
4.8 5.3 5.8
Frequency (GHz)
Figure 5.27 Computed slot resistance for two offset values, x0 ¼ 0 mm and 5 mm.
b ¼ 20 mm, other data as in Figure 5.26. IEEE 1993. Reprinted
from [43], with permission
shows the dependence on the offset. Only minor changes can be realised using this
parameter.
There is no simple way of varying the excitation of a transverse slot. There is
also no phase reversal technique available, such as alternating the offset direction
or changing the sign of the inclination angle as in other types of slots. In arrays, this
will lead to the generation of grating lobes if resonant slot spacing shall be used,
that is, one guide wavelength for broadside radiation. Parallel plates in the form of
baffles can be used to suppress the grating lobes in the E-plane of an array of
transverse slots; more about this in Section 9.3.
References
[23] Eshrah I. A., Yakovlev A. B., Kishk A. A., Glisson A. W., Hanson G. W.
‘The TE00 waveguide mode – The ‘‘complete’’ story’. IEEE Antennas and
Propagation Magazine. 2004;46(5):33–41.
[24] Clemmow P. C. The Plane Wave Spectrum Representation of Electro-
magnetic Fields. Pergamon Press, Oxford, 1966.
[25] Stinson D. C. Intermediate Mathematics of Electromagnetics. Prentice-Hall,
Englewood Cliffs, NJ, 1976.
[26] Meixner J. ‘The behavior of electromagnetic fields at edges’. IEEE Trans-
actions on Antennas and Propagation. 1972;20(4):442–446.
[27] Amitay N., Galindo V., Vu C. P. Theory and Analysis of Phased Array
Antennas. Wiley Interscience, New York, 1972.
[28] Butler C. M., Wilton D. R. ‘General analysis of narrow strips and slots’.
IEEE Transactions on Antennas and Propagation. 1980;28(1):42–48.
[29] Richmond, J. H. ‘On the edge mode in the theory of TM scattering by a strip
or strip grating’. IEEE Transactions on Antennas and Propagation. 1980;
28(6):883–887.
[30] Park K. P., Stern G. J., Elliott R. S. ‘An improved technique for the eva-
luation of transverse slot discontinuities in rectangular waveguide’. IEEE
Transactions on Antennas and Propagation. 1983;31(1):148–154.
[31] Rengarajan S. R., Josefsson L., Petersson R. ‘Recent developments in broad
wall slots in rectangular waveguides for array applications’. Seventh Inter-
national Conference on Antennas and Propagation, ICAP 91 (IEE). 1991,
pp. 729–732.
[32] Harrington R. F. Time-Harmonic Elctromagnetic Fields. IEEE Press,
New York, 2001.
[33] Josefsson L. ‘Analysis of longitudinal slots in rectangular waveguides
including the effect of wall thickness’. Proceedings of the URSI Electro-
magnetic Theory Symposium. Budapest, Hungary. August 1986, pp. 367–369.
[34] Zhang M., Hirano T., Hirokawa J., Ando M. ‘Method of Moments analysis
of a waveguide round-ended wide slot by using numerical-eigenmode
basis functions’. IEEE Topical Conference on Wireless Communication
Technology. 2003, pp. 360–361.
[35] Stegen R. J. Longitudinal Shunt Slot Characteristics. Hughes Tech. Memo
No. 261, November 1951.
[36] Derneryd, A. G., Lorentzon T. C. ‘Design of a phase/frequency scanned
array antenna with non-resonant slotted ridge waveguide elements’.
Proceedings of the IEEE Antennas and Propagation Symposium, London,
Ontario, June 1991, pp. 1728–1731.
[37] Karlsson I. ‘Applications of waveguide arrays in commercial and military
radars’. Proceedings of the 1993 Antenna Applications Symposium, Allerton
Park, IL, USA, pp. 1–22, 1993.
[38] Marcuvitz N. (ed.) Waveguide Handbook. MIT Rad.Lab.Series, Vol. 10,
1951.
[39] Rengarajan S. R., Derneryd A. ‘Improved scattering model for weak broad-
wall slots’. Microwave and Optical Technology Letters. 1993;6(8):504–507.
Slot models 91
Based on the analysis of the single-slot antenna in the previous chapters we can
now approach the problem of designing linear and planar slot arrays. We will first
study longitudinal slot arrays and start with the simple case with one row of slots in
a rectangular waveguide. For this case mutual coupling appears mainly in the
H-plane and can in many cases be neglected. Some examples of computed and
measured performance will be presented. Procedures for designing linear arrays of
slots, including some examples, will be discussed.
6.1 Introduction
At the design frequency the slots are resonant so that their admittances are real.
With a uniform amplitude distribution (i.e. with equal slot conductances ¼ G0/N) a
narrow broadside beam is obtained in the H-plane with first sidelobes about
13 dB below the main beam peak. With a tapered distribution the slot con-
ductances shall be chosen according to the desired slot voltage distribution; the slot
conductance is assumed to be proportional to the square of the slot voltage as an
approximation. The relationship between the slot conductance and slot voltage will
be further discussed in Section 6.3.
94 Slotted waveguide array antennas: theory, analysis and design
φ1 φ2 φN
y1 y2 y3 yN
Figure 6.1 A linear slot array and its network representation with shunt
admittances
–10
Reflection coefficient in dB
–20
–30
N=2 N=4
N=8 N = 16
–40
–10 dB ref
–50
0 4 8 12 16
Per cent bandwidth
Figure 6.2 Reflection coefficient bandwidth of a standing wave array for different
numbers N of slot elements (10 dB ref is a baseline reference)
self-admittance, that is, that of an isolated slot, and geometrical parameters such as
offsets and lengths of all slots in the array. This expression includes external mutual
coupling between all slots. Internal TE20 mode coupling between adjacent slots is
derived later. The TE20 mode coupling between the slot closest to the short circuit
and the short may be included using a similar approach. The design procedure uses
an iterative technique as discussed in Sections 6.3, 6.5 and 6.6. Convergence is
reached generally after a few iterations.
The design approach using the incremental conductance, discussed in Sec-
tion 6.7, is especially useful in relatively large arrays in geometries where it is
difficult to determine the self and mutual admittances. The incremental con-
ductance is defined as the difference between the total input conductance of two
arrays, one having N resonant slots and the other N þ 1. In both cases the slots are
identical (same values of offset, length or tilt) with adjacent offsets or tilts alter-
nating positive and negative values. Incremental conductance data is computed or
measured for a range of values of slot offsets or tilts. It is then possible to design
arrays that have slowly varying aperture distributions. It is possible to use this
approximate technique as a starting design, which may be improved with more
rigorous design techniques, if necessary.
The Elliott design technique models longitudinal radiating slots in the form of
shunt admittances in a transmission line. For arrays having slots with large offsets
in reduced height waveguides and compound slots characterised by offsets and tilts,
a better slot model will be in the form of a Tee or Pi network. For such slots a
design procedure using scattering parameters of the slot is discussed in Section 6.8.
The Elliott design procedure requires scattering characteristics of isolated
longitudinal radiating slots as a function of slot offsets and lengths. Ideally a slot in
a waveguide should be modelled as a four-port network, with two ports on the
waveguide on either side of the slot and two ports for the two apertures of a slot
when wall thickness is included in the model. For practical values of wall thickness,
the voltages of both slot apertures are nearly the same. Therefore, the wall thickness
is ignored in the Elliott design procedure except when determining the scattering
data of isolated slots. It should be mentioned that the resonant length of a slot is
sensitive to the wall thickness. A two-port model of a slot with both ports in the
waveguide, with the knowledge of the slot voltage is adequate. The slot aperture
distribution is assumed to be a half cosine distribution with a complex coefficient,
and it is related to the type of excitation such as the incident TE10 mode, higher-
order modes from adjacent slots or external radiation from all other slots. With the
availability of many commercial computational electromagnetic (CEM) codes it is
possible to compute slot data very accurately. The mutual coupling may be com-
puted accurately as discussed in Section 6.4.
The design of slot arrays involves the determination of waveguide dimensions
and slot parameters to achieve the required radiation characteristics, for example,
gain, sidelobe level and cross-polarisation level, and the return loss as a function of
frequency. On the other hand, in the analysis process one is interested in deter-
mining the radiation characteristics and the return loss, given the waveguide and slot
parameters. Design techniques generally use transmission line models with lumped
The linear slotted waveguide array antenna 97
circuit loads representing the slots. While some analysis models also use similar
loaded transmission line models, rigorous analyses employing finite element meth-
ods, integral equations/method of moments (MoM) and mode matching techniques
provide very accurate solutions. The analysis procedure employing the MoM solu-
tion of the integral equations for the slot aperture fields is discussed in Chapter 8.
0
With MC
Pattern level in dB –10 Without MC
–20
–30
–40
–50
–60
–90 –60 –30 0 30 60 90
Theta in degrees
4.21 showed that the coupling level in the H-plane beyond the nearest slot was less
than 40 dB. However, for stringent pattern shapes, for example, low sidelobes,
the coupling in the H-plane may have to be included, cf. [14]. In the E-plane the
coupling is much stronger and should be accounted for. Therefore the mutual
coupling has typically to be included in the design of planar arrays, where both
E-plane and H-plane coupling is present, see Chapter 7.
An illustration of the coupling effect on an eight-element equal amplitude
H-plane linear slot array is shown in Figure 6.3. The array was designed first
without including the mutual coupling effects, and in a subsequent design mutual
coupling was included. A moment method analysis of both designs yielded the
patterns shown in Figure 6.3. The MoM uses nine entire domain sinusoidal basis
functions, similar to the ones discussed in Chapter 5, for each slot aperture. In
addition, the MoM solution accounts for external mutual coupling and internal
higher-order mode coupling. A further discussion on MoM for slot arrays is found
in Chapter 8. As seen in Figure 6.3, the mutual coupling effect on the radiation
pattern is very small and of no importance in this case. The example shown is for a
half-height X-band (9.3 GHz) waveguide with slot widths 1.6 mm and a wall
thickness of 0.5 mm and sum of conductances G0 ¼ 1.0. The principles for
including the mutual coupling in the theoretical design are elaborated in
Section 6.3.
Let us look at an example with a tapered excitation for lower sidelobes where a
greater impact of the mutual coupling could be expected. A result for a 30 dB
Dolph–Chebyshev pattern is shown in Figure 6.4. Here the mutual coupling effect
is noticeable in the first sidelobe. Although the effect of the error is to reduce the
first sidelobe level in this particular example, it may have an undesirable effect in
other examples by increasing the sidelobe levels. Further details of the design of
The linear slotted waveguide array antenna 99
–10
Pattern level in dB MC
–20
No MC
–30
–40
–50
–60
–90 –60 –30 0 30 60 90
Theta in degrees
this array are found in Section 6.5. The sidelobe taper is due to the slot element
pattern. Slight asymmetry in the pattern is found in the array designed without
mutual coupling. This asymmetry is caused by the internal TE20 mode coupling
generated at the slot closest to the short, which couples to the same slot after getting
reflected by the short. The internal coupling will be discussed in Section 6.4.
The asymmetry is not found for the array designed using the mutual coupling
effects since the above-mentioned TE20 mode coupling was taken into account in
the design. Therefore the slot voltages and the pattern are close to the ideal for
this design.
Slot # i
A+i A
Yi Ys Vn G0
Ai– B
(a) (b)
The reflection coefficient must be the same in the two representations, thus
G ¼ A þ
i =Ai ¼ B=A (6.4)
1 2 1
jA j Yi ¼ jBj2 G0 (6.7)
2 i 2
That is,
pffiffiffiffiffiffiffiffiffiffiffiffi
B ¼ Yi =G0 A i (6.8)
This gives
rffiffiffiffiffiffi
Ys 2 Yi
¼ A (6.9)
G0 Vn G0 i
We have already an expression for the scattered mode amplitude A
i in (4.74).
Thus we obtain
Ys Vs
¼ K1 f ðx0 ; LÞ (6.10)
G0 Vn
The linear slotted waveguide array antenna 101
This expression relates the slot admittance Ys, the slot voltage Vs and the
modal voltage Vn in the first design equation of Elliott [11, Eq. (10)]. Equation
(6.10) is valid for the isolated slot admittance as well as the active slot admittance.
In the expression for the isolated slot admittance the slot voltage does not include
mutual coupling whereas for the active admittance, mutual coupling is accounted
for. An explicit expression for the mutual coupling is derived in the second design
equation.
The constant
rffiffiffiffiffiffi pffiffiffi
Yi jp 2
K1 ¼ 2 pffiffiffiffiffi ; (6.11)
G0 b10 k0 a ab
is the same as that in [11, Eq. (11)], while
p
cos b10 L px0
2kL
f ðx0 ; LÞ ¼ 2 cos (6.12)
p 2 b10 a
2kL k
Equation (6.10) differs from Elliott’s equation (10) in [11] by a minus sign
because of sign differences in the normalisations for the TE10 mode in the two deri-
vations. The sign difference does not cause any problem in the design of slot arrays
since we always work with the ratio of active admittances to realise the aperture
distribution, for example, (6.16). We also observe that the term (k/k0) in the numerator
inside the square root of Elliott’s equation (11) in [11] for K1 should not be present so
that the corrected expression will be valid for dielectric filled waveguides also.
Coupling
Slot
A+
Waveguide
A–
Figure 6.6 Slot excited internally from the waveguide and externally from the
mutual coupling
102 Slotted waveguide array antennas: theory, analysis and design
The term A is proportional to the slot voltage Vs, which likewise can be
divided into two components: Vs ¼ Vsisol þ Vscoupl.
We can write
G a =G ¼ A
isol þ Acoupl =Aisol ¼ Vs =Vsisol ¼ Vs = Vs Vscoupl ¼ 1= 1 Vscoupl =Vs
(6.14)
Ya 2Ga 2
¼ ¼
G0 1 þ G a 1 þ 1=G a
Ya 2
¼
G0 2 1 Vscoupl (6.15)
þ
Ys =G0 G Vs
which essentially is the second design equation according to Elliott [11, Eq. (33)].
Yna fn Vns VN
¼ (6.16)
YNa fN VNs Vn
The linear slotted waveguide array antenna 103
The subscript n refers to the slot number n while N refers to the Nth slot, the one
closest to the short used as a reference. Note that fN and fn are defined in (6.12).
Vns and VNs are slot voltages while Vn and VN are mode voltages. For standing wave
arrays with a slot spacing of half guide wavelength, the mode voltage magnitudes
are the same and their signs alternate and so do the signs of fn s. For a travelling
wave array, expressions for the mode voltages and the admittances are described in
Section 6.6. For the Nth slot, that is, n ¼ N, (6.16) is not relevant. Enforcing the
active admittance of the Nth slot to be real yields one equation. One more equation
is obtained by enforcing the total active admittance at the input port to the char-
acteristic admittance of the TE10 mode, G0 so as to achieve a match at the input
port. Thus there are 2N non-linear equations in 2N unknowns for the array design.
The simultaneous non-linear equations may be solved iteratively by a quasi-
Newton technique [16]. Convergence is reached when the changes in the values of
slot offsets and lengths between two successive iterations are less than the desired
manufacturing tolerance, D, or if the desired performance is met. Figure 6.7 shows
the flow chart for the design algorithm.
Start
Choose N,
a, b, f, w, t
Read slot data
Iter = 1
Solve 2N
non-linear
Iter = Iter+1 equations
Iter >1
and
No ∆ < tol ?
Yes
Stop
1 Ys =G0 X N
0 s
Vscoupl ¼ Vs ext ¼ jðb10 =k Þðk0 bÞða=lÞ3 V gmn ðLm ; Ln ; Xmn ; Zmn Þ
fn Ys =G0 þ 2 m¼1 m
2
(6.17)
where Lm and Ln are the half lengths of slots m and n, respectively, and Xmn and Zmn
are the spacings between slots m and n along the E-plane and H-plane directions,
respectively [11, Eq. (29)]. The prime in the summation means that the m ¼ n term
is excluded. The integral gmn cast as a dimensionless quantity in terms of normal-
ised lengths by Elliott [11] is related to the mutual admittance between apertures in
a ground plane, Ymn, in (4.115), repeated next.
jb
Ymn ¼ gmn (6.18)
pah
For the convenience of calculation of Ymn it has been assumed that Lm ¼ Ln. The
use of (6.17) in (6.15) will yield the second equation given by Elliott [11, Eq. (33)].
It may be seen that the mutual admittance derived in Chapter 4, (4.85) and (4.86) treat
the case of slots on a ground plane. The presence of the waveguide under the slot n
in (6.17) is responsible for the additional terms in (6.17) [11, Eqs. (29) and (33)].
The linear slotted waveguide array antenna 105
We consider the mutual coupling integral gmn specified in [11] and defined in
(4.114) and (4.115) in Chapter 4. Equation (4.114) is used here to express gmn in
terms of slot length 2L and width w. Since slot lengths in arrays vary by a small
amount, one may calculate gmn for slots m and n by keeping their lengths equal by
averaging the lengths of slot m and n. There are two terms in gmn shown next, one
containing a single integral and the other a double integral. We will simplify gmn2 to
a single integral so that gmn can be computed rapidly by numerically evaluating the
single integrals.
p L
ð pz expðjkR Þ expðjkR Þ
1 2
gmn1 ¼ cos þ dz (6.20a)
2k0 L L 2L R1 R2
and
p 2 1 ð L pz ð L 0
pz expðjkRÞ 0
gmn2 ¼ k02 cos cos dz dz
2L k0 L 2L L 2L R
(6.20b)
where
h i
2 1=2
R ¼ Xc2 þ ðZc þ z0 zÞ ; (6.21a)
h i1=2
R1 ¼ Xc2 þ ðZc þ L zÞ2 (6.21b)
and
h i1=2
R2 ¼ Xc2 þ ðZc L zÞ2 : (6.21c)
Figure 6.8 shows the geometry of the coupling problem. Two sub-domains of
the double integral in gmn2, evaluated in terms of difference coordinate t ¼ z0 z,
are shown in Figure 6.9.
Z'
Z
X'
w Zc
X 2L
Xe
z'
τ = 2L
L
τ=0
2
–L z
L
1
–L τ = –2L
Figure 6.9 Two sub-domains of the coupling integral; t < 0, and t > 0
Aperture
number: 1 2 n N
V+ V–
Figure 6.10 Waveguide array cross section with slot apertures and feeding
waveguides
(alone, and in an array) and their dependence on their separation. An expression for
the calculation of the mutual admittance Ymn was given by (4.114).
The scattering matrix can be computed once we know the mutual admittance
matrix [Y e] between the slot apertures. However, to fill [Y e] we also need the self-
e
admittance Ynn of a single slot-aperture before we can make the calculation:
Note that in Compton’s expression b is the total width of the waveguide (¼ slot
length) and a is the total height of the waveguide (¼ slot width), just the opposite of
the definitions used in this book. It is easily shown that this solution is identical to
our (5.43) with p ¼ s ¼ 1 inserted.
Compton transforms his solution into the following form, which can be eval-
uated numerically more easily.
ð ða pffiffiffiffiffiffiffiffiffi
ejk x þy
2 2
16 jb b
Y self
¼ g ðxÞhðyÞ pffiffiffiffiffiffiffiffiffiffiffiffiffiffi dxdy (6.27)
awm0 p y¼0 x¼0 x2 þ y2
The parameters are
8
> p
< g ðxÞ ¼ ða xÞ
2 (6.28)
> py py
: hðyÞ ¼ D1 ðb yÞcos þ D2 sin
b b
108 Slotted waveguide array antennas: theory, analysis and design
where
8
>
> 1 k2 p
> D ¼
< 1 b2 4p 4b2
(6.29)
>
> 1 k2 p
>
: D2 ¼ þ
pb 4p 4b2
In the graphs by Compton the aperture admittance values are normalised to
the wave admittance of the medium outside the aperture. Lewin’s graphs are
normalised to the mode admittance of the feeding waveguide. In Figure 6.11 the
1.2
b/λ = 0.6
b/λ = 0.4
0.8
Conductance
0.6
b/λ = 0.2
0.4
0
0.5 0.7 0.9 1.1 1.3 1.5
(a) a/λ
0.6
0.5
b/λ = 0.2
b/λ = 0.4
0.4
Susceptance
b/λ = 0.1
0.3
b/λ = 0.6
0.2
0.1
0
0.5 0.7 0.9 1.1 1.3 1.5
(b) a/λ
Figure 6.11 Aperture conductance (a) and susceptance (b) versus slot length a/l
and width b/l. The aperture field is assumed to be that of a TE10
mode. Values are normalised to the TE10 modal admittance
The linear slotted waveguide array antenna 109
× 10–3
2.5
2
Conductance
1.5
Admittance
1
0.5
0
Susceptance
–0.5
–1
0.4 0.5 0.6 0.7 0.8 0.9 1
a/λ
Figure 6.12 Calculated aperture admittance (Siemens) for a square aperture fed
from a same size square waveguide, thus a ¼ b. Dashed lines are
measured. Reprinted from [24]
calculated admittance for some typical apertures sizes is plotted. (The definitions
of aperture dimensions a/l (length) and b/l (width) are the ones chosen in this
book.)
This relatively simple theory assuming just the fundamental TE10 mode gives
quite good agreement with more complex models and measured results [22,23].
One further example is shown in Figure 6.12, where the admittance is given in
Siemens (i.e. not normalised).
using the waveguide Green’s function (2.61). This can also be obtained from (4.72)
for the TE20 mode of the rectangular waveguide.
110 Slotted waveguide array antennas: theory, analysis and design
The TE20 mode fields propagating in the z direction in the waveguide are
given next.
~ wm0
E b ¼ ^y Ba sinð2p x0 =aÞexpðg20 z0 Þ (6.31)
2p=a
~ b ¼ Ba ^z j cosð2p x0 =aÞ ^x jg20 sinð2p x0 =aÞ expðg20 z0 Þ
H (6.32)
2p=a
The TE20 mode back scattering coefficient of slot n is
ð2p=aÞ2
Ba ¼ Vns hn (6.33)
g20 k0 h0 abk
where
p coshðg20 Ln Þ
hn ¼ 2 cosð2pdn =aÞ (6.34)
2kLn 2 p 2
ðg20 =k Þ þ
2kLn
and k ¼ k0 if the waveguide is air-filled. The TE20 mode forward scattering coef-
ficient is determined as Ca ¼ Ba.
s
The TE20 mode induced voltage, Vn;20 in the nth slot is obtained by comparing
s
the TE10 mode induced voltage Vn10 as follows. Equation (7) in [11], combined
with (6.5), yields (6.35)
Gn n
s
Vn10 ¼ A (6.35)
Kfn 10
where fn is defined as f (x0, L) in (6.12).
s
Similarly Vn20 , the voltage induced by the TE20 mode wave of amplitude An20 is
given by
s
Vn10 I20
s
Vn20 ¼ (6.36)
I10
where
ðð
pzn
I10 ¼ Hz10 cos ds0 (6.37)
slot n 2Ln
and
ðð
pzn
I20 ¼ Hz20 cos ds0 (6.38)
slot n 2Ln
Hz10 is the magnetic field of the TE10 mode wave and Hz20 is that of the TE20 mode
wave in the region of slot n. Combining (6.35) through (6.38) we obtain
The internal coupled voltage is added to the external coupled voltage, (6.17),
thus
Vscoupl ¼ Vs ext þ Vns int (6.40)
V ¼ S Vþ (6.41)
V ot ¼ ðI þ S Þ V þ (6.42)
where I is the identity matrix. Now, suppose we want to have a desired excitation
V des. We should then feed with
The matrix (I þ S)1 provides the desired correction [27,28]. It could in theory
be realised as a network (Figure 6.13).
This corrective matrix would restore the excitation of the slot radiators to the
desired values in the presence of mutual coupling. It would work even for different
tapers and even phase steering of the array, etc. In practice it can hardly be realised
in analogue form. However, for receive only as in digital beam forming it could be
done [27].
112 Slotted waveguide array antennas: theory, analysis and design
Possible corrective
Feed network Radiators
network
The self-admittance is determined from S11 of the two port scattering matrix of
an isolated radiating slot as shown next.
Y 2S11
¼ (6.45)
G0 1 þ S11
Thus the design data needed are (i) resonant lengths, 2Lres, of longitudinal
radiating slots computed at the design frequency for a range of values of slot off-
sets, |d| up to a value typically less than 0.35a. In addition, for each value of |d| we
need computed values of (ii) |S11| as a function of the normalised slot length,
L ¼ L=Lres , in the range of 0.9–1.10 and (iii) the phase of S11 for the same range of
offsets and normalised slot lengths.
In the iterative design, the required S11 for any offset and length is determined
from a two-dimensional interpolation of |S11| and arg Sl1 from the computed (or
measured) data as a function of offset and normalised length. The self-admittance
of the slot may then be determined from (6.45). Alternatively the slot admittance
may be plotted in terms of Stegen normalisation as shown in Figure 5.1(a).
The conductance and susceptance curves may be fitted into polynomials for rapid
calculations for any given length and offset.
V s1 V s2 V s3 V s4 V s5 V s6 V s7 V s8
0.262 0.519 0.812 1.0 1.0 0.812 0.519 0.262
114 Slotted waveguide array antennas: theory, analysis and design
L ¼ L=Lres 0.94 0.955 0.97 0.985 1.0 1.015 1.03 1.045 1.06
jdj
0.444 0.0016 0.0019 0.0022 0.0024 0.0024 0.0023 0.0022 0.0020 0.0018
0.889 0.0065 0.0076 0.0086 0.0093 0.0096 0.0093 0.0086 0.0078 0.0071
1.333 0.0146 0.0169 0.0190 0.0206 0.0211 0.0204 0.0190 0.0173 0.0157
1.778 0.0255 0.0294 0.0331 0.0358 0.0366 0.0355 0.0332 0.0304 0.0276
2.222 0.0392 0.0449 0.0503 0.0542 0.0540 0.0538 0.0505 0.0464 0.0424
2.667 0.0557 0.0634 0.0705 0.0754 0.0768 0.0747 0.0702 0.0648 0.0594
3.111 0.0740 0.0837 0.0924 0.0984 0.1000 0.0975 0.0921 0.0854 0.0787
3.556 0.0941 0.1057 0.1159 0.1227 0.1245 0.1216 0.1153 0.1075 0.0995
4.000 0.1156 0.1288 0.1402 0.1476 0.1495 0.1461 0.1391 0.1304 0.1213
L ¼ L=Lres 0.94 0.955 0.97 0.985 1.0 1.015 1.03 1.045 1.06
jdj
0.444 228.7 219.5 207.9 194.4 180.0 166.5 155.1 146.0 138.9
0.889 228.4 219.2 207.7 194.3 180.1 166.7 155.4 146.3 139.2
1.333 227.6 218.4 207.0 193.7 179.8 166.7 155.6 146.6 139.6
1.778 227.2 218.0 206.8 193.9 180.3 167.5 156.5 147.6 140.6
2.222 226.2 217.1 206.1 193.5 180.3 167.9 157.2 148.5 141.5
2.667 224.9 215.8 204.8 192.5 179.8 167.9 157.6 149.1 142.3
3.111 223.7 214.7 204.0 192.1 179.9 168.5 158.5 150.3 143.6
3.556 222.4 213.4 203.0 191.6 180.0 169.0 159.4 151.4 144.8
4.000 220.8 212.0 201.9 190.9 179.8 169.4 160.2 152.5 146.0
Dolph–Chebyshev array shown in Table 6.1 whereas the design without mutual
coupling is found to have small amounts of error, especially near the edges. There
is a slight asymmetry in the excitations obtained from the analyses because
the structure is not strictly symmetric with a short at one end of the waveguide.
The linear slotted waveguide array antenna 115
Table 6.5 Offsets and lengths in mm of slots of the 30 dB sidelobe Dolph–
Chebyshev array, with and without mutual coupling (MC) (slot
excitation phase is in degrees)
The higher-order modes generated by the slot closest to the short get reflected
by the short and couples to the slot. The asymmetry is slightly greater for the design
without mutual coupling.
Figure 6.4 in Section 6.2 shows the computed H-plane pattern produced by the
two designs at 9.3 GHz. The sidelobe levels decrease away from broadside because
of the element pattern of the slot. There is a small difference between the two
designs at some of the sidelobes. Table 6.6 shows the pattern characteristics at the
centre frequency and at the band edges, where the return loss is 10 dB for the end-
fed Dolph–Chebyshev array designed with mutual coupling. At the higher end of
the band, the directivity is almost the same as that at 9.3 GHz. The increase in the
electrical size of the aperture at 9.50 GHz is offset by the phase errors in the
aperture distribution due to the design at a different frequency (9.3 GHz).
The centre-fed standing wave arrays do not exhibit beam squint with frequency.
However, there is a small amount of beam squint with frequency in the end-fed
array, with positive values corresponding to the squint towards the short at higher
frequencies while negative values of squint are found at lower frequencies. At the
design frequency we have a broadside beam. At other frequencies, the aperture
phase distribution deviates from being equal, thereby creating a squint. The travel-
ling wave arrays to be discussed in Section 6.6 exhibit significant beam squint as a
116 Slotted waveguide array antennas: theory, analysis and design
function of frequency. The beam squint effect has been eliminated in a recent work
by placing a meta-material, exhibiting negative values of permittivity and perme-
ability in the frequency range of interest, between slots in the waveguide [29].
where q0 is measured from the z-axis, pointed along the incident TE10 wave propa-
gation direction. As an example, if b/k0 ¼ 0.707, q0 ¼ 45 . The element spacing
should be greater than the typical slot length so that adjacent slots do not overlap in
their z-coorodinates or introduce substantial amount of higher-order mode coupling
but small enough to avoid grating lobes. For a travelling wave array the element
spacing is chosen such that the reflected waves from all slots very nearly cancel which
results in good input match.
For a travelling wave array consisting of slots with alternating offsets, the
beam pointing angle is given by
0.9
0.85
d in wavelength
0.8
0.75
0.7
83 84 85 86 87 88
(a) Beam pointing angle in degrees
0.7
d in wavelength
0.6
0.5
0.4
93 98 103 108 113 118 123
(b) Beam pointing angle in degrees
Figure 6.15 Element spacing in terms of wavelength for the required beam
pointing: (a) beam towards the load; (b) beam towards the feed port
118 Slotted waveguide array antennas: theory, analysis and design
wavelength are shown for the assumed value of b ¼ 0.707 k0. For q0 < 83 ,
d becomes large enough to allow grating lobes whereas for large values of q0 the
smaller spacing might create severe higher-order mode coupling effects.
Equation (6.50) is the first design equation for the travelling wave array, similar to
(6.16) for the standing wave array. The second design equation used for standing
wave arrays, (6.15) with coupled voltages specified by (6.17) and (6.39) is valid for
the travelling wave arrays also.
determine the mode voltage of slot 2 using (6.48) and the offset and length of slot 2
are determined using (6.15), (6.17) and (6.39) so as to satisfy (6.50). The mutual
coupling term is computed assuming that slots 3 to N have their desired excitations
and are centred in the waveguide and that their lengths are equal to half wavelength
in free space. These computations will be more accurate in subsequent iterations
when the parameters get updated. This process continues until we reach slot N
when we can determine VN. The total normalised input admittance, YNa =G0 , is
determined using (6.49) recursively. The input reflection coefficient and incident
wave power are now calculated. This process is iterated a few times starting from
slot 1 again, until convergence is reached. Between iterations, d1 is adjusted to have
a good compromise between the load power and the input match condition.
A greater value of d1 may reduce the load power but increase the input reflection
coefficient. As a rule of thumb, the number of slots and the spacing may be chosen
such that the scattered wave voltage phasors from all the slots span the complex
plane an integer number of times as suggested by Elliott [2]. However, since the
scattering from different slots have different amounts; this is not a strict rule.
Element # 1 2 3 4 5 6 7
Amplitude 0.334 0.279 0.378 0.485 0.595 0.701 0.800
Element # 8 9 10 11 12 13 14
Amplitude 0.883 0.947 0.986 1.000 0.986 0.947 0.883
Element # 15 16 17 18 19 20 21
Amplitude 0.800 0.701 0.595 0.485 0.378 0.279 0.334
120 Slotted waveguide array antennas: theory, analysis and design
Table 6.8 Offsets and lengths of the radiating slots, designed by including TE20
mode coupling
Element # 1 2 3 4 5 6 7
Offset (mm) 3.000 1.134 1.714 2.225 2.736 2.690 2.413
Length (mm) 15.425 15.368 15.347 15.416 15.442 15.391 15.378
Element # 8 9 10 11 12 13 14
Offset (mm) 2.400 2.521 2.380 2.022 1.812 1.765 1.657
Length (mm) 15.420 15.434 15.383 15.349 15.367 15.378 15.337
Element # 15 16 17 18 19 20 21
Offset (mm) 1.381 1.128 0.989 0.873 0.682 0.366 0.829
Length (mm) 15.296 15.306 15.320 15.288 15.240 15.275 15.093
The measured reflection coefficient was low over 5% frequency band. However,
measured patterns exhibited sidelobe levels in the range of 22 to 25 dB. Elliott
attributed part of the discrepancy between theory and experiment to the internal
coupling between slots via the TE20 mode that was not accounted for in the design
[2]. In addition, manufacturing errors and inaccuracy in the slot data used in the
design procedure and errors in measurement could also account for the discrepancy
between theory and experiment.
In order to test the accuracy of the design procedure for the travelling wave array
and to understand the higher-order mode coupling effects, we investigated the same
21-element array using accurate slot data generated by the MoM solution to the integral
equation of the slot aperture electric field, similar to those shown in Tables 6.2–6.4 in
Section 6.4 but calculated at 9.375 GHz. In addition, an accurate MoM analysis pro-
gram was used to assess the accuracy of the design procedure and various phenomena
that cause discrepancy between the design and analysis. This MoM analysis program
had been verified against very carefully measured experimental data and also with the
commercial code HFSS (www.ansys.com) which is used as an industry standard [31].
We designed two slot arrays. In the first design the internal higher-order mode
coupling via the TE20 mode was included. The second one did not include the
internal TE20 mode coupling to mimic the process that was used in [2]. The values
of the slot spacing (0.389 lg or 17.4 mm) and the first slot offset were chosen to be
the same as in [2].
Using (6.46) we obtain the beam direction measured from the z-axis to be 44.4 .
Table 6.7 shows the relative values of slot voltages, that is, the aperture distribution,
of the array to produce the 30 dB Dolph–Chebyshev pattern. The offsets and
lengths of slots obtained by the accurate design that included the TE20 mode cou-
pling are given in Table 6.8. The slot lengths in Table 6.8 assume square ended slots
used in the MoM computations while those in [2] refer to round ended slots used in
the measured slot data. In addition, our design included internal TE20 mode cou-
pling while Elliott did not account for internal higher-order mode coupling.
The linear slotted waveguide array antenna 121
0
Dsn incl 20
–10 Dsn excl 20
Pattern level in dB
Ideal
–20
–30
–40
–50
0 30 60 90 120 150 180
Angle from the waveguide axis
Figure 6.16 Computed far field pattern of the 21-element travelling wave array.
Dotted line – design with TE20 mode; solid line – design without TE20
mode, both analysed using full wave MoM including all higher-order
mode coupling; ideal – pattern of a 21 element collinear slot array
with ideal aperture distribution
Figure 6.16 shows two radiation patterns produced by the full wave MoM
analysis for the two array designs mentioned above. The MoM solution employs
nine entire domain sinusoidal basis functions for each slot aperture (cf. Section 5.2)
and is discussed in detail in Chapter 8 for the case of planar arrays. The pattern with
the legend ‘Dsn incl TE20’ refers to the array designed with the TE20 mode cou-
pling while ‘Dsn excl TE20’ refers to the array design that did not account for the
internal TE20 mode coupling. Both MoM solutions account for external coupling
between slots and accurately model the self-term of each slot as well as its exci-
tation by the TE10 mode in the waveguide. Clearly the design that includes the TE20
mode coupling is closer to the ideal patterns and demonstrates the need for
including the TE20 mode coupling between adjacent slots. The directivities of the
three patterns had small differences and therefore in order to compare the patterns,
all pattern peaks were normalised to 0 dB.
Figure 6.17 shows two radiation patterns produced by the MoM for the array
that was designed including the TE20 mode coupling. The pattern with the legend
‘TE20 MoM’ was analysed with TE20 mode only for the internal coupling. The
other, with the legend ‘full wave MoM’, was analysed with all higher-order modes
for internal coupling. The two patterns had slightly different directivities. In order
to compare the patterns, the peaks were normalised to 0 dB in both cases. Small
deviation between the two patterns shows the contribution of higher-order modes
other than TE20 that have not been accounted for in the design procedure. Among
the contribution from all higher-order modes for this array, TE20 accounts for about
a third while TE01, TE11 and TE21 modes account for about half. Modes up to m ¼ 4
and n ¼ 2 captured most of all higher-order mode coupling effects. Most standing
wave array designs work well with TE20 mode alone accounted for and occasion-
ally one needs a few other modes such as TE01, TE11, and TE30. Sidelobes in the
angular region q < 45 are substantially lower since the element pattern approaches
122 Slotted waveguide array antennas: theory, analysis and design
–30
–40
–50
0 30 60 90 120 150 180
Angle from the waveguide axis in degrees
Figure 6.17 Computed far field patterns of the 21-element travelling wave array
designed with TE20 mode coupling. Dotted line – full wave MoM
analysis including coupling via all modes; solid line – TE20 mode
MoM analysis including internal coupling between adjacent slots via
TE20 mode only
a null as q ? 0, whereas sidelobes in the angular range q > 45 keep increasing in
levels up to 90 and then keep going down because of the element pattern.
Figure 6.18(a) shows the analysed aperture amplitude distributions using the
full wave MoM, and TE20 MoM of the array designed by including the TE20 mode
internal coupling. Both distributions are compared to the ideal 30 dB Dolph–
Chebyshev distribution. The ideal aperture distribution is symmetric, whereas
the other two exhibit slight asymmetry because of imperfections in the design. The
maximum amplitude is found at different elements for the three cases. All aperture
distributions have been normalised to a peak value of unity. We notice that all
amplitude distributions are nearly the same with some deviations between the ideal
and the other two distributions. The effect of higher-order modes other than the
TE20 mode is found to be very small.
Figure 6.18(b) compares the computed values of the aperture phase distribu-
tion using MoM with the ideal linearly progressive phase of the TE10 mode wave
subtracted out. The difference between the phase data shown in Figure 6.18(b) of
the full wave MoM and TE20 MoM is found to be as high as 17 for some slots. The
variation from the ideal linear phase is as high as 66 over a phase range of about
3,000 . It appears difficult to realise the ideal linear phase distribution because of
the presence of a reflected TE10 mode wave. Figure 6.19 exhibits the input
reflection coefficient, and the load power relative to the incident wave power, both
in dB as a function of frequency. These values were also computed by the full wave
moment method.
Approximate values of the beam pointing angle determined by (6.46) have
been used to find the beam squint as a function of frequency plotted in Figure 6.20.
The beamwidth is about 7.8 so this array can be used over 5% bandwidth without
significant loss of gain. While an acceptable design of a travelling wave array can
be accomplished using the procedure described here, further improvements can be
The linear slotted waveguide array antenna 123
1.00
0.60
Full wave
0.40 TE20 MoM
Ideal
0.20
1 6 11 16 21
(a) Element number
40
30
Phase error in degrees
20
10
0 Full wave
–20
1 6 11 16 21
(b) Element number
Figure 6.18 Computed values of the aperture amplitude distribution and phase
error relative to the ideal aperture phase: (a) amplitude distribution
computed by two analysis methods; (b) phase error relative to the
ideal aperture phase computed by two analysis methods. Squares
correspond to the full wave MoM analysis, and diamonds correspond
to the TE20 MoM analysis
made using the global optimisation technique with a full wave MoM code, dis-
cussed in Chapter 8. Such an optimisation technique has been used to reduce
the power dissipated in the load substantially in the traveling wave array [32].
Mode matching technique and generalised scattering matrix method discussed in
Chapter 7 can also be used for accurate design and optimisations.
Elliott’s equations were cast in a form for the analysis of a slot array by
Hamadallah [33]. The external mutual coupling was incorporated in Hamadallah’s
analysis [33] but the internal higher-order mode coupling was not. Reasonably
124 Slotted waveguide array antennas: theory, analysis and design
–5
Load power
–35
9.1 9.2 9.3 9.4 9.5 9.6
Frequency in GHz
Figure 6.19 Reflection coefficient and load power relative to incident power in dB
2
Beam squint in degrees
–1
–2
–3
8.9 9.1 9.3 9.5 9.7 9.9
Frequency in GHz
good agreement was obtained between measured and computed patterns at the
design frequency as shown in Figure 6.21.
Hamadallah’s analysis equations are repeated next by including the TE20 mode
coupling.
The slot voltages of an N-element linear array are calculated from the fol-
lowing matrix equation:
k2
½ G ½V s ¼ V1 ½ E (6.51)
k1
where
k expðg20 d Þ
Gmn ¼ gmn þ hn fdm n1 hn1 þ dm nþ1 hnþ1 g if m n (6.52)
g20 ðk0 bÞða=lÞ3
The linear slotted waveguide array antenna 125
00
Computed
–10
Measured [2]
–20
dB
–30
–40
–50
00 30 60 90 120 150 180
Angle (deg.)
Figure 6.21 Computed and measured patterns of the 21-element travelling wave
array. IEEE 1989, reprinted from [33]
and, if m > n,
k expðg20 d Þ
Gmn ¼ gmn þ hn fdm n1 hn1 þ dm nþ1 hnþ1 g
g20 ðk0 bÞða=lÞ3
dm n1 and dm nþ1 are Kronecker delta functions (¼1 if the two subscripts are equal,
otherwise 0).
[V s] is the column matrix containing the N slot voltages.
The nth element in the column matrix [E] is given by
sin½b10 fd0 þ ðn 1Þgdg
En ¼ fn (6.54)
sinðb10 d0 Þ
for a standing wave array. The distance from the short to the centre of the closest
slot is d0.
For a travelling wave array with a match termination
En ¼ exp½ jb10 ðn 1Þd fn (6.55)
2kG0
k2 ¼ (6.55a)
jb10 ðk0 bÞða=lÞ3
and
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 2G0
k1 ¼ (6.55b)
jða=lÞ hðb10 =k ÞðkaÞðkbÞ
Edge slots
voltage magnitudes are equal with alternating signs for half wavelength spacing;
cf. (6.16).
Yna Y1a j fn j Vns
= ¼ (6.58)
G0 G0 j f1 j V1s
Equation (6.58) serves as the design equation for a linear array. One needs to
obtain the values of active admittances of slots so as to realise the required aperture
distribution and an input match. The aperture distribution may be complex but the
amount of phase variation achievable with rotated slots is limited. This method may
be used for an initial design of an array and subsequently one can improve the
design using a full wave analysis code and an optimisation procedure. It has been
shown that for linear uniformly excited arrays this method provides good accuracy
whereas for nonuniform excitations better methods are needed; see e.g. [37].
ds
D10,N–1 D10,N
A10,N–1 A10,N
B10,N–1 B10,N
C10,N–1 C10,N
Figure 6.23 Incident, transmitted and scattered TE10 mode waves of slots
later. The scattering parameters of the isolated slot N are determined from
computed slot data for the assumed values of offset and length.
2. We then obtain the contribution Vs N coup due to external mutual coupling from
all other slots assuming that all other slots are centred in the waveguide with
their voltages specified by the aperture distribution and lengths equal to half
wavelength in free space. These values will be updated later in the iterative
process. The isolated and active reflection coefficients of slot N are related as
shown in (6.14) which is repeated below.
GaN =GN ¼ VsN =ðVsN VsNcoupl Þ (6.59)
The slot voltage due to mutual coupling, VsNcoupl , may be obtained for com-
pound slots in a manner similar to the derivation for longitudinal slots.
3. Steps 1 and 2 are repeated until slot N is resonant in an active sense, i.e. in the
presence of mutual coupling. We then determine the active two-port scattering
matrix ½SNa . We set A10;N þ C10;N to 1 as a reference.
4. If there is a match termination at the end, D10,N ¼ 0 whereas if there is a short,
D10;N ¼ ½A10;N þ C10;N expð2b10 ds Þ (6.60)
5. By converting ½SNa to a transmission matrix, we can determine the TE10 wave
amplitudes and phases at the input side of slot N, A10,N and B10,N þ D10,N. The
slot voltage VsN is also determined.
6. We now determine the parameters of slot N 1 such that its voltage VsN 1
relative to VsN satisfies the required aperture distribution after including the
mutual coupling. Its scattering parameters in the absence of mutual coupling
are obtained from computed slot data.
7. Use the procedure stated in step 2 for slot N 1 to calculate, VsN 1 , and ½SN 1a .
8. The TE10 mode waves incident and leaving the output side of slot N 1 are
obtained from the corresponding wave amplitudes and phases at the input side
130 Slotted waveguide array antennas: theory, analysis and design
Elliott’s design procedure uses a shunt admittance model for the longitudinal
radiating slots. For slots in reduced height waveguides and for slots with large
offsets, the shunt model has been found to introduce errors. A Tee network for each
radiating slot may be derived from the computed scattering parameters of an iso-
lated slot. Such a Tee network consists of a dominant term which is the shunt
admittance and two series reactances as shown in Figure 6.24.
The series and shunt elements of the Tee network are readily obtained from the
two-port scattering matrix parameters of the slot. The ABCD matrix of the Tee
network is obtained by multiplying the ABCD matrices of the series and shunt
elements in Figure 6.24 as given in (6.61) [42]. By comparing the elements of the
ABCD matrix of the Tee network to the scattering parameters of the slot, we obtain
the values of the network elements given in (6.62) [42].
A B 1 Z 1 0 1 Z 1 þ ZY 2Z þ Z 2 Y
¼ ¼ (6.61)
C D 0 1 Y 1 0 1 Y 1 þ ZY
Z = jX Z = jX
YS2 YS1
Figure 6.25 Circuit model of a linear array of slots modelled as Tee networks
with different spacings. IEEE 2010, reprinted from [41]
132 Slotted waveguide array antennas: theory, analysis and design
References
[1] Stevenson A. F. Series of Slots in Rectangular Waveguides. National
Research Council of Canada Radio Reports 12 and 13, 1944.
[2] Elliott R. S. ‘On the design of traveling-wave-fed longitudinal shunt slot
arrays’. IEEE Transactions on Antennas and Propagation. 1979;27(5):
717–720.
[3] Montgomery C. G., Dicke R. H., Purcell E. M. Principles of Microwave
Circuits, MIT Radiation Laboratory Series, vol. 8. McGraw-Hill, New York,
1948.
[4] Collin R. E. Foundations for Microwave Engineering. McGraw-Hill, New
York, 1966.
[5] Stegen R. J. ‘Slot radiators and arrays at X-band’. IRE Transactions on
Antennas and Propagation. 1952;1(1):62–84.
[6] Stegen R. J., Reed R. H. ‘Arrays of closely-spaced nonresonant slots’. IRE
Transactions on Antennas and Propagation. 1954;2(3):109–113.
[7] Kaminow I. P., Stegen R. J. Waveguide Slot Array Design. Hughes Aircraft
Company, Culver City, CA, Tech. Memo No. 348, 1954.
[8] Dion A. ‘Nonresonant slotted array’. IRE Transactions on Antennas and
Propagation. 1958;6(4):360–365.
[9] Elliott R. S. ‘A note on the relation between slot conductance and slot
voltage’. IEEE Transactions on Antennas and Propagation. 1984;32(10):
1133–1134.
[10] Elliott R., Kurtz L. ‘The design of small slot arrays’. IEEE Transactions on
Antennas and Propagation. 1978;26(2):214–229.
[11] Elliott R. S. ‘An improved design procedure for small arrays of shunt slots’.
IEEE Transactions on Antennas and Propagation. 1983;31(1):48–53.
[12] Harrington R. F. Time-Harmonic Electromagnetic Fields. IEEE Press,
New York, 2001.
[13] Gulick J., Stern G., Elliott R. S. ‘The design of linear and planar arrays
of waveguide-fed longitudinal slots’. IEEE International Antennas and
Propagation Symposium Digest. 1986;2:685–688.
[14] Kay A. F., Simmons A. J. ‘Mutual coupling of shunt slots’. IRE Transactions
on Antennas and Propagation. 1960;8(4):389–400.
[15] Elliott R. S. Antenna Theory and Design. IEEE Press, Hoboken, NJ, 2003.
[16] Press W. H., Teukolsky S. A., Vetterling W. A., Flannery B. P., Metcalf M.
Numerical Recipes: The Art of Scientific Computing, 3rd Edition. Cam-
bridge University Press, Cambridge, 2007.
[17] Mazzarella G., Panariello G. ‘On the evaluation of mutual coupling between
slots’. IEEE Transactions on Antennas Propagation. 1987;35(11):2189–
2192.
[18] Nauwelaers B., van de Capelle A. ‘Series expressions for mutual coupling
between magnetic currents’. Proceedings of the International Conference on
Aerospace Application, Torino, Italy, 1989, pp. 199–202.
The linear slotted waveguide array antenna 133
[35] Silver S. Microwave Antenna Theory and Design. MIT Radiation Laboratory
Series. vol. 12, McGraw-Hill, New York, 1949.
[36] Yee H. Y. ‘The design of a large waveguide array of shunt slots’. IEEE
Transactions on Antennas and Propagation. 1992;40(7):775–781.
[37] Bucci O. M., Mazzarella G., Panariello G. ‘A note on edge-shunt slot array
design’. IEEE International Antennas and Propagation Symposium Digest.
vol. 3, 1997. pp. 1946–1949.
[38] Rengarajan S. R., Derneryd A. G. ‘Application of compound coupling slots
in the design of shaped beam antenna patterns’. IEEE Transactions on
Antennas and Propagation. 1993;41(1):59–65.
[39] Rubio J., Corcoles J., de Aza M. A. G. ‘Inclusion of feeding effects in the
generalized scattering matrix formulation of a finite array’. IEEE Antennas
and Wireless Propagation Letters. 2009;8:819–822.
[40] Montesinos-Ortego I., Zhang M., Sierra-Perez M., Hirokawa J., Ando M.
‘Systematic design methodology for one-dimensional compound slot-arrays
combining Method of Moments, equivalent circuit model and forward
matching procedure’. IEEE Transactions on Antennas Propagation. 2013;
61(1):453–458.
[41] Costanzo S., Casula G. A., Borgia A., et al. ‘Synthesis of slot arrays on
integrated waveguides’. IEEE Antennas and Wireless Propagation Letters.
2010;9:962–965.
[42] Pozar D. M. Microwave Engineering. John Wiley and Sons, Hoboken, NJ,
2012.
[43] Shavit R., Elliott R. ‘Design of transverse slot arrays fed by a boxed
stripline’. IEEE Transactions on Antennas and Propagation. 1983;31(4):
545–552.
[44] Robertson R., Elliott R. ‘The design of transverse slot arrays fed by the
meandering strip of boxed stripline’. IEEE Transactions on Antennas and
Propagation. 1987;35(3):252–257.
[45] Elliott R. S., Stern G. J. ‘The design of microstrip dipole arrays including
mutual coupling’. IEEE Transactions on Antennas and Propagation. 1987;
35(3):252–357.
Chapter 7
Design of planar slotted waveguide
array antennas
In this chapter, we will present Elliott’s design procedure for planar standing wave
slot arrays including a detailed design example with computed and measured
results. For large planar arrays a modification to Elliott’s design procedure using an
infinite array model is sometimes preferred. Large standing wave arrays can be
broken up into sub-arrays with a parallel feed network to improve bandwidth.
Important parameters are the total normalised slot conductance of radiating wave-
guides and the total normalised resistance of feed waveguides. Additional examples
of slot array designs will be presented, including a procedure for designing a tra-
velling wave feed to excite radiating waveguides with either standing wave or
travelling wave slot arrays. Other design and analysis methods in the literature will
also be reviewed.
7.1 Introduction
Planar slot arrays are employed in applications requiring pencil beams by stacking
several linear arrays or waveguide sticks containing radiating slots next to each
other. They are preferable to conventional parabolic reflector antennas since they
can be designed to provide greater control of the pattern, especially of the sidelobe
levels. They are used in numerous radar and communication systems in ground and
space applications. The use of all metal waveguides in slot arrays is attractive in
deep space applications where severe radiation environment is encountered. Since
slot arrays occupy small volumes, they are easy to deploy.
Figure 7.1 shows a planar array consisting of six radiating waveguides stacked
next to each other with six longitudinal radiating slots cut in the broad wall of each.
Each radiating waveguide is shorted at a distance of a quarter guide wavelength
from the centre of the radiating slot at each end. The radiating slots are cut on a flat
plate which will form the top broad walls of all the radiating waveguides stacked
next to each other. Therefore a planar array like this is also called a flat plate array.
The radiating waveguides are fed by coupling slots cut on top broad wall of a feed
waveguide placed underneath the radiating waveguides and orthogonal to them.
Resonant centred-inclined coupling slots are commonly used in the design. Other
types of coupling slots will be discussed in Chapter 8.
136 Slotted waveguide array antennas: theory, analysis and design
In planar arrays, mutual coupling between radiating slots in the E-plane is sig-
nificant and needs to be included in the design and analysis. Inclusion of mutual
coupling is generally a challenging process. Planar slot arrays have been reported in
the literature since 1958 [1–4]. The design of a planar array was presented by Elliott
[5] that included the model of the coupling slots between the feed waveguide and the
radiating waveguides. The method of accounting for the internal higher-order mode
coupling, discussed in Chapter 6 may be applied to the design of planar arrays as
well. We will discuss Elliott’s procedure for planar slot arrays in detail since it
provides an accurate procedure for the designs of slot arrays. For large arrays
the computational burden of Elliott’s method becomes impractical, because of the
large amount of external mutual coupling computations. The method based on the
infinite array mutual coupling model is a viable alternative. Large arrays are often
broken into sub-arrays. The sub-array architecture is discussed with some examples,
followed by travelling wave feeds used for planar arrays. The chapter concludes with
a discussion on various design and analysis methods found in the current literature.
6
(6,1) (6,3) (6,5)
(6,2) (6,4) (6,6)
4
(4,1) (4,3) (4,5)
(4,2) (4,4) (4,6)
(3,1) (3,3) 3
(3,2) (3,4) (3,5)
(3,6)
2
(2,1) (2,3) (2,5) (2,6)
(2,2) (2,4)
for complex excitations as well. However, it should be remembered that the amount
of phase control achievable for resonant elements is limited, generally within 60
from the resonant values.
Let the radiating waveguides be numbered from i ¼ 1 through M with the first
one near the input end of the feed waveguide. There is a short circuit at a distance
equal to a half guide wavelength from the centre of the last coupling slot. The
centred-inclined coupling slot behaves like a series impedance because of its odd-
symmetric scattering, as explained in Chapter 8. Therefore a short is placed a half
guide wavelength away from the last coupling slot in the standing wave array.
Radiating slots in each radiating waveguide are numbered with two indices, the first
one, i being the number of the radiating waveguide while the second, j ¼ 1 through
N, denotes the radiating slot number in waveguide i. The first radiating slot is on the
left of the feed waveguide when the feed waveguide propagation direction is upward
as shown in Figure 7.2. For simplicity the value of N is assumed to be the same for
each radiating waveguide but the design procedure is valid even if it is not the case.
138 Slotted waveguide array antennas: theory, analysis and design
The first design equation for linear arrays given in (6.16) becomes (7.1) for
planar arrays.
The subscripts i and j denote the radiating waveguide number and the slot number in
that waveguide respectively. The reference slot 11 is the first slot in the first radiating
waveguide. The magnitude of the mode voltage of all slots in a radiating waveguide
is the same because of the half guide wavelength spacing between adjacent slots at
the design frequency. Therefore mode voltages do not appear in (7.1). A resonant
coupling slot acts as a transformer with the coupling coefficient ci given by
where S11 is the reflection coefficient of the four-port centred-inclined coupling slot
with all the ports match-terminated. In Chapter 8 coupling slot models are dis-
cussed in detail. Clearly the active admittance of a radiating slot is proportional to
the slot voltage or the excitation coefficient whereas it is inversely proportional to
the coupling coefficient of the coupling slot in that radiating waveguide. Obviously
if the coupling coefficient is greater, a smaller value of active admittance is needed
to realise a given slot excitation.
The terms fij and Yija =G0 in (7.1) have been previously defined in (6.12) and
(6.15), respectively, with a single subscript for each. They are reproduced in (7.3)
and (7.4) in terms of double subscripts corresponding to those of each radiating
slot. Equation (7.4) is a combination of (6.15), (6.17) and (6.39).
p
cosðb10 Lij Þ
2kLij
fij ¼ fij ðx0ij ; Lij Þ ¼ cosðp x0ij =aÞ 2 2 (7.3)
p b10
2kLij k
The mutual coupling term in (7.4) is obtained by combining (6.17) and (6.39),
and it is given by
ða=lÞ3 X M X N
0 s
MCij ¼ jðb10 =kÞðk0 bÞ V gmn;ij ðLmn ; dmn ; Lij ; dij Þ
Vij m¼1 n¼1 mn
s
b h i
þ j 10 eðg20 dÞ hij hi; j1 Vi;s j1 þ hij hi; jþ1 Vi;s jþ1 (7.4a)
g20
Design of planar slotted waveguide array antennas 139
where
p coshðg20 Lij Þ
hij ¼ 2 cosð2pdij =aÞ (7.4b)
2kLij p 2
ðg20 =kÞ2 þ
2kLij
and gmn;ij is the expression for external mutual coupling between radiating slots mn
and ij, defined previously in (6.18) and (4.115). The prime in the summation
indicates that m 6¼ i and n 6¼ j simultaneously. In a planar array, the E-plane
coupling between slots in different waveguides generally becomes substantial.
The normalised self-admittance of the longitudinal radiating slot ij, Yij =G0 , is
obtained from the two-port scattering parameter of the slot with the reference plane
passing through the centre of the slot. S11ij is given by (7.5) which is the same as
(6.45) with double subscripts.
S11ij
Yij =G0 ¼ 2 (7.5)
1 þ S11ij
Start
Choose M, N,
a, b, f, w, t
Read slot data
lter = 1
Solve 2N or 2N + 1
simultaneous eqns. for rad
lter = lter +1 waveguides 1 through M
lter > 1
and
No Δ < tol?
Yes
Stop
total of 2N unknowns for this waveguide. For the slot 11 (i ¼ 1 and j ¼ 1), (7.1) is
not relevant. However, we obtain one equation for slot 11 by choosing the ima-
a
ginary part of the active admittance of Y11 to be zero. Thus for the first radiating
waveguide we have 2N equations in 2N unknowns. Thus there are as many equa-
tions as the number of unknowns for each radiating waveguide. However, the
simultaneous equations are nonlinear. The total normalised slot conductance G tot
mentioned above is discussed in [7] and in Section 7.4.
At the start of the first iteration, the initial values of alternating slot offsets for
all radiating slots are chosen arbitrarily and their lengths have been set to equal
their resonant values. The choice of these values is not critical since they will be
updated to more accurate values in subsequent iterations. Recall that in Section 6.3
there was a discussion on the iterative design for linear standing wave arrays where
it was stated that 2N equations in 2N unknowns were solved using a quasi-Newton
method. The same technique is used in the planar array by solving the non-linear
equations of one radiating waveguide at a time [8]. After solving for the unknowns
in all radiating waveguides, we iterate the process. The iteration is stopped when
Design of planar slotted waveguide array antennas 141
changes in linear dimensions of slots between successive iterations are less than
achievable values of manufacturing tolerance. In order to determine the manu-
facturing tolerance required to achieve the desired antenna performance, designers
evaluate how changes in slot parameters such as lengths, offsets and tilts on
the antenna characteristics affect the antenna performance, for example, gain,
sidelobe level and cross-polarisation. Based on such studies it may be possible to
set the tolerance needed to achieve the required performance. At low frequencies,
tolerance may not be an issue. The stopping criterion for iterations may be when
change in linear dimensions is below a predetermined level, obtained by a process
similar to that used to determine the tolerance. Experience has shown that good
design solutions are obtained typically after five iterations. The iterative design
process yields the offsets and lengths of all slots and all the coupling coefficients
relative to that of the first coupling slot. By enforcing a match at the input port or by
equating the total normalised resistance of the feed waveguide to unity, we deter-
mine the coupling coefficient c1 from (7.6), and hence all ci as follows.
X
M
in ¼ c21
R tot;i c
G 2i ¼ 1 (7.6)
i¼1
where
i ¼ ci =c1
c (7.7)
is the normalised coupling coefficient of the coupling slot i relative to that of the
coupling slot 1 and
X
N
tot;i ¼
G Yij =G0 (7.8)
j¼1
Using (7.2), the scattering parameter S11i of the ith coupling slot in a four-port
coupler is determined. Then from the original computed data of coupling slots one
can interpolate the resonant length and the tilt angle corresponding to each S11i .
45.46 mm. The wall thickness of each radiating waveguide is 0.508 mm while the
coupling slot thickness is 1.27 mm. When radiating waveguides are stacked next
to each other spacing between centres of adjacent guides is equal to 21.71 þ 2
0.508 ¼ 22.73 mm. This is exactly equal to half guide wavelength in the feed
waveguide at 9.3 GHz and is also the spacing between adjacent coupling slots. There
is a short at a distance of half guide wavelength from the coupling slot farthest from
the feed end. The radiating slots have alternating offsets with half guide wavelength
spacing, which is 24.055 mm. At each end of every radiating waveguide, there is a
short at a distance of a quarter guide wavelength from the nearest slot centre.
We obtain computed slot data using the moment method solution to the perti-
nent coupled integral equations of isolated radiating or coupling slot. Such data may
be obtained from any computational electromagnetics code or from experimental
measurements. For a range of tilt angles of coupling slots, resonant lengths and |S11|
are determined with all the four ports match-terminated. The reference planes of all
the four ports pass through the centre of the coupling slot. Table 7.1 shows computed
data for coupling slots. For radiating slots, we determine the resonant length for a
range of values of offsets. Then for each offset we determine |S11| and arg S11 for a
range of values of slot lengths normalised to the resonant length. Radiating slot data
in Tables 7.2–7.4 are similar to those in Tables 6.2–6.4 in Chapter 6. In the design,
the total normalised slot conductance in each radiating waveguide was set at 2.8 so
that magnitudes of slot offsets would not be too small. The total normalised
Table 7.3 Magnitude of S11 as a function of offset (mm) and normalised length
(mm)
¼ L=Lres 0.94
L 0.955 0.97 0.985 1.0 1.015 1.03 1.045 1.06
jdj
0.333 0.0027 0.0031 0.0034 0.0037 0.0038 0.0037 0.0034 0.0032 0.0029
0.667 0.0107 0.0122 0.0136 0.0146 0.0148 0.0144 0.0136 0.0126 0.0115
1.000 0.0239 0.0271 0.0300 0.0320 0.0326 0.0318 0.0300 0.0278 0.0256
1.333 0.0421 0.0474 0.0521 0.0553 0.0562 0.0548 0.0519 0.0483 0.0446
1.667 0.0646 0.0722 0.0789 0.0833 0.0845 0.0826 0.0786 0.0735 0.0683
2.000 0.0908 0.101 0.109 0.115 0.116 0.114 0.109 0.102 0.0957
2.333 0.121 0.133 0.142 0.149 0.150 0.147 0.141 0.134 0.126
2.667 0.153 0.166 0.177 0.184 0.185 0.182 0.176 0.168 0.159
3.000 0.187 0.201 0.212 0.219 0.221 0.217 0.211 0.202 0.192
Table 7.4 Phase of S11 in degrees as a function of offset and normalised length
¼ L=Lres
L 0.94 0.955 0.97 0.985 1.0 1.015 1.03 1.045 1.06
jdj
0.333 225.5 216.4 205.4 193.0 180.1 167.9 157.5 148.9 142.0
0.667 225.0 216.0 205.1 192.8 180.1 168.2 157.8 149.3 142.5
1.000 224.2 215.2 204.5 192.5 180.2 168.6 158.5 150.1 143.3
1.333 222.8 213.9 203.3 191.7 179.8 168.6 158.9 150.8 144.2
1.667 221.5 212.7 202.4 191.3 179.9 169.3 160.0 152.1 145.6
2.000 220.0 211.3 201.4 190.8 180.1 170.1 161.1 153.6 147.2
2.333 217.9 209.4 199.9 189.9 179.9 170.5 162.1 154.9 148.8
2.667 216.1 207.9 198.9 189.4 180.1 171.4 163.5 156.7 150.8
3.000 213.8 205.9 197.4 188.6 180.0 171.9 164.6 158.2 152.6
resistance in the feed waveguide was set to unity to achieve a match. Table 7.5 shows
the tilt angles and lengths of the six coupling slots. The values of offsets and lengths
of radiating slots are shown in Table 7.6. All data in Tables 7.5 and 7.6 were
obtained using the design procedure mentioned in Section 7.2.1.
144 Slotted waveguide array antennas: theory, analysis and design
Table 7.6(a) Offsets and lengths of radiating slots obtained from the design
Table 7.6(b) Offsets and lengths of radiating slots obtained from the design
0 H-plane realised
H-plane ideal
–10 E-plane realised
Pattern level in dB
E-plane ideal
–20
–30
–40
–50
–90 –60 –30 0 30 60 90
Theta in degrees
E-plane
–10
H-plane
Pattern level in dB
–20
–30
–40
–50
–90 –60 –30 0 30 60 90
Theta in degrees
Chapter 8. It is assumed that the slots are embedded in an infinite ground plane and
therefore radiation is confined to the front half space only. The ideal patterns in
Figure 7.4 are obtained with the specified aperture distribution with the actual slot
locations. The electric field of each slot is assumed to have a half-cosine amplitude
and uniform phase distribution in the ideal pattern whereas in the realised pattern
the computed aperture distribution for each slot is used. Small discrepancies
between the ideal and realised patterns are attributed to imperfections in the design
procedure and implementation. Figure 7.5 shows the measured pattern of an array
fabricated with the use of design data shown in Tables 7.5 and 7.6. The discrepancy
between computed and measured patterns, especially in the far out sidelobe region,
is attributed primarily due to diffraction from the edges of the finite ground plane.
Manufacturing tolerances may also account for some discrepancy between theory
and measurement.
In large arrays with slowly varying aperture distributions, all slots except the
ones near the edges are surrounded by approximately identical neighbouring
elements but for alternating signs for the offsets. The external mutual coupling
expression for a representative slot n is then obtained using a simplified infinite
array model. Infinite array mutual coupling model has been used in a number of
antenna arrays. In microstrip reflectarrays the infinite array mutual coupling
model has been used successfully for many designs and analyses [10]. It has
been shown that the infinite array model for slot arrays yields errors for only a
couple of rows of slots near the E-plane edges. The accuracy of the model has
been established for moderate to large arrays [11,12]. In this model, slot n is
embedded in two doubly periodic infinite arrays, one with positive offsets and
146 Slotted waveguide array antennas: theory, analysis and design
a/2 + d Slot n dx
dz
the other with negative offsets, both in rectangular lattice with period 2dz along
the radiating waveguides and dx across. Figure 7.6 shows a few of the slots in the
infinite array model. The magnitude of the offset of all slots, dn, and their half
lengths, Ln, are those of slot n.
ð1Þ ða=lÞ3 XM X N
0 s
MCij ¼ jðb10 =kÞðk0 bÞ V gmn;ij ðLmn ; dmn ; Lij ; dij Þ (7.9)
Vijs m¼1 n¼1 mn
In the infinite array environment, it is assumed that all slot voltages are equal or
slowly varying. In the latter case all slot voltages are approximated to be equal.
Therefore all slot lengths are the same and the magnitudes of slot offsets are equal.
Then (7.9) becomes
ð1Þ
X
1 X
1
0
MCij ¼ jðb10 =kÞðk0 bÞða=lÞ3 gmn;ij ðLij ; dij ; Lij ; dij Þ (7.10)
m¼1 n¼1
The sign for the offset of slot mn is chosen based on its location because of alter-
nating slot offsets. We express (7.10) as follows.
ð1Þ 4ph0
MCij ¼ ðb10 =kÞðk0 bÞða=lÞ3 T (7.11)
w2
where
X
1 X
1
0
T¼ Imn;ij ðLij ; dij ; Lij ; dij Þ (7.12)
m¼1 n¼1
Design of planar slotted waveguide array antennas 147
and Imn;ij ðLij ; dij ; Lij ; dij Þ ¼ Ymn Lij w=2. Ymn is specified in (4.113) for a pair of
slots with single subscripts whereas we have double subscripts for the planar
array. It is depicted in Figure 4.15 with slot lengths a and widths b whereas in
(7.12) the half length and width are Lij and w, respectively.
Equation (7.12) may be expressed as
ðð
pzij
T¼ HV cos ds0 (7.13)
2Lij
slot ij
where the integral is carried out over slot ij, zij is its local coordinate along the axis
of the slot, and cosine weighting corresponds to its aperture distribution. HV is the
magnetic field along the location of slot ij, produced by half cosinusoidal magnetic
currents of unit peak strength at the location of all other slots, radiating in free
space in the absence of the ground plane. Clearly each term inside the summation in
(7.12) corresponds to the integral in (7.13) when the contribution to HV due to unit
magnetic current in slot mn only is considered.
2 ðð
1 @ pzmn
HV ¼ þ k 2
S cos ds0 (7.14)
j4pwm0 @V2 0
2Lij
slot mn
where
X X expðjk0 R1 Þ expðjk0 R2 Þ
S¼ þ (7.15)
m n
R1 R2
h i
2 1=2
R1 ¼ fV ðV0 þ 2mdz Þg þ fx ðx0 þ ndx Þg
2
(7.16)
h i
2 1=2
R2 ¼ fV ðV0 þ ð2m þ 1Þdz Þg þ fx ðx0 2d þ ndx Þg
2
(7.17)
and the integral is carried out over all the slot apertures except that of ij. In the
double summation in (7.15) all positive and negative integer values and zero
are used for m and n except m ¼ 0 and n ¼ 0 in (7.16). Using Poisson’s sum formula
the double summations S may be expressed in terms of double infinite Floquet
series [13] as shown in (7.18), where all positive and negative integer values and
zero are used for p and q.
pl0 0 ql0 0
exp jk ðV V Þ þ ðx x Þ
j2p X X
0
2dz dx
S¼ " 2 2 #1=2
2dz dx p q
pl0 ql0
k0 1
2dz dx
4pd
1 þ ð1Þp exp jq (7.18)
dx
148 Slotted waveguide array antennas: theory, analysis and design
" p
2 #
pl0 Lij ppLij
k02 1 w 2 2 cos (7.21)
2dz p pp dz
2Lij dz
Using (7.22) in (7.11) we obtain the expression for the external mutual coupling for
an infinite array model. It is possible to obtain scattering data for a slot embedded
in an infinite array as a function of offset and length and use such data to design a
slot array. For a large array such a design works very well and for a moderate size
array it may still yield acceptable results.
Design of planar slotted waveguide array antennas 149
In the design of a planar standing wave array, the total slot conductance of each
radiating waveguide, G, is one of the design parameters. A linear array can be
designed with a specified value of G and with an impedance transformer to match
the input port. The frequency behaviour of linear arrays has been analysed by
considering equally spaced constant series or shunt loaded transmission lines [7].
The frequency response of the shunt or series load is ignored since the bandwidth of
the array is usually small compared to that of a load. The analysis employs the
transmission matrix of each unit cell composed of a length of transmission line
equal to the spacing between successive slots and the load impedance or admittance
[14]. A larger value of G yields a greater bandwidth for return loss whereas
the bandwidth for pattern performance such as the sidelobe level is lower.
A smaller value of G will produce the opposite effect. In the case of a planar array,
the total normalised shunt conductance G and the total normalised series resistance
R of the feed waveguide are design parameters. If R is not equal to 1, there is a need
to use a transformer. Derneryd and Petersson [7] have calculated the effects of
overloading on the return loss and sidelobe performance over 5% bandwidth for a
slot array of 10 wavelength diameter circular aperture designed to provide a 30
dB sidelobe level. We reproduce their results in Figures 7.7–7.9.
–20
Sidelobe level (dB)
–25
–30
0 1 2 3 4
Overloading G
Figure 7.7 First sidelobe level in the H-plane at the upper frequency limit as a
function of overloading in radiating waveguides. Reprinted from [7],
with permission
150 Slotted waveguide array antennas: theory, analysis and design
–20
–25
–30
0 4 8 12 16
Overloading R
Figure 7.8 First sidelobe level in the E-plane at the upper frequency limit as a
function of overloading in the feed waveguide. Reprinted from [7],
with permission
4
Voltage standing wave ratio
1
0 10 20 30 40
Overloading G∙R
Figure 7.9 Voltage standing wave ratio at the upper frequency limit as a function
of overloading of the waveguides. Reprinted from [7], with permission
Figures 7.7 and 7.8 show that in order to produce good pattern performance,
especially low sidelobes over the desired bandwidth G and R have to be small.
However, a better VSWR over the desired bandwidth is obtained for a value of the
product GR ¼ 8 as shown in Figure 7.9. A choice of G ¼ 2 yields the best
Design of planar slotted waveguide array antennas 151
compromise for the performance of VSWR and sidelobe level over frequencies.
Therefore it is recommended that the values of G and R be chosen as 2 and 4,
respectively. These values are guidelines for designs since Derneryd and Peters-
son’s approximate analysis did not consider the frequency response of radiating and
coupling slots and that of the impedance transformer.
hence the maximum length of line in two orthogonal directions corresponds to one
and two guide wavelengths respectively. This architecture with small sub-arrays
was able to meet low sidelobe levels and low levels of reflection coefficients over
approximately 5% bandwidth. This array designed with an even symmetry for the
slot offsets in the H-plane helped in reducing the grating lobes called butterfly lobes
produced by the conventional slot arrays with alternating positive and negative
offsets. Butterfly lobes are discussed in Section 9.9. Section 7.6.2 presents results
for sidelobe levels of the slot arrays exhibiting symmetry in the two halves of the
H-plane and for the array with alternating slot offsets along the entire H-plane.
(a)
(b)
Figure 7.11 One of four modules of a stick, each consisting of four 10 10 sub-
arrays: (a) Radiating side of the module. Four modules are combined
in one stick. (b) Rear side of one module showing power dividers.
IEEE 2009, reprinted from [15], with permission
Angled
Radiating slot layer, coupling slot
single-clad braze
sheet, AI Brazing short
Main body,
double-sided AI Offset input slot
hog-out Power divider,
Input sheet, double- single-sided AI
clad braze sheet, AI hog-out
Figure 7.12 An 8 8 array divided into two sub-arrays. IEEE 2010, reprinted
from [16], with permission
The lower layer contains the corporate power divider made up of H-plane
T junctions. A single H-plane T junction shown in Figure 7.15(a) divides
the incident power into the desired power ratio. In this array all T junctions are
designed to provide equal power division. The centred septum length lf and the iris
width wf are chosen to provide a match at the input port as long as the branch ports
154 Slotted waveguide array antennas: theory, analysis and design
75 mm
Aperture size
16 × 4.2 = 67.2 mm
76 mm
Figure 7.13 A 16 16 array divided into two sub-arrays. IEEE 2011, reprinted
from [17], with permission
Radiating slot
Cavity
Coupling aperture
Full-corporate-feed
waveguide
Feed aperture
(WR15)
z Radiating slot
Figure 7.14 A schematic view of a quadrant of the array with H-plane T junctions
in the lower layer in black and the upper layer consisting of radiating
slots. IEEE 2011, reprinted from [17], with permission
Design of planar slotted waveguide array antennas 155
Port 4 Port 2
lf
wf
Figure 7.15 H-plane power dividers: (a) a single H-plane T-junction divider;
(b) three H-plane T’s for four-way power. IEEE 2011,
reprinted from [17], with permission
Magnetic field
are matched. It is generally possible to design such a T junction with greater than
20 dB return loss over a broad bandwidth. Figure 7.15(b) shows that a combination
of three H-plane T’s can provide one to four way power division. With the use of
additional H-plane T junctions one can achieve the required power division for the
entire array.
Figure 7.16 shows the method of exciting each of the two radiating waveguides
in each sub-array. The two radiating waveguides are excited by a coupling window
which is a longitudinal offset slot in the waveguide in the lower layer. The coupling
window with its magnetic field in the longitudinal direction will excite each of the
156 Slotted waveguide array antennas: theory, analysis and design
two radiating waveguides equally in amplitude but with 180 out of phase.
Therefore, in order to excite the four slots with equal phase, the offsets of slots in
the same waveguide are on the same side of the centreline whereas the offsets of
slots in the two waveguides are on opposite sides. The slot spacing is between
0.812 and 0.925 wavelengths in the frequency range of operation. In the array
design, external mutual coupling was modelled based on an infinite array of four-
element sub-arrays. Uniform aperture distribution was desired but because of lim-
itations in the infinite array mutual coupling model the sub-arrays near the edges
had significant error. This array achieved a return loss better than 14 dB over 8.3%
bandwidth and better than 80% efficiency. Efficiency is the ratio of the antenna
gain to the ideal directivity of the antenna assuming uniform aperture distribution.
An efficiency of 80% corresponds to a value of gain equal to 1 dB below the ideal
directivity. It is possible to achieve such a broadband performance in a slot array if
many small sub-arrays are used. However, for large arrays containing thousands of
elements, the power divider may become very complicated and one may have to
use somewhat larger sub-arrays.
Radiating Input
slot slot
Coupling
slot
Feed
guide
Radiating Input
guide guide
Figure 7.17 A sub-array containing 100 radiating slots, 10 coupling slots and an
input shunt-series coupling slot
with respect to the axis of the feed waveguide. At the centre of the feed waveguide
of each 10 10 sub-array, there is a shunt-series coupling slot that matches a
normalised resistance of 4.0 to the input port on one side. The other side of the
input waveguide is shorted at a distance of a quarter guide wavelength from the
centre of the slot. The shunt series coupling slot has an offset of 2.081 mm and
length equal to 4.242 mm. It is centred transverse to the feed waveguide. Fig-
ure 7.17 shows the slots in a sub-array.
Figures 7.18 and 7.19 show the E-plane and H-plane patterns of a 10 40
module, respectively. The computed patterns are obtained from the moment
method analysis of an infinite array of 10 10-element sub-arrays with uniform
excitations of all sub-arrays. From the moment method solution we determine the
pattern of each sub-array and then determine the pattern of four sub-arrays in the
10 40 module. Measured patterns are in very good agreement with computed
plots except in the far-out sidelobe regions. The discrepancies may be attributed to
the fact that the computation assumes an infinite ground plane and therefore it does
not account for the edge diffraction from the finite ground plane. In addition,
imperfections in the fabrication process, especially the 1 mil (25 m) mechanical
tolerance in the manufacturing process may account for some error. Reflection
158 Slotted waveguide array antennas: theory, analysis and design
0
Meas, 35.66 GHz...... calc
–10
Relative magnitude (dB)
–20
–30
–40
–50
–60
–90 –60 –30 0 30 60 90
Theta (deg)
0
Meas, 35.66 GHz...... calc
–10
Relative magnitude (dB)
–20
–30
–40
–50
–60
–90 –60 –30 0 30 60 90
Theta (deg)
coefficient plots shown in Figure 7.20 indicate some discrepancy, especially at the
frequency where the 10 40 module is tuned, partly because of errors in the fab-
rication process. Figure 7.21 shows the H-plane patterns of 16 sticks and the pattern
computed by the moment method. The light traces corresponding to the patterns of
Design of planar slotted waveguide array antennas 159
0
Measured
10 × 10s
–10
Refl. coefficient (dB)
Calculated
–20 10 × 10
Measured
–30 power divider
–40
33.66 34.16 34.66 35.16 35.66 36.16 36.66 37.16 37.66
f (GHz)
Figure 7.20 Reflection coefficient of each sub-array and the power divider of a
10 40 module. IEEE 2009, reprinted from [15], with permission
–10
dB
–20
–30
–40
–9 –6 –3 0 3 6 9
Theta (deg)
different sticks are not identical because of the repeatability problem associated
with the fabrication process. Grating lobes near 8 are due to the wall thickness
between sub-arrays. A key requirement of the array design and fabrication was to
align the 16 beams in the E-plane radiated by each of the 16 sticks, within 1/10th of
the beamwidth or 0.042 . This objective was met.
160 Slotted waveguide array antennas: theory, analysis and design
Uniform phase distribution of slot apertures was achieved with 180 phase
difference between the excitations of the two feed waveguides as shown in
Figure 7.23. Tilt angles of coupling slots in the two sub-arrays are in the same
direction in Figure 7.22 whereas they are in opposite directions in Figure 7.23.
Table 7.8 shows the sidelobe levels achieved for the two arrays shown in Fig-
ures 7.22 and 7.23. The symmetric design yields lower sidelobes over the angular
ranges beyond 30 whereas it increases the sidelobe level slightly for angular
region below 30 .
Far-field patterns of the S-band antenna, both computed by the finite element
technique HFSS and measured using the spherical near-field technique are found
to have very good agreement over nearly 60 dB dynamic range [16]. They are
reproduced in Figures 7.24 and 7.25. The spherical far field pattern is mapped onto
a circle, with q varying from 0 to 180 radially and f varying from 0 to 360
azimuthally. q ¼ 180 point in the far field sphere is stretched into the cir-
cumference. Average sidelobe levels in different angular ranges obtained from
computed and measured data are compared in Table 7.9. Excellent agreement
between two sets of data is noted and the design margin is found to be more than
adequate.
162 Slotted waveguide array antennas: theory, analysis and design
Figure 7.23 Symmetric array architecture. IEEE 2010, reprinted from [16],
with permission
180° 0
150
–10
100
–20
50
θsinφ
dB
0° –30
–50
–40
–100
–50
–150
–180° –60
180°
–180°
180° 0
150
–10
100
–20
50
θsinφ
dB
0° –30
–50
–40
–100
–50
–150
–180° –60
–150 –100 –50 0° 50 100 150
–180°
180°
θcosφ
Figure 7.25 Far-field pattern of the S-band antenna measured at 2.6 GHz.
IEEE 2010, reprinted from [16], with permission
Table 7.9 Average sidelobe for the S-band antennas ( IEEE 2010, reprinted
from [16] with permission)
Figure 7.26 A planar array of longitudinal slots with a travelling wave feed.
IEEE 2011, reprinted from [22], with permission
Y
4
f
X
1 2
feed waveguide and two in the radiating waveguide. The scattering matrix of such a
coupler at resonance (see Figure 7.27) is given by [23]
2 3 2 3
S11 S12 S13 S14 r p q q
6 S21 S22 S23 S24 7 6 p
7 6 r q q 7 7
½S ¼ 6
4 S31 S32 S33 S34 5 ¼ 4 q q p (7.23)
r 5
S41 S42 S43 S44 q q r p
where p ¼ (1 r) and q ¼ [r(1 r)]1/2. If the S-matrix for a positive value of tilt
angle f is S(r,p,q), then for a negative value of f the S-matrix becomes S(r,p,q).
The tilt angle f is shown in Figure 7.27 with ports 1 and 2 in the feed waveguide
and 3 and 4 in the radiating waveguide. If ports 3 and 4 are terminated in reflection
coefficients G3 ¼ G4 ¼ G (¼0 if ports 3 and 4 are matched), the coupler may be
represented as a two-port network.
" 0 #
r 1 r0
½S2 ¼ (7.24)
1 r0 r0
where
0
–10
Pattern level in dB
–20
2.2% PL
–30 3.6% PL
Desired
–40
–50
–60
–90 –60 –30 0 30 60 90
Theta in degrees
Figure 7.28 E-plane pattern of the planar array with an 80-element travelling
wave array feed (PL ¼ load power/incident power in per cent).
IEEE 2011, reprinted from [22], with permission
360
Excitation phase (deg.)
270
2.2% PL
180
3.6% PL
90
0
0 20 40 60 80
Element number
Figure 7.29 Excitation phase of the radiating waveguides (PL ¼ relative power in
the load). IEEE 2011, reprinted from [22], with permission
35
30
Tilt angle (deg)
25
2.2% PL
20 3.6% PL
15
10
5
0 20 40 60 80
Element number
Figure 7.30 Tilt angles of coupling slots (PL ¼ relative power in the load).
IEEE 2011, reprinted from [22], with permission
168 Slotted waveguide array antennas: theory, analysis and design
wavelength ¼ 1.4741 l0 and l0 is the free space wavelength. In this case the main
beam in the E-plane will be at q ¼ 93 or 3 off normal if the radiating slots in each
waveguide are designed to be of standing-wave type with uniform phase aperture
distribution. Figure 7.28 shows the E-plane pattern of the array shown in Figure 7.26
for two designs. The power dissipated in the load relative to that at the input port for
the two cases are 2.2% and 3.6%, respectively. The desired case shows the pattern
with ideal phase distribution of the incident TE10 mode wave but for the 180 phase
reversal due to alternating slot offsets. The actual phase distribution deviates from
the ideal linear phase due to the presence of reflections at each slot. The phase
distributions for the two designs are illustrated in Figure 7.29. The effect of such
phase deviations on the patterns is not significant. Figure 7.30 shows the tilt angles of
coupling slots. Clearly the elements close to the feed have small amounts of coupling
and hence relatively small values of tilt angles as discussed in Chapter 6. Slots with
small values of tilt are sensitive to manufacturing tolerances. In the design of shaped
patterns asymmetric amplitude distributions with larger amplitudes near the feed end
are preferred to reduce the effects of manufacturing tolerances.
t1 t2
Feed B1 B2 B3
Inclined slot
Figure 7.31 A planar array of slots modelled by three blocks. IEEE 2001,
reprinted from [42], with permission
[a1i ] [ao1]
Slot 1
[S1]
[b1i ] [bo1] Radiation
Feed
network [a2i ] [ao2] network
[ai] [SF] [SH]
Slot 2
[S2]
[b2i ] [bo2]
[bi] [aH]
[bH]
[api ] [aop]
Slot p
[Sp]
[bpi ] [bop]
an inclined coupling slot of a certain thickness t1. The coupling slot is treated as a
waveguide of length t1. Block B3 represents the exterior half space region. It is
connected to each branch line or radiating waveguide through radiating slots of
thickness t2. Each block is characterised by generalised scattering matrices deter-
mined by a full wave method such as the commercial code HFSS. The inter-
connection between blocks by slots or waveguides is modelled in terms of a few
modes. Thus the generalised scattering matrix can be used for analysis as well as
design and optimisation. One of the limitations of this model is that the singularity in
the coupling term, if present, cannot be modelled in terms of the eigen function
expansion or modes used in the generalised scattering matrix. The singularity in the
Green’s functions in the coupling terms is present when a coupling slot is projected
on the wall containing radiating slot overlaps with the radiating slot [9]. This is
discussed in detail in Chapter 8.
Figure 7.32 shows the generalised scattering matrix (GSM) model for a planar
slot array developed by Enneking et al. [43]. The GSM of the feed network is
Design of planar slotted waveguide array antennas 171
–5
–10
–15
–20
dB
–25
–30
–35
–40
–80 –60 –40 –20 0 20 40 60 80
(a) Theta (deg.)
–5
–10
–15
dB
–20
–25
–30
–35
–40
–80 –60 –40 –20 0 20 40 60 80
(b) Theta (deg.)
References
[1] Saltzman H., Stavis G. ‘A dual beam planar antenna for Janus type Doppler
navigation systems’. IRE National Convention Record. 1958, pp. 240–247.
[2] Simmons A. J., Giddings O. M. ‘A multiple-beam two-dimensional wave-
guide slot array’. IRE International Convention Record. 1963, pp. 56–69.
[3] Miller J. R., Forman R. J. ‘A planar slot array with four independent beams’.
IEEE Transactions on Antennas and Propagation. 1966;14(5):560–566.
[4] Elliott R., Kurtz L. ‘The design of small slot arrays’. IEEE Transactions on
Antennas and Propagation. 1978;26(2):214–229.
[5] Lo Y. T., Lee S. W. (eds.). Antenna Handbook. Van Nostrand Reinhold,
New York, 1988, p. 12–1.
[6] Elliott R. S., O’Loughlin W. R. ‘The design of slot arrays including internal
mutual coupling’. IEEE Transactions on Antennas and Propagation. 1986;
34(9):1149–1154.
[7] Derneryd A., Petersson R. ‘Bandwidth characteristics of monopulse slotted
waveguide antennas’. Proceedings of the Fourth International Conference
on Antennas and Propagation, 1985, pp. 27–30.
[8] Press W. H., Teukolsky S. A., Vetterling W. T., Flannery B. P. Numerical
Recipes: The Art of Scientific Computing, 3rd edition. Cambridge University
Press, Cambridge, 2007.
[9] Rengarajan S. R., Shaw G. M. ‘Accurate characterization of coupling junc-
tions in waveguide-fed planar slot arrays’. IEEE Transactions on Microwave
Theory and Techniques. 1994;42(12):2239–2248.
[10] Pozar D. M., Targonski S. D., Syrigos H. D. ‘Design of millimeter wave
microstrip reflectarrays’. IEEE Transactions on Antennas and Propagation.
1997;45(2):287–296.
[11] Yee H. Y. ‘The design of a large waveguide arrays of shunt slots’. IEEE
Transactions on Antennas and Propagation. 1992;40(7):775–781.
[12] Scharstein R. W. ‘Mutual coupling in a slotted phased array, infinite E-plane
and finite in H-plane’. IEEE Transactions on Antennas and Propagation.
1990;38(8):1186–1191.
Design of planar slotted waveguide array antennas 173
[30] Ando M., Zhang M., Lee J., Hirokawa J. ‘Design and fabrication of milli-
meter wave slotted waveguide arrays’. Proceedings European Conference
on Antennas and Propagation, 2010, pp. 1–6.
[31] Volakis J. (ed.). Antenna Engineering Handbook. McGraw-Hill, New York,
2007. p. 9–1.
[32] Enjiu R. K., Protoni M. B., Andre S. ‘Slotted waveguide antenna design
using 3D EM simulation’. Microwave Journal. 2013;56(7):72–83.
[33] Gulick J. J., Elliott R. S. ‘A general design procedure for small slot arrays’.
Proceedings of the IEEE International Antennas and Propagation Sympo-
sium, 1987, pp. 302–305.
[34] Rengarajan S. R. ‘Improved design procedure for waveguide-fed slot arrays
using the method-of-moments analysis’. Electromagnetics. 2012;32(4):
221–232.
[35] Hamadallah M. ‘Frequency limitations on broad-band performance of shunt
slot arrays’. IEEE Transactions on Antennas and Propagation. 1989;37(7):
817–823.
[36] Coetzee J. C., Joubert J., McNamara D. ‘Off-center-frequency analysis of a
complete planar slotted-waveguide array consisting of sub-arrays’. IEEE
Transactions on Antennas and Propagation. 2000;48(11):1746–1755.
[37] Casula G. A., Mazzarella G. ‘A direct computation of the frequency
response of planar waveguide slot arrays’. IEEE Transactions on Antennas
and Propagation. 2004;52(7):1909–1912.
[38] Manuilov M., Lerer V., Sinyavsky G. ‘Fast and accurate full-wave analysis
of large slotted waveguide array antennas’. Proceedings of the 37th
European Microwave Conference, 2007, pp. 360–363.
[39] Liu J., Long Y. ‘A full-wave numerical approach for analyzing rectangular
waveguides with periodic slots’. IEEE Transactions on Antennas and Pro-
pagation. 2012;60(8):3754–3762.
[40] Rao S., Wilton D., Glisson A., ‘Electromagnetic scattering by surfaces of
arbitrary shape’. IEEE Transactions on Antennas and Propagation. 1982;
30(3):409–418.
[41] Rengarajan S. R. ‘Genetic algorithm optimization of a planar array using full
wave method of moments analysis’. International Journal of Computer
Aided RF and Microwave Engineering. 2013;23(4):430–436.
[42] Chiarandini S., Morini A. ‘Full wave CAD oriented technique for the ana-
lysis of airborne flat plate arrays’. Proceedings of the IEEE International
Antennas and Propagation Symposium, 2001, pp. 210–213.
[43] Enneking A., Beyer R., Arndt F. ‘Rigorous analysis of large finite wave-
guide-fed slot arrays including the mutual internal and external higher order
mode coupling’. Proceedings of the IEEE International Antennas and
Propagation Symposium, 2000, pp. 74–77.
[44] http://www.wasp-net.com. Retrieved March 2018.
Chapter 8
Concepts and models for advanced designs
Radiating slots
Coupling
Main slot
waveguide
Branch
waveguides
Figure 8.1 A planar array of radiating slots and coupling slots. IEEE 1994,
reprinted from [5], with permission
Figure 8.2 A centred-inclined slot coupler in the common broad wall of two
orthogonal waveguides, with ports 1 and 2 in the main waveguide
and ports 3 and 4 in the branch waveguide. IEEE 1989,
reprinted from [1], with permission
with one on top of the other. Such a slot transmits the entire power from a
waveguide in one layer to the next layer. In that case the coupling slot acts as a
one-to-one transformer.
2L
w
z
Figure 8.3 Slot coordinates at the upper broad wall of the main waveguide with
ports 1 and 2. IEEE 1989, reprinted from [1], with permission
Assuming that the reference planes of all four ports pass through the centre of
the slot, a coupling slot at resonance may be characterised by a scattering matrix
with only one unknown, that is, S11 as given in (8.1) [1].
2 3
r p q q
6 p r q q7
6 7
S ðqÞ ¼ f ðr; p; qÞ ¼ 6 7 (8.1)
4 q q p r5
q q r p
where r ¼ S11(q), p ¼ 1 – r and q ¼ [r(1 – r)]1/2.
From symmetry considerations we can obtain the scattering matrix for com-
plementary tilt angles and negative tilt angles as given by (8.2) and (8.3) [1].
S ðqÞ ¼ f ðr; p; qÞ (8.2)
S ðp=2 qÞ ¼ f ðp; r; qÞ (8.3)
Thus a q range of 0 to p/4 radians is sufficient to characterise the coupling slot.
Equation (4.68) in Chapter 4 may be used to determine the TE10 mode wave
scattered by a centred-inclined slot in the broad wall of a rectangular waveguide.
It may be shown that the forward and backward scattered waves are equal in
magnitude but 180 out of phase. A similar odd-symmetric scattering behaviour is
found when a wave is incident in a transmission line loaded by a series impedance.
Therefore a centred-inclined slot in the broad wall of a rectangular waveguide may
be modelled as a series impedance in an equivalent transmission line. The scattered
TE10 mode waves in the backward and forward direction are in phase and 180 out
of phase, respectively, with respect to the incident TE10 mode wave at resonance.
Analysis of the centred-inclined slot coupler using the MoM shows that the slot
may be modelled as a series impedance in the main line as well as in the branch
line with excellent accuracy. If the branch line ports 3 and 4 are match-terminated,
the series resistance normalised to the characteristic impedance of the TE10 mode
wave in the main line may be expressed in terms of S11 as
¼ 2S11 =ð1 S11 Þ
R (8.4)
178 Slotted waveguide array antennas: theory, analysis and design
3 4
n:1
1 jB 2
ratio n is determined at the resonant frequency. Over a frequency range of 6.6% below
the resonant frequency L ¼ L1 and C ¼ C1 were found to produce accurate values of
the susceptance. Similarly L ¼ L2 and C ¼ C2 yielded accurate susceptances over a
6.6% frequency band above resonance. By choosing the values of L and C between
the two sets they were able to produce an accurate model over a 10% frequency
band [2]. The value of n at resonance is the coupling coefficient c defined in (8.5).
1
(a)
w
δ
2L
z
(b)
Figure 8.5 Shunt-series coupling slot.: (a) cut-away view of the slot;
(b) slot parameters and localised slot coordinates. IEEE 1989,
reprinted from [3], with permission
180 Slotted waveguide array antennas: theory, analysis and design
The scattering matrix of the four-port resonant coupler with all ports defined to
be planes passing through the centre of the slot is given by
2 3
r p jq jq
6 p r jq jq 7
6 7
S¼6 7 (8.8)
4 jq jq p r5
jq jq r p
where r ¼ S11. If the shunt model is acceptable p ¼ 1 – r, and q ¼ [r(1 – r)]1/2. For a
negative offset, q will be replaced by q in (8.8). A resonant shunt-series coupling
slot was used (see Figure 7.12 in Chapter 7) at the centre of a feed waveguide of a
planar array as a transformer to match the total resistance seen by the feed waveguide
into a match at the input port [4]. In that case the normalised input admittance is given
by (8.9) with the coupling coefficient determined by (8.10) using S11 at resonance.
3
Slot
2
(a)
Short
(b)
scattering matrix of the junction with ports 1 and 4 referenced to the centre of the
slot is
" #
1 þ ejy 1 þ ejy
S ¼ 0:5 (8.12)
1 þ ejy 1 þ ejy
For the special case of s ¼ 0, the slot with half the original width will be in
front of the short. The previous result may be shown to be valid for this special case
by considering the image of the equivalent magnetic current of the slot. In this
case the two-port device has zero reflection at ports 1 and 4 and the transmission
coefficient between ports 1 and 4 is unit amplitude and 180 phase. Such a device is
conveniently used in multi-layer power dividers for total transmission from one
layer to another and also to introduce the required amounts of phase shifts through
extra lengths of waveguides between ports 1 and 4 and the slot centre [8].
Figure 8.8 An inclined slot cut in the edge wall of a rectangular waveguide
Concepts and models for advanced designs 183
slots in the broad walls, the scanning range is limited because of the broad wall
dimension being greater than half wavelength in free space.
The edge wall slot may be modelled as a shunt element since it scatters sym-
metrically. Since the edge wall dimension is less than a half wavelength in free
space, a slot cut in that wall may have to be wrapped around both broad walls
in order to accommodate the required resonant slot length. This presents a problem
in stacking waveguides one on top of another. The tilted slots introduce a certain
amount of cross-polarisation. Untilted edge wall slots excited by probes, wires or
irises do not radiate cross-polarisation. Figure 8.10 shows a concept for producing
dual polarisation. One set of waveguides with longitudinal offset slots cut in the
broad walls produce linear polarisation perpendicular to the waveguide axes. The
second set of waveguides with edge wall slots will produce the orthogonal polar-
isation along the waveguide axes. See also Chapter 9.
In the design of edge wall slot arrays, mutual coupling can be incorporated by
using an infinite array model [9,10]. It is possible to compute the mutual coupling
by the so-called element by element method by considering a pair of slots at a time
using the spectrum of two-dimensional solutions [11,12].
Figure 8.9 The aperture electric field in two edge wall slots with a half guide
wavelength spacing
3
δ
θ
δ 1 – Longitudinal
2 – Centred inclined
t
3 – Compound
Figure 8.11 Broad wall radiating slots. IEEE 1989, reprinted from [13],
with permission
Concepts and models for advanced designs 185
(a) (b)
(c) (d)
1
0.5
Normalised offset
Quadrant 2 Quadrant 1
0
–90 –60 –30 0 30 60 90
Quadrant 3 Quadrant 4
–0.5
–1
(e) Tilt in degrees
Figure 8.12 Four ranges of phase for the aperture electric field and possible range
of offsets and tilts for compound slots. (a) d > 0, 0 q 90 , 0
arg E 90 ; (b) d > 0, 90 q 0 , 90 arg E 180 ;
(c) d < 0, 90 q 0, 180 arg E 270 ; (d) d < 0, 0 q
90 , 270 arg E 360 ; (e) possible range of normalised slot
offsets (vertical axis) and tilts (horizontal axis) for a compound slot
Although the compound radiating slot was introduced in the 1940s [7], it was not
used in an antenna application until 1960 [14], when a linear travelling wave array of
compound radiating slots was designed, and built. It is ideally suited for travelling
wave arrays such as the fuze antenna of a missile system in which the beam is at an
angle with respect to the missile axis. Compound slots are also useful as coupling
elements. Although an isolated resonant compound slot can be designed to produce any
value of the aperture amplitude and phase, the amount of realisable phase values will be
limited by the reflection coefficient at the output port, with a greater range of phase
values for smaller reflection coefficients. This result was shown for a compound cou-
pling slot in [15] but it is true for compound radiating slots also. Compound slot arrays
may be designed using the procedure described in Section 6.8 in Chapter 6 in terms of
scattering wave parameters. An array of compound slots all with a tilt of 45 but
different offsets has been used in an automotive collision avoidance radar [16].
186 Slotted waveguide array antennas: theory, analysis and design
I (mm) C (mm)
5.0 4.7
6.4 6.1
7.5 7.5
8.8 9.1
10.0 10.8
Source: 1991, I. Landmark/Nyström. Modified from [21],
with permission.
Table 8.2 Values of I, C and the length of a centred slot to produce resonance at
5.3 GHz compared to an equivalent longitudinal offset slot producing
the same resonant conductance
0.6
0.4
Normalised admittance
0.2
0
G1
G2
–0.2
B1
B2
–0.4
4.8 5 5.2 5.4 5.6 5.8
Frequency in GHz
Figure 8.14 Susceptance (B1) and conductance (G1) versus frequency for an
iris-excited centred slot and the susceptance (B2) and conductance
(G2) of an equivalent conventional offset longitudinal slot for
case A in Table 8.2. 1991, I. Landmark/ Nyström. Modified
from [21] with permission
0.2
0.1
Normalised admittance
–0.1
G1
G2
–0.2
B1
B2
–0.3
4.8 5 5.2 5.4 5.6 5.8
Frequency in GHz
Figure 8.15 Susceptance (B1) and conductance (G1) versus frequency for an
iris-excited centred slot and the susceptance (B2) and conductance
(G2) of an equivalent conventional offset longitudinal slot for
case C in Table 8.2. 1991, I. Landmark/ Nyström. Modified
from [21] with permission
In (8.13) fin is a function of iris dimensions, slot length and width. Similar to the
first design equation (6.10) in Chapter 6 and in [23] we derive
Yna Vns
¼ fin Ki (8.14)
G0 Vn
where
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1 2wm0 b10 ab
Ki ¼ (8.15)
p=a G0
The second design equation (6.15) derived in Chapter 6 is valid for the iris-
excited slot also and it is reproduced as (8.16).
Ya 2
¼
G0 2 1 Vscoupl (8.16)
þ
Ys =G0 G Vs
However, the voltage coupled to the slot from the external mutual coupling,
given in (6.17), gets modified for iris-excited slots as
jp 1 Ys =G0 X N
0
Vscoupl ¼ Vs ext ¼ 2 f 2 Y =G þ 2
Vms gmn ðLm ; Ln ; Xmn ; Zmn Þ
2ðk0 bÞb10 a ðwm0 Þ i s 0
3
m¼1
(8.17)
Using the design equations given here, one can design linear and planar arrays
of iris-excited slots using computed values of resonant lengths and admittances as
functions of normalised slot lengths for different couplets I and C.
190 Slotted waveguide array antennas: theory, analysis and design
(a) (b)
The longitudinal slot cut in the broad wall of a ridge waveguide scatters
symmetrically, just like the longitudinal slot in a rectangular waveguide and may be
modelled in the form of a shunt admittance on an equivalent transmission line.
Using Montgomery’s expressions [27] for the TE10 mode electric field, the first
design equation similar to (6.10) of the longitudinal slot in the conventional
rectangular waveguide is obtained next.
Yna Vs
¼ K1 fn n (8.18)
G0 Vn
where
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2kt kt2
K1 ¼ (8.19)
k0 wm0 b10 G0
Ya 2
¼
G0 2 1 Vscoupl (8.20)
þ
Ys =G0 G Vs
However, the voltage coupled to the slot from the external mutual coupling,
given in (6.17) gets modified for the ridge waveguide slot as [24]
A planar array consisting of two ridge waveguides with eight radiating slots in
each was designed using (8.18) and (8.20), and it was built and tested [24]. The
measured results for the radiation pattern and the voltage standing wave ratio
(VSWR) compared well with the theory.
0.42
2Lres
λ0 0.38
0.34
0.30
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
h/λd
Figure 8.18 The effect of the dielectric cover thickness on the resonant length
of the slot. IEEE 1990, reprinted from [31], with permission
Experiment [8]
1.00 This theory
0.08
Gr
0.06
G0
0.04
0.02
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
h/λd
Figure 8.19 The effect of the dielectric cover thickness of the resonant
conductance of the slot. IEEE 1990, reprinted from [31],
with permission
dielectric cover is used. The resonant slot conductance also decreases in presence
of the dielectric cover (see Figure 8.19). For both of these plots, the slot length is
12.065 mm, and the slot offset is 2.36 mm. The dielectric constant is 3.31. The
waveguide inner dimensions are 22.86 mm 10.16 mm and the slot width is 1.59 mm.
Casula and Montisci have modified the Elliott’s design procedure for a slot array
with a dielectric cover [32]. The modification is needed only in the second design
Concepts and models for advanced designs 193
–5
Far-field pattern [dB]
–10
SLL = –13 dB
SLL = –14.1 dB
–15
–20
–25
–30
–90 –75 –60 –45 –30 –15 0 15 30 45 60 75 90
Angle from broadside [deg.]
Figure 8.20 Simulated radiation pattern in the E-plane. Solid line (with
dielectric); dashed line (without dielectric). IEEE 2009,
reprinted from [32], with permission
equation involving the external mutual coupling. The external mutual coupling is
evaluated in the spectral domain in the form of numerical double integrals in polar
coordinates. The radial integral exhibits highly oscillating terms and has poles due to
surface waves. For a thin dielectric cover, only the TM0 surface wave pole is present.
The residue contribution of the pole has to be considered in addition to its principal
value [33] or the external mutual coupling integral in spectral domain may be deformed
to avoid the pole singularity [30]. The highly oscillating integral is evaluated by the
method of weighted averaging and by the use of Shank’s transforms [34,35].
Casula and Montisci demonstrate the significance of a dielectric cover of
thickness as small as 0.012 free space wavelength and a dielectric constant equal to
2.2 by comparing the radiation pattern and input reflection coefficient of a 10 5
array (five waveguides with ten slots in each) simulated with and without the
dielectric [32]. Those results are reproduced in Figures 8.20 and 8.21.
–5
–10
–15
–20
–25
–30
–35
–40
–45
8.7 8.75 8.8 8.85 8.9 8.95 9 9.05 9.1 9.15 9.2 9.25 9.3
Frequency [GHz]
Slot n + 1 Branch
waveguides
Slot n
Slot n – 1
Feed
waveguide
height waveguides, in addition to the TE20 mode coupling, there may be some
significant TE01 mode coupling as well. The higher-order mode coupling between
adjacent shunt-series coupling slots may be accounted for by using a procedure
similar to that for such coupling between adjacent radiating slots [37]. In this
section we will discuss a method for incorporating the higher-order mode coupling
between centred-inclined coupling slots [38].
Concepts and models for advanced designs 195
where Gn is the reflection coefficient of the slot without accounting for higher-order
mode coupling, Zn is the impedance of the coupling slot in the absence of higher-
s
order mode coupling and Vnh is the slot voltage induced by the higher-order mode
coupling from adjacent slots. Using (6.36) in Chapter 6 and (8.22) and (8.23) we
s
obtain an expression for Vnh as
s
Vn1 V s Ih Gn
s
Vnh ¼ Ih ¼ n1 ¼ Ih (8.26)
I10 A10 Cn K2 Cn2
196 Slotted waveguide array antennas: theory, analysis and design
where Ih is given by the reaction integral (8.27) with Hhz the magnetic field along
s
the axis of slot n due to higher-order modes scattered from adjacent slots. Vn1 is the
slot voltage induced by a TE10 mode wave of amplitude A10.
ð Ln
p
Ih ¼ Hhz cos z dz (8.27)
Ln 2Ln
Zna 1 1
¼ ¼ h s i (8.28)
Z0 Z0 =Zn 2K12 C 2 Ih =Vns Z0 =Zn 1
2 I 1
Vn1
s þ I 2
s
Vnþ1
s
n 2K2 Cn V
n Vn
The dominant contribution to Ih is expected to be from the TE20 mode but it is easy
to account for a few other modes to get an accurate evaluation. Using the dyadic
Green’s function for the magnetic field of the magnetic current given in (2.62) in
Chapter 2 we can determine I1 for the coupling between slots n 1 and n and I2
between slots n þ 1 and n, shown in Figure 8.22 using (8.29). The term I1 is non-
existent for the first coupling slot and likewise I2 is not relevant for the last one.
ð Ln ð Ln1
p p
I1 ¼ cos z cos z0
Ln 2Ln Ln1 2Ln1
G1xx ðx; x0 ; z; z0 Þ sinðqn Þ sinðqn1 Þ
þ G1xz ðx; x0 ; z; z0 Þ sinðqn Þ cosðqn1 Þ
þ G1zx ðx; x0 ; z; z0 Þ cosðqn Þ sinðqn1 Þ
where x ¼ a/2 þ z sin qn, x0 ¼ a/2 þ z0 sin qn1, z ¼ z cos qn, z0 ¼ z0 cos qn–1 – D,
and D is the slot spacing. The tilt angles of slots n and n – 1 are qn, and qn–1,
respectively. The expression for I2 is obtained readily from I1 given in (8.29)
by replacing subscripts n – 1 by n þ 1 and –D by D in z0 in (8.29). The terms
G1xx, G1xz, G1zx and G1zz are the ^x^x ; ^x^z ; ^z^x and ^z^z components respectively of the
dyadic Green function G 1 given in (2.63) and (2.64) in Chapter 2.
Main
guide
θ
1
a
X01 2Lc W
2
X02
Branch
guide
(a)
Main
guide
2
2Lr2
2Lr1
1
Branch
guide
(b)
Figure 8.23 Geometry of the coupling junction: (a) hard coupling; (b) soft
coupling. IEEE 1994, reprinted from [5], with permission
are primarily confined to the coupling and radiating slots in the junction regions. If
the vast majority of the radiating slots elsewhere have close to the desired excita-
tions, the radiation pattern is expected to be of acceptable accuracy in the H-plane.
However, the errors in the coupling slots may affect the input match and affect the
pattern in the E-plane. In a small array such errors may become substantial since
the percentage of total number of radiating slots in the coupling junction region is
greater.
There is a simple method to include the higher-order mode coupling discussed
earlier in the design of planar arrays. Let us consider the problem of three slots in
the junction region using an MoM solution to the coupled integral equations of the
aperture fields of slots. The calculation can be made for two sets of excitations,
with and without higher-order mode coupling in the junction region. Then one can
perturb the coupling slot’s tilt and length and the offsets and lengths of the radiating
slots, such that the excitations of the perturbed slots with higher-order mode
Concepts and models for advanced designs 199
coupling are the same as the excitations of the original slot parameters without
higher-order mode coupling. We then replace the original slots in Elliott’s design
by the perturbed slots. The rationale for this design is that higher-order mode
coupling is a localised phenomenon and therefore we can account for it by
considering only the three slots in the junction region, without accounting for
the presence of other slots. In some cases, however, this method does not produce
satisfactory results and one may need to use a full wave method to analyse the
entire slot array to account for the coupling effects. Such a technique is discussed in
Section 8.10 [39].
n' ρ
γ s
n ρ
s1
ρ1
an additional term for the magnetic field in the region of slot n due to a unit
magnetic current in slot n, due to edge diffraction, derived later.
The magnetic field incident at the edge is given by
The parameter n is given by the interior wedge angle (2 – n)p. For a ground
plane of zero thickness, n ¼ 2 whereas for a thick edge, the interior wedge angle is
90 and n ¼ 0.5. j and j0 are the angles that the incident and diffracted rays make
with respect to the plane surface, and they are zero in this case. Because of grazing
incidence, a factor ½ has been included in the diffracted field since the total inci-
dent field at grazing incidence is the sum of an incident and a reflected field [43].
Thus we obtain the diffracted field as
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
exp ½ jk ð s þ s1 Þ ð 1 j Þ ðs þ s1 Þ
H d ¼ H0 pffiffiffiffiffiffiffiffi (8.33)
ðs þ s1 Þ 16p krr1
The diffracted field at slot n can be written as the incident field at the image
slot n0 multiplied by the edge diffraction parameters. Therefore the original mutual
coupling term gmn discussed in Section 6.4 in Chapter 6 is augmented by adding an
extra term containing the integral between slots m and n0 (image of n) as follows.
gmn ! gmn þ H0 gmn0 (8.34)
where
rffiffiffiffiffiffiffiffiffiffiffiffi
1 þ j s þ s1
H0 ¼ pffiffiffiffiffiffiffiffi (8.35)
16p krr1
The field of slot n, incident at the edge, after getting diffracted may couple to
itself. This term can be added to the summation in (6.46b) in Chapter 6 as
Vns gnn0 ðLn ; Ln Þ (8.36)
The gnn0 integral will be carried out using the procedure discussed in Section 6.4
in Chapter 6 for slots n and n0 of half-length Ln, separated by a distance 2r in the
E-plane direction and aligned in the H-plane direction, as shown in Figure 8.24.
Concepts and models for advanced designs 201
Section 5.2 in Chapter 5 discusses the MoM solution to the integral equation of the
aperture field of a single slot. This is the most efficient technique to compute the
characteristics of a slot since the unknowns are the equivalent magnetic currents in
the slot apertures, as opposed to techniques such as the finite element method and
finite difference method where the unknowns are in a volume. In addition, when
one uses entire domain sinusoidal basis functions for the longitudinal variations of
the magnetic current, the first term, a half cosinusoidal distribution, is an excellent
approximation to the aperture electric field of a slot near resonance. Therefore a
few unknowns are usually adequate to obtain good accuracy.
The slot thickness is accounted for by considering the two (inner and outer)
apertures of the slot. The formulation of the coupled integral equations starts with
shorting out each slot aperture by an infinitesimally thin perfectly conducting sheet.
By invoking Schelkunoff’s equivalence principle (see Section 2.7), we place
equivalent magnetic currents on each such conducting sheet. Enforcement of the
continuity of the tangential electric fields across each aperture of the original pro-
blem requires that the magnetic current on either side of the conducting sheet shall
be equal in magnitude but opposite in direction. Figure 8.25 shows the cross section
of a thick radiating slot which we assume is the mth slot.
By enforcing the continuity of the longitudinal component of the aperture
magnetic field for the mth slot we obtain (8.39) and (8.40).
X
extþ ~ ext ext ~ ext ext ~ int extþ ~ int
Hzm M m Hzm M m Hzm Mm þ Hzm Mq ¼ 0 (8.39)
q
X
intþ ~ int intþ ~ ext int ~ int int ~ int
Hzm M m þ Hzm M m Hzm Mm Hzm M n ¼ Hzm
inc
(8.40)
n6¼m
Thus, there are two coupled integral equations for slot m, one for the exterior
aperture facing the exterior half space and the other for the interior aperture facing
the waveguide. There are two unknowns for slot m, M ~ ext and M~ int , the magnetic
m m
currents in the exterior (ext) and interior (int) aperture, respectively. The slot m is
coupled to all the radiating slots in the entire array through external mutual cou-
pling in the summation over q. Similarly the slot m is coupled to all the other
radiating slots and the coupling slot in that radiating waveguide, in the summation
over n, through TE10 mode scattering. In addition, slot m is coupled to adjacent
radiating slots and to the coupling slot, if it is in the immediate vicinity, through
higher-order modes. The magnetic field of the magnetic current in each of the three
regions (waveguide interior, exterior half space and the cavity region to account for
the slot thickness) is expressed in terms of integrals over the magnetic current
with the appropriate Green’s function for the waveguide, half space and cavity,
respectively.
For a linear array of standing wave or travelling wave type directly fed by an
incident TE10 mode wave, an additional contribution on the right side of (8.40)
exists. This source term is absent in planar standing wave arrays where all the
ext
+
–
+
–
Thick wall int
Figure 8.25 The internal and external apertures of the mth radiating slot
Concepts and models for advanced designs 203
radiating waveguides are supposed to be shorted at both ends and excited by cou-
pling slots. A similar pair of coupled integral equations results when the continuity
of the longitudinal component of the magnetic field is enforced across the apertures
of each coupling slot. The incident field term is absent if the feed waveguide
containing the coupling slots is shorted at both ends and excited by an input
slot coupling to another waveguide layer. An example of such an input shunt-series
coupling slot is found in Figure 7.12. The source term then appears in the integral
equation for this slot.
Short Short
m C10 n
Figure 8.26 Slots cut in a waveguide wall with shorts at both ends
204 Slotted waveguide array antennas: theory, analysis and design
known. The radiation pattern may be computed by summing the radiated far fields
of the magnetic currents of all slots. The reflection coefficient at the input port is
found from the scattering from the coupling slots. Directivity is readily found since
the incident wave power and reflected power are known and the power density in
the main beam peak direction can easily be calculated.
Therefore all coupling slots need to be detuned so that the input reflection coeffi-
cient will be equal to that given by (8.44).
G0 ¼ r þ s expðjaÞ (8.44)
Z ¼ ð1 þ G0 Þ=ð1 G0 Þ (8.45)
In the initial design using the Elliott’s procedure, the total normalised slot
conductance of all radiating slots in each radiating waveguide was assumed to be 2
with an input match at the feed port enforced. Figure 8.28 shows the E-plane
pattern of this array with the legend ‘original’. A substantial amount of higher-order
mode coupling is present in the slots in the five coupling junctions, thereby affecting
the return loss and the E-plane pattern. It was concluded that the major source of error
in the original design was the higher-order mode coupling in the five coupling
junctions.
Figure 8.28 shows also the E-plane patterns computed with an approximate
method of accounting for the higher-order mode coupling in the junction slots in
the original design. In this method for which results are shown by the legend
‘approximate’, one perturbs each coupling slot and a pair of straddling radiating
slots by considering one junction at a time. The perturbation compensates for the
errors introduced by the higher-order mode coupling in the original design. This is
a major simplification considering only six coupled integral equations of the three
slots for each junction. However, the method is approximate and, as shown by
Figure 8.28, the improvement produced by this method is marginal. Obviously the
iterative method of perturbing all slots using the moment method analysis of the
entire array antenna, shown by the legend ‘improved’, produces very good results.
Concepts and models for advanced designs 207
20
10 Improved
Approximate
Pattern level in dB 0 Original
–10
–20
–30
–40
–90 –60 –30 0 30 60 90
Theta in degrees
–5
–10
|S11| in dB
–15
–20
–25
–30 Original
Improved
–35
13 13.1 13.2 13.3 13.4 13.5 13.6
Frequency in GHz
Figure 8.29 Magnitude of the reflection coefficient of the original and the
improved designs. Reprinted from [39], with permission
For arrays made up of sub-arrays, the proposed method can be applied by con-
sidering one sub-array at a time while treating all other sub-arrays as additional
source terms or external excitations.
The improved design procedure using MOM applied to this antenna yielded
radiation patterns close to the ideal patterns. Since the higher-order mode coupling
in the junctions did not have a significant effect in the H-plane patterns, they are not
shown here. Figure 8.29 shows the reflection coefficient in dB over a range of
frequencies. The original design has the best match at 1% below the centre of the
desired frequency band. The improved design places the best performance at the
208 Slotted waveguide array antennas: theory, analysis and design
20.3
20.2
20.1
Directivity in dB
20
19.9
19.8
Original
19.7
Improved
19.6
13 13.1 13.2 13.3 13.4 13.5 13.6
Frequency in GHz
Figure 8.30 Directivity of the original and the improved designs. Reprinted
from [39], with permission
Table 8.4 Coupling slot parameters in the original and improved designs
centre frequency and achieves the desired 10 dB return loss over 4% frequency
band. Similarly, the improved design achieves between 0.1 and 0.235 dB increase
in the values of directivity within the required 4% frequency band compared to
the original design. Figure 8.30 compares the directivity of the two designs.
E-plane patterns of the improved design at band edges, not shown here, were closer
to the desired patterns.
Table 8.4 shows the lengths and tilt angles with respect to the feed waveguide
axis of the coupling slots for the original and improved designs. Since coupling
slots exhibit symmetry with respect to the centre only the three elements closest to
the feed port are specified. The perturbations are found to be small. Table 8.5
shows similar results for the offsets and lengths of radiating slots in the three
radiating waveguides starting from the one nearest to the feed port. In the original
design radiating slot parameters exhibit symmetry with respect to the diagonal, but
for the sign of the offset. In the improved design such symmetry is present in an
approximate sense. Radiating slot perturbations are significant compared to those
of coupling slots.
Concepts and models for advanced designs 209
Table 8.5 Radiating slot parameters in the original and improved designs
mode internal higher-order mode coupling terms. The slot voltages in the left
column matrix are given by
k2
½G ½V s ¼ V1 ½ E (8.48)
k1
where
Gmn ¼ gmn if n
m (8.49)
Gmn ¼ gmn jk2 fm fn sinfðm nÞb10 d g if n < m (8.50)
of 4.6%. When the design was optimised for 5% bandwidth, the SLL was better than
21.9 dB, while Pload/Pin was below 6.2% within the band. The VSWR was better than
1.02 in the frequency range of interest. For 10% bandwidth optimisation, SLL was
better than –20.9 dB and Pload/Pin was lower than 7.1%. VSWR values are higher than
those for the 5% bandwidth case but still below 1.04 in the entire band.
The optimisation exercises show the power of the GA when combined with an
antenna analysis technique. However, this theoretical exercise did not include the
internal higher-order mode coupling and has the limitations of the Elliott’s design
procedure discussed in Chapter 6.
Width
Width
d2 d1
Output port Input port
Figure 8.32 A waveguide H-plane section with two thick inductive irises
212 Slotted waveguide array antennas: theory, analysis and design
four parameters, widths of the two irises in the H-plane, the spacing between the
irises, d2, and the distance between the input port and the nearest iris, d1, that may be
optimised to maximise the return loss over a frequency band. Iris thickness is usually
fixed. The input port is connected to the slot array to be matched.
Mode matching (MM) is a fast and accurate technique for solving boundary
value problems for waveguide discontinuities [55]. It is ideally suited to the analysis
of waveguides with thick irises. GA optimisation in conjunction with MM may be
used to improve the impedance match over the required frequency bandwidth.
We start with the design of an 8 10 array consisting of eight radiating
waveguides, each containing ten radiating longitudinal slots, using Elliott’s pro-
cedure. The aperture distribution is assumed to be uniform. The dimensions of the
input and radiating waveguides are 5.70 mm 2.85 mm while those of the feed
waveguide are 5.673 mm 2.836 mm. The slot widths and thicknesses are
0.508 mm and the centre frequency is 35.75 GHz. Diagonal symmetry is assumed
in the array for the radiating slots whereas centred-inclined coupling slots are
assumed to be symmetric with respect to the centre but for alternating signs for tilt
angles. In the GA design there were 85 parameters, offset and length for each of the
40 radiating slots, tilt angle for each of the four centred-inclined coupling slots and
offset from the centreline for the input shunt-series coupling slot. Other radiating
and coupling slot data are obtained from symmetry. Coupling slots are all assumed
to be resonant. Binary GA with 7 bits for each parameter is employed so that the
length of each chromosome is 595. This allows a resolution in the variable size to
be less than the machine tolerance, assumed to be 1 mil (~25 m).
For each element of the population (chromosome) GA performs the moment
method analysis of the slot array at nine equally spaced frequency points within a
4% frequency band centred at 35.75 GHz. The lowest values of the broadside gain
and the return loss are weighted to specify the fitness parameter in GA. Thus we
optimise the worst case values of gain and return loss within the bandwidth for a
nominally uniform aperture distribution. In the GA the fitness parameter is not
allowed to increase if the return loss is greater than 12 dB. A population size of 10,
tournament selection, mutation probability of 0.1, and probability of crossover of 0.5
with uniform crossover are used with the best individual replicated into the next
generation [46–48]. Optimum solutions are achieved typically after 600 generations.
The range of allowed values for the offset from the centreline for radiating slots
and the input shunt-series coupling slot is about 1/10th of the waveguide ‘a’ dimen-
sion. The slot length and the distance between the short and the nearest slot centre are
allowed to have a range of about 5% of their nominal values while the tilt angle of
each centred-inclined coupling slot is allowed to vary 5 with respect to the initial
value. The computer time for each GA execution for about 600 generations is about
one week in a dedicated PC with an Intel Core Duo processor T9600 and a clock
speed of 2.8 GHz with 4GB of RAM. The computer time may be drastically reduced
by using many parallel processors or using resources such as cloud computing.
In the case of matching sections with irises, the range of values for the spacing
between irises and the spacing between the array feed port and the nearest iris is a
maximum of half a guide wavelength at the centre frequency. The irises are
Concepts and models for advanced designs 213
assumed to be 20 mil (~0.5 mm) thick and their widths are allowed to have a range
of 1/3 of the waveguide ‘a’ dimension. There are four variables and each is 10 bits
long so that the length of the chromosome is 40. Other parameters of GA in
this case are identical to the previous GA MOM optimisation. The fitness
parameter is the lowest value of the return loss computed at nine equally spaced
frequency points within 4% bandwidth centred at 35.75 GHz. GA mode matching
(MM) is an extremely fast computational process.
Figure 8.33 shows gain values for three cases. The original design using Elliott’s
procedure has a gain plot with a wide variation in the frequency range. A GA solution
obtained with one basis function for each slot aperture has nearly a constant gain over
a 4% bandwidth. For each radiating slot and the input shunt-series coupling slot in
this GA optimisation, we obtain an equivalent slot with nine basis functions while for
each coupling slot the equivalent slot with three basis functions is obtained from
Equation (8.52). The gain values calculated for an array with these equivalent slots
is shown using the legend ‘nine bases GA’. The latter two plots are very similar.
Figure 8.34 shows results for the reflection coefficients for the three cases. These
results further justify the equivalence established by (8.52).
27
26.5
Gain in dB
26 Orig design
1 basis GA
9 bases GA
25.5
34.5 35 35.5 36 36.5 37
Frequency in GHz
Figure 8.33 Gain plots of GA designs with one basis function and an equivalent
array with nine basis functions. The legend ‘orig design’ used
Elliott’s procedure. Reprinted from [54], with permission
–5
–10
Reflection coefficient in dB
–15
–20
–25
–30 Orig design
1 basis GA
–35 9 bases GA
–40 –10 dB ref coef
–45
34.5 35 35.5 36 36.5 37
Frequency in GHz
Figure 8.34 Reflection coefficient plots of GA designs with one basis function
and an equivalent array with nine basis functions. The legend
‘orig design’ used Elliott’s procedure. Reprinted from [54],
with permission
Concepts and models for advanced designs 215
27
26.5
Gain in dB
26
Orig design
25.5 GA
25
34.5 35 35.5 36 36.5 37
Frequency in GHz
–5
Reflection coefficient in dB
–10
–15
–20
Orig design
–25
GA
–30
–10 dB ref coef
–35
34.5 35 35.5 36 36.5 37
Frequency in GHz
30
Orig design
20
GA
Pattern level in dB
10
–10
–20
–30
–90 –60 –30 0 30 60 90
(a) Theta in degrees
30
Orig design
20 GA
Pattern level in dB
10
–10
–20
–30
–90 –60 –30 0 30 60 90
(b) Theta in degrees
Figure 8.37 Radiation patterns at 35.75 GHz; ‘orig design’ used Elliott’s
procedure and GA optimised the gain and the reflection
coefficient: (a) H-plane pattern; (b) E-plane pattern.
Reprinted from [54], with permission
within the 5% frequency band. The initial GA optimisation that produced no higher
than –12 dB reflection coefficient is further optimised by GA with a separate design
of two irises. Thus Figure 8.36 shows a nearly flat reflection coefficient of –15 dB
and exhibits greater than 6% bandwidth for input match with better than –10 dB
reflection coefficient.
Typical radiation patterns computed from the GA optimised design are shown in
Figure 8.37 for the centre frequency. The pattern peaks are normalised to 0 dB for easy
comparison of sidelobe levels. There is no cross-polarised radiation in the principal
planes. GA patterns generally exhibit lower levels of sidelobe radiation. The GA
optimised the gain over a 5% bandwidth. The pattern with the legend ‘orig design’ was
obtained from the array designed by the Elliott procedure. The GA design produced
lower sidelobes compared to the original design using the Elliott procedure.
Concepts and models for advanced designs 217
References
[1] Rengarajan S. R. ‘Analysis of a centered-inclined waveguide slot coupler’.
IEEE Transactions on Microwave Theory and Techniques. 1989;37(5):
884–889.
[2] Mazzarella G., Montisci G. ‘Wideband equivalent circuit of a centered-
inclined waveguide slot coupler’. Journal of Electromagnetics and Appli-
cations. 2000;14:133–151.
[3] Rengarajan S. R. ‘Characteristics of a longitudinal/transverse coupling slot
in crossed rectangular waveguides’. IEEE Transactions on Microwave
Theory and Techniques. 1989;37(8):1171–1177.
[4] Rengarajan S. R., Zawadzki M. S., Hodges R. E. ‘Waveguide-slot array
antenna designs for low-average-sidelobe specifications’. IEEE Antennas
and Propagation Magazine. 2010;52(6):89–98.
[5] Rengarajan S. R., Shaw G. M. ‘Accurate characterization of coupling junc-
tions in waveguide-fed planar slot arrays’. IEEE Transactions on Microwave
Theory and Techniques. 1994;42(12):2239–2248.
[6] Senior D. C. Higher-order mode coupling effects in a shunt-series coupling
junction of a planar slot array. Ph.D. dissertation, University of California,
Los Angeles, CA, 1986.
[7] Watson W. H. ‘Resonant slots’. IEE Journal. 1946;93(Part 3A):747–777.
[8] Zawadzki M. S., Rengarajan S. R., Hodges R. E. ‘The design of H- and
V-pol waveguide slot array feeds for a scanned offset dual-polarized
reflectarray’. IEEE Antennas and Propagation Society International
Symposium. 2005;2B:417–420.
[9] Silver S. Microwave Antenna Theory and Design. MIT Radiation Laboratory
Series, vol. 12, McGraw-Hill, New York, 1949.
[10] Yee H. Y. ‘Slot antenna arrays’ in Johnson R. C., Jasik H. (eds.). Antenna
Engineering Handbook. McGraw-Hill, New York, 1984.
[11] Kildal P.-S., Rengarajan S. R., Moldsvor A. ‘Analysis of nearly cylindrical
antennas and scattering problems using a spectrum of two-dimensional
solutions’. IEEE Transactions on Antennas and Propagation. 1996;44(8):
1183–1192.
[12] Rengarajan S. R. ‘Mutual coupling between slots cut in rectangular
cylindrical structures: spectral domain technique’. Radio Science. 1996;31(6):
1651–1661.
[13] Rengarajan S. R. ‘Compound radiating slots in a broad wall of a rectangular
waveguide’. IEEE Transactions on Antennas and Propagation. 1989;37(9):
1116–1123.
[14] Maxum B. J. ‘Resonant slots with independent control of amplitude and
phase’. IRE Transactions on Antennas and Propagation. 1960;8(4):384–389.
[15] Derneryd A., Petersson R. ‘Bandwidth characteristics of monopulse slotted
waveguide antennas’. Proceedings of the Fourth International Conference
on Antennas and Propagation, 1985, pp. 27–30.
218 Slotted waveguide array antennas: theory, analysis and design
[16] Montesinos-Ortego I., Zhang M., Sierra-Perez M., Hirokawa J., Ando M.
‘Systematic design methodology for one-dimensional compound slot-arrays
combining method of moments, equivalent circuit model and forward
matching procedure’. IEEE Transactions on Antennas Propagation. 2013;
61(1):453–458.
[17] Gruenberg H. ‘Second order beams of slotted waveguide arrays’. Canadian
Journal of Physics. 1953;31(1):55–59.
[18] Gruenberg H. ‘Theory of waveguide-fed slots radiating into parallel plate
regions’. Journal of Applied Physics. 1952;23(7):733–737.
[19] Forooraghi K., Kildal P.-S., Rengarajan S. R. ‘Admittance of an isolated
waveguide-fed slot radiating between baffles using a spectrum of two-
dimensional solutions’. IEEE Transactions on Antennas and Propagation.
1993;41(4):422–428.
[20] Tang R. ‘A slot with variable coupling and its application to a linear array’.
IRE Transactions on Antennas and Propagation. 1960;8(1):97–101.
[21] Nystrom I. Excitering av centrerad slits i vågledare. Report R/AG-91:188,
Ericsson Radar Electronics AB, Sweden, 1991.
[22] Marcuvitz N. Waveguide Handbook, MIT Radiation Laboratory Series,
Vol. 10, Boston Technical Publishers, Inc., 1964.
[23] Elliott R. S. ‘An improved design procedure for small arrays of shunt slots’.
IEEE Transactions on Antennas and Propagation. 1983;31(1):48–53.
[24] Kim D., Elliott R. S. ‘A design procedure for slot arrays fed by a ridge
waveguide’. IEEE Transactions on Antennas and Propagation. 1988;36(11):
1531–1536.
[25] Falk K. ‘Conductance of a longitudinal resonant slot in a ridge waveguide’.
IEE Proceedings, part H. 1987;134(1):98–100.
[26] Falk K. ‘Admittance of a longitudinal slot in a ridge waveguide’. IEE
Proceedings, part H. 1988;135(4):263–268.
[27] Montgomery J. P. ‘On the complete eigenvalue solution of ridge waveguide’.
IEEE Transactions on Microwave Theory and Techniques. 1971;19(6):547–555.
[28] Bailey M. C. ‘Design of dielectric-covered rectangular resonant slots in a
rectangular waveguide’. IEEE Transactions on Antennas and Propagation.
1967;15(5):594–598.
[29] Bailey M. C. ‘The impedance properties of dielectric-covered narrow
radiating slots in the broad face of a rectangular waveguide’. IEEE Trans-
actions on Antennas and Propagation. 1970;18(5):596–603.
[30] Rexberg L. ‘Vector Fourier transform analysis of dielectric-covered slot in
the broad wall of a waveguide’. Microwave and Optical Technology Letters.
1988;1(10):360–363.
[31] Katehi P. B. ‘Dielectric-covered waveguide longitudinal slots with finite
wall thickness’. IEEE Transactions on Antennas and Propagation. 1990;
38(7):1039–1045.
[32] Casula G. A., Montisci G. ‘Design of dielectric-covered planar arrays of
longitudinal slots’. IEEE Antennas and Wireless Propagation Letters.
2009;8:752–755.
Concepts and models for advanced designs 219
[50] Schaffner J., Kim D., Elliott R. S. ‘Compromises among optimum sum and
difference patterns for planar waveguide-fed slot arrays’. Alta Frequenza.
1981;L(6):312–319.
[51] Jensen A., Rengarajan S. R. ‘Genetic algorithm optimization of a traveling
wave array of longitudinal slots in a rectangular waveguide’. Applied
Computational Electromagnetics Society Journal. 2006;21(3):337–341.
[52] Elliott R. S. ‘On the design of traveling-wave-fed longitudinal shunt
slot arrays’. IEEE Transactions on Antennas and Propagation. 1979;27(5):
717–720.
[53] Hamadallah M. ‘Frequency limitations on broad-band performance of shunt
slot arrays’. IEEE Transactions on Antennas and Propagation. 1989;37(7):
817–823.
[54] Rengarajan S. R. ‘Genetic algorithm optimization of a planar array using full
wave method of moments analysis’. International Journal of Computer
Aided RF and Microwave Engineering. 2013;23(4):430–436.
[55] Wexler A. ‘Solution of waveguide discontinuities by modal analysis’. IEEE
Transactions on Microwave Theory and Techniques. 1967;17(9):508–517.
[56] Ando M., Zhang M., Lee J., Hirokawa J. ‘Design and fabrication of millimeter
wave slotted waveguide arrays’. Proceedings of the European Conference on
Antennas and Propagation, 2010, pp. 1–6.
Chapter 9
Antenna systems and special requirements
So far we have mainly discussed flat slotted waveguide array antennas consisting of
rectangular waveguides, typically with slots in the broad wall, radiating a fixed
narrow beam. Many other configurations are possible, however, and even more
suitable in special cases.
In this chapter we will investigate how slotted waveguide arrays are used in
system applications where more advanced antenna functions or special configura-
tions are required. As this is a very broad subject we will concentrate on cases
where advantages and perhaps difficulties are of particular interest.
Waveguide elements
Phase shifters
Feed network
Figure 9.1 Phase steered array with slotted waveguides as radiating elements
–1
–2
d/l = 0.7 0.6 0.5
dB
–3
–4
–5
0 10 20 30 40 50 60 70
Degrees
Figure 9.2 Embedded element patterns for the centre element of a 15-element
array with different element spacings d/l
edge slotted waveguides [10–12]. Figure 9.3 shows a typical size reduction using
ridge-loaded waveguides. The width of the ridge guide is in this example 35% less
than the width of the corresponding rectangular guide for the same guide wavelength.
Shnitkin [9] describes an interesting ridge waveguide configuration (see
Figure 9.4), where the waveguide heights alternate on the two sides of the ridge.
The longitudinal slots remain on the physical centreline, and no ‘butterfly lobes’
(see Section 9.9) are generated.
Antenna systems and special requirements 223
A B
SEC AA SEC BB l l
High Low
Low High
A B
Electrical
neutral plane
(zero H-field)
Figure 9.4 Asymmetric ridge waveguide with slots along the centreline.
1990 IEEE, reprinted from [9], with permission
Numerical methods have been developed for the analysis of ridge waveguide
characteristics; see, for example, Conciauro et al. [13]. Also slots in the ridge
waveguide can be analysed [14,15]. Slot data for array design can of course also
be obtained from measurements. Commercial codes using finite elements (e.g. HFSS)
are used as well.
Systems with phase steering in elevation combined with mechanical rotation in
azimuth is a common way of covering all coordinates, for example, in a 3D radar.
The antenna may be composed of several slotted ridge waveguides with individual
phase control; see Figure 11.7 in Chapter 11. Another interesting solution is to
combine phase steering with frequency scanning [16,17].
A large S-band radar with wide angle phase steering in azimuth is used in an
airborne early warning (AEW) system [18,19]. About 200 transmit/receive modules
feed two slot arrays on each side of the unit carried by the aircraft. The vertical
slotted ridge waveguides allow almost 360 of azimuth coverage (Figure 9.5).
(a)
Air inlet
RF network
Radome
Slotted waveguides
TR modules
Slotted
(b) waveguides
Figure 9.5 AEW radar with a phase steered slotted ridge waveguide
array antenna. (a) The dorsal unit in a near field test chamber.
(b) AEW demonstrator cross section showing TR modules in centre
with slotted waveguide array panels on each side of the dorsal unit.
Courtesy of Saab AB
Antenna systems and special requirements 225
between the radiating slots. Thus, the beam direction can be steered by changing
the frequency. This type of electronic beam steering is fairly simple and econom-
ical; it requires no electronic phase shifters, and the antenna can be fabricated with
relative ease. Several linear arrays can of course be combined to form a planar
array. Frequency scanned slotted waveguide arrays have been developed for fre-
quencies as high as at least 180 GHz [20].
The general problem of frequency scanning has been treated by many authors
[21–24]. There are several basic types of frequency scanned linear arrays; the most
important are
1. The travelling wave array
with matched termination and no phase reversal
(no alternating slot offsets or slot inclination angles)
2. The travelling wave array
with matched termination and phase reversal
3. The resonant series-fed array
short instead of a matched termination
f
Slots θs
Load
d
The beam direction from the normal qs is related to the phase delay:
2p
Df ¼ d sin qs (9.3)
l
yielding the beam direction
sin qs ¼ l=lg (9.4)
Thus, the beam direction is independent of the spacing d, but it depends on the
frequency. In a typical case with lg ¼ l.5l we obtain qs ¼ 42 . The antenna is called
a leaky wave antenna or fast wave antenna. It is interesting to note that the beam
has the same angle versus the y-direction as the constituent TE10 plane waves
versus the x-direction (the waveguide mode can be seen as plane waves inside the
waveguide bouncing between the waveguide side walls).
hence depending on both the frequency and the spacing d. Differentiating the last
expression gives the approximate scan versus frequency sensitivity:
Df
½Dqdegrees (9.7)
f0 %
that is, about 1 scan for 1% frequency variation. The closer to waveguide cut-
off the larger is the beam squint. An example is shown in Figure 9.7 for 5%
frequency scan yielding about 5 of beam scanning. It is here assumed that the
beam shall be directed normal to the array at the frequency f ¼ f0.
The useful angular regions for this type of array are indicated in Figure 9.8. It
scans backwards (towards the feed input) when the slot spacing is less than lg/2
and forward when the spacing is larger than lg/2. See also Figure 6.15. Three
‘forbidden’ regions are indicated: (1) slot spacing less than the slot length is
assumed not possible, (2) slot spacing d ¼ lg/2 implies that the reflections from
all slots add in phase with a resulting high VSWR at the input and (3) too large
Antenna systems and special requirements 227
–2
–4
–6
–8
0.94 0.96 0.98 1 1.02 1.04 1.06
f/f0
Figure 9.7 Frequency scanning of a travelling wave array antenna with phase
reversal. Standard waveguide with a dimension ¼ 22.86 mm. Centre
frequencies are: solid line 7.5 GHz, dash-dotted 8.5 GHz and dashed
line 9.5 GHz
Poor VSWR
spacing causes a grating lobe to radiate. To prevent grating lobe radiation the
requirement is
1:5
d=l < (9.8)
1 þ l=lg
228 Slotted waveguide array antennas: theory, analysis and design
d'
d'
d
d
Figure 9.9 A serpentine feed. Note the increased path d0 compared to the slot
spacing d. 1981 J. H. Elliott. Adapted from [24], with permission
An example of a phased array antenna system using the phase reversal slotted
travelling wave arrays as the phase steered elements was illustrated in Chapter 3
(Figure 3.5). In this case two-dimensional electronic scan was obtained: frequency
scan in elevation and phase scan in azimuth. If the phase delay between slots is
increased a larger scan range is obtained for the same frequency variation. This can
be achieved with a folded, serpentine feed (Figure 9.9). In the figure the serpentine
waveguide has radiating edge slots. It could also feed a second set of slotted
waveguides to form a planar array; an example is shown in Figure 9.10. More than
90 of scan can be obtained with practical serpentines for 10% frequency variation.
With a serpentine (sinuous) feed of length d0 the beam direction can be written
d0l 1 1
sin qs ¼ (9.9)
d lg lg0
where it is assumed that the beam direction qs ¼ 0 when lg ¼ lg0. From (9.9) it
appears that sin qs might exceed unity for large phase delays. However, phase is
ambiguous and we need to pick the right branch, that is inside [0,2p] or [p, þp].
Compare the real versus imaginary space for array patterns in Figure 9.11. The
necessary modification to (9.9) for any beam direction is to include an integer
parameter m as shown in (9.10):
l d0 1
sin qs ¼ m (9.10)
d lg 2
The scanning range with serpentine lengths of 0.5, 3 and 5 guide wavelengths
is illustrated in Figure 9.12.
Antenna systems and special requirements 229
Figure 9.10 The ITT Gilfillan AN/SPS-48 3D radar antenna. Note serpentine
on the left. Photo courtesy of Troy Prince [25]
l/d l/d
0
dB
Visible space
–10
–20
–30
–3 –2 –1 0 1 2 3
sin q
sin qs
45
30
15
Scan angle [deg]
–15
–30
–45
0.94 0.96 0.98 1 1.02 1.04 1.06
f/f0
θs
Short
d circuit
∆θ
The array will produce one single beam only in a narrow frequency range
where the two beams overlap satisfactorily. Assuming a maximum beam separation
of half the three dB beamwidth we get a useful bandwidth BW of about
50
BW ¼ % (9.11)
N
where N is the number of slots. The number 50 is approximately valid for no taper
and will be slightly larger in an amplitude tapered array. We find that the band-
width in per cent approximately equals the beam width in degrees.
We remarked before in relation to Figure 9.7 that scanning in the broadside
direction would cause prohibitively large reflections, since all slot reflections add
up in this case. Especially in long waveguides this could be problematic. However,
certain slot configurations have been found to reduce the reflections considerably.
In one case an inductive post adjacent to the slot is used to tune out the slot
reflection [26]. Another possibility is to use pairs of transverse slots (closely
spaced) that both radiate but the respective reflections are out of phase [27,28].
A variation still is the offset crossed slot radiating circular polarisation which
scatters only in the forward direction [29,30].
f f
Figure 9.14 Principle for the squintless antenna. DSTO Australia (1990),
reprinted from Radarcon 90 [31], with permission
232 Slotted waveguide array antennas: theory, analysis and design
When the frequency is increased both beams scan towards their respective
waveguide loads with the result that the composite beam remains relatively
stable in the far field. The bandwidth realised before the beam splits depends on the
array length, but can in a practical case be around 3–4%.
This antenna has several interesting features from a theoretical viewpoint.
Since the slot spacing is different in the two halves the usual design methods are not
directly applicable. The problem has been addressed by O’Loughlin et al. [32]. See
also Figure 11.7 (Chapter 11).
Septum slot
E
Transverse
E currents
Figure 9.15 The bifurcated waveguide. With even excitation (shown) vertical
polarisation is created. Odd excitation creates horizontal
polarisation. IEEE, adapted from [38], with permission
a1
y h
h1
z
2L
x
a
(a) (b)
Odd and even excitations of the dual waveguide provide the means for con-
trolling the polarisation. The waveguide can be low height so that arraying in a
phase scanned antenna is possible. Related schemes, also based on bifurcated
waveguides, have been proposed by Sangster et al. [39]. It remains, however, to
find a solution to the grating lobe problem caused by the lg spacing of the septum
slots for in-phase excitation.
A combination of two separate waveguides with respectively longitudinal slots
and transverse slots could provide the two orthogonal polarisation components
required for polarisation agility. In order to accommodate the waveguide pair in
an array with sufficiently dense spacing several solutions have been proposed.
Figure 9.16 shows one breadboard C-band array studied for a polarimetric SAR [40].
The building block is shown to the right in Figure 9.16. Both polarisations
excite the region between the vertical baffles that are formed by the waveguides
standing on the narrow side. Inside the baffle region the fundamental propagating
234 Slotted waveguide array antennas: theory, analysis and design
dB
0
–10
–20
θ
–80° –40° 0° 40° 80°
Figure 9.17 Measured patterns for a six-element array with transverse slots
spaced 1.4 l0. Broken curve without baffle, and solid curve with
baffle. IEEE 1991, reprinted from [41], with permission
modes are a TEM parallel plate mode excited by the longitudinal slots and a TE(1)
parallel plate mode excited by the transverse slots. The latter mode has a wavelength
l1 larger than the wavelength lg of the feeding waveguide (since a1 < a, Figure 9.16).
The slots are resonantly spaced at lg distance. Hence the grating lobes inside the
baffle region are evanescent and will be attenuated provided that a > a1 > l0/2, with
l0 the free space wavelength. In a typical case a baffle height of one-half wavelength
gives about 20 dB grating lobe attenuation; see Figure 9.17 [41,42].
The baffles work as a spatial filter, with a pass band centred at normal inci-
dence. Other spatial filter structures could be selected for the same purpose, based
on, for example, dielectric layers or wire grid layers [43].
We give one more example of a proposed dual polarisation waveguide element
using baffles (Figure 9.18). Here lg spaced transverse slots have the grating lobes
suppressed as before using baffles. Longitudinal slots at lg/2 spacing on the
waveguide centreline would normally not be excited. However, in this case the
lower array, also with centred longitudinal slots with post excitation radiates
through the upper longitudinal slots [44].
Similar to Figure 9.16 two waveguides, one for each polarisation could be used
with the lower waveguide having longitudinal slots and the other waveguide
standing on its side having edge slots. This simple principle is illustrated in
Figure 9.19 [45].
Unfortunately, the inclined edge slots will radiate some cross-polarisation due
to the inclination angle. However, this problem can be circumvented by exciting
Antenna systems and special requirements 235
E5
E3 C
lg
7
6
5
lg
10 2
lg
2
11
2
4 E4 E1 1
8
3
E2 9
B
(a) (b)
Figure 9.18 Dual polarised slotted waveguide using baffle for grating lobe
suppression. A breadboard sub-array is shown on the right.
Courtesy of RUAG Space AB
Figure 9.19 Longitudinal and inclined edge slot radiators in a dual polarised
array. IEEE 1996, reprinted from [45], with permission
Rear array
Front array
Figure 9.22 Crossed edge slot arrays according to Alexander and Sichelstiel.
Reprinted from [50]
238 Slotted waveguide array antennas: theory, analysis and design
Directional couplers
Waveguides
Radiating
elements
(a) To receivers
–10
dB –20
–30
–40
–45 –30 –15 0 15 30 45
(b) (c) Degrees
Figure 9.23 The Blass concept, example with five beam ports. McGraw-Hill
Education 1962, reprinted from Skolnik [54] with permission.
(b) Dual beam slotted waveguide array. Courtesy of RUAG Space
AB. (c) Measured dual beam patterns. IEEE 1960, reprinted
from [56], with permission
narrow beam and one shaped broad beam were successfully tested [56,57]; see
Figure 9.23(b) and (c).
In Section 9.2 on resonant travelling wave-type slotted waveguide arrays it was
mentioned that the reflection from the shorted termination causes a second beam to
radiate. Thus, by feeding the two ends of a non-resonant array we can construct a
dual beam antenna with the two beams symmetrically directed relative to broad-
side. However, a high-efficiency symmetrical taper will be difficult to arrange [23].
Classical navigational systems for aircraft normally include an X-band
Doppler radar with four beams pointing towards ground in the directions forward
right and left, and rear right and left. By measuring the Doppler shift of the four
returns the speed and direction of the aircraft can be determined with high
precision. Multi-beam arrays using non-resonant slotted waveguide arrays are
suitable for this, considering their inherent stability and the high precision
Antenna systems and special requirements 239
Port 4 Port 1
Feeding array
20''
Coaxial loads
Waveguide
structure
23''
Radiating
slots
Dielectric
Primary wedges
feed
Geodesic Parallel plate region
lens
Figure 9.25 A cylindrical slotted waveguide array with a lens type rotating
feed. IEEE 1980, reprinted from [61], with permission
In linear and planar array antennas all elements point in the same direction and
we usually write the antenna pattern function as a product of an element factor and
an array factor. Thus, for a linear array we have
X
EðfÞ ¼ ELðfÞ Vn ejnkd sin f (9.13)
n
where EL(f) is the element pattern function, Vn is the element excitation amplitude,
d is the element spacing and k ¼ 2p/l.
For a corresponding circular ring array (cf. Figure 9.27(b)), an element factor
can in general not be factored out as was done in (9.13). We have instead
X
EðfÞ ¼ Vn ELðf nDjÞejkR cosðfnDjÞ (9.14)
n
where Dj is the element angular spacing and R is the circular array radius.
Antenna systems and special requirements 241
Figure 9.26 Curved slotted waveguide array antenna for radar cross section
studies. IEEE 1997, reprinted from [62], with permission
n
d R
n n∆j f
(a) (b)
Figure 9.27 (a) Linear array; (b) part of a circular ring array
One consequence is that for conformal arrays in general the polarisation is not
constant as the beam is steered to different angular positions. This may result in a need to
adjust the polarisation individually for each radiating element [63]. Cf. also Section 9.3
on dual polarised waveguide arrays. However, in the special case of a circular or
cylindrical array of vertical electric dipole elements parallel to the cylinder axis there is
a common element factor with constant polarisation. For the array in Figure 9.25 we can
see the slots as magnetic current sources, all pointing vertically. This results in a con-
stant polarisation but there is still no common element factor that can be factored out.
Stripline-fed tapered slots [64] have been used in broadband cylindrical arrays
[65]. For more complex shapes the polarisation problem gets worse. Circularly
polarised elements is one possible solution. Linear polarisation from slot elements
on a cone is a difficult case [66] as illustrated by Figure 9.28.
242 Slotted waveguide array antennas: theory, analysis and design
Monopulse antennas are used in tracking systems such as radar and missile seekers
where the angular coordinates of detected targets shall be measured, in addition to
measuring the distance. Early tracking systems used conical scan, typically a
reflector antenna with a rotating feed causing the received signal to be amplitude
modulated. Zero modulation resulted when the antenna was pointed correctly on
the target. With monopulse (or simultaneous lobing) a single pulse provides in
principle the necessary information by processing three simultaneously received
signals from the antenna. This minimises the effect of amplitude noise on the
measurement accuracy [54, p. 175].
Figure 9.29 illustrates the principle for a phase monopulse array antenna. The
aperture is divided into four quadrants which are combined in a comparator network
to provide the sum, the azimuth difference and the elevation difference signals.
The lower part of Figure 9.29 is the comparator consisting of four hybrids
(magic Tees or equivalent) that form the sum and difference of the quadrant sig-
nals. The output signals are labelled S, Daz, and Del.
The antenna patterns of the sum and difference signals depend on the illumi-
nation function. For the difference signal one aperture half is 180 out of phase with
respect to the other half, that is, there is a strong discontinuity across the centre
of the aperture where the amplitude is maximum. Very high sidelobes result.
One could of course design an illumination function which is optimum for the
difference pattern, but that would penalise the sum pattern with high sidelobes
(see Figure 9.30).
A typical slotted waveguide array with its feed network is shown in Figure 9.31
[67]. The four quadrant outputs shall be connected to a waveguide comparator.
Optimum distributions for the sum function are, for example, the Taylor
(sampled Taylor) and Dolph–Chebyshev distributions and for the difference func-
tion, the Bayliss distribution [68,69]. The best solution would be if the optimum
distributions could be realised simultaneously and independent of each other.
This is in theory possible since the distributions are orthogonal to each other.
However, the resulting waveguide circuitry would be very complex.
Derneryd [70] proposed a separate central part of the array which is not
included in forming the difference illumination functions (Figure 9.32). For the
sum function the central part is included together with the quadrant parts. The result
Antenna systems and special requirements 243
Array aperture
Comparator
Σ Δ Σ Δ
Σ Σ Δel
Σ
Δaz
Δ Δ
Figure 9.29 A dual plane monopulse antenna system with the aperture
divided into four quadrants
Excitation Pattern
Σ Low sidelobes
High sidelobes
Δ
Δ Low sidelobes
High sidelobes
Inclined series-series
coupling slot
Radiating
slot
Quad. 1 Quad. 2
Radiating
waveguide
Quad. 4 Quad. 3
Figure 9.31 A typical slotted waveguide array antenna with one feeding point in
each quadrant. One quadrant is shown to the left; the complete
antenna (rear view) on the right. Courtesy of Microwave Journal,
reprinted from [67], with permission
Figure 9.32 Partitioning of a monopulse array antenna with a fifth sector added.
Reprinted from [70]
Figure 9.33 Complex waveguide feed network on the rear of a large planar array.
Raytheon Company. Reproduced from [72], with permission
9.7.1 Introduction
A stripline slotted array antenna has slots in a metal ground plane fed from stripline
circuitry in a layer below the slots [Figure 9.34(b)]. Just like the linear slotted
waveguide array a linear stripline slot array can be envisioned with either long-
itudinal or transverse slots [73]. The slot positions are not bound to straight feeding
waveguides, but can be chosen more freely, depending on the pattern requirements
and the layout of the feeding circuit [74]. Some of the main characteristics of
waveguide and stripline technologies are compared in Table 9.1.
Stripline circuits can be fabricated using printed circuit technology and are
therefore less expensive than waveguide circuits. On the other hand, metal
246 Slotted waveguide array antennas: theory, analysis and design
x0
x0
(a) (b)
Figure 9.34 (a) Waveguide slot, offset from the waveguide centreline.
(b) Stripline slot with the feed line offset from the slot centre
waveguides can withstand higher temperatures. There are also other technologies
such as waveguides made of metallised CFRP, see Chapter 11, and metal-plated
plastic waveguide antennas, described in Chapter 10.
In a stripline array the TEM mode is the dominant transmission line mode. In
fact, the structure is similar to a coaxial line, albeit in a flat configuration. Side
walls are necessary for preventing parallel plate modes to escape. The characteristic
impedance of a stripline transmission line can be calculated or obtained from
published graphs, see, for example, [75]. As an example we assume the following
parameters (Table 9.2).
A calculation gives the result: w ¼ 2.56 mm for 50 W impedance.
The slot can be symmetrically fed from the stripline in the centre of the slot.
An offset feed position reduces the impedance [76]. The impedance depends also
Antenna systems and special requirements 247
(a) (b)
on the dimensions of the cavity formed under the slot, and the dielectric material
[77,78]. At resonance a sinusoidal field distribution results along the slot. As we
know from Chapter 4.2, the slot antenna has by itself rather high radiation impe-
dance, typically around 500 W and some way of transforming the impedance to
match the stripline feed is necessary. An example is shown in Figure 9.35, where
the circuit layout of a stripline flat plate antenna [79] is shown. Each slot is fed
from a narrow strip of length lg/4, which acts as an impedance transformer. The
stripline continues as an open stub about lg/4 beyond the slot. The stub length is
adjusted to tune the reactive part of the slot impedance. Alternatively, an off-centre
feed position can be chosen as in Figure 9.34(b). However, this can cause problems
with the circuit layout since a great deal of symmetry is lost.
Δ Δ
Σ
Δ Δ
ΔE Σ ΔH
Δ Δ
Σ
Figure 9.36 Feed principle for independent excitations with three sectors
per quadrant. Reprinted from MEDE ’77 [80]
dB
0
In vacuum bag
–10
After bonding
–20
–30
Figure 9.37 Measured difference pattern (H-plane). The dashed line was
measured before bonding of the antenna layers. Reprinted from
MEDE ’77 [80]
Antenna systems and special requirements 249
Figure 9.38 Slotted monopulse stripline array antenna with three functional
layers. Reprinted from MEDE ’77 [80]
–20
–25
–30
σTETE (dBsm)
–35
–40
–45
φ
–50
–55
–60
–180 –140 –100 –60 –20 20 60 100 140 180
(a) (b) φ (degrees)
3 4
Figure 9.40 Slot boundary problems numbers 3 and 4, waveguide slot and
slot in a ground plane
y k
q
cf. (5.44).
The problem no. 4 in Figure 9.40, the slot in a ground plane (no waveguide),
is easily solved in the same manner. We have Hzi ¼ Hze and thus
14 GHz
9.375 GHz
0 –5 –10 –5 0
(dB)
Figure 9.42 Bistatic scattering from a longitudinal slot in a waveguide. The angle
of incidence is q ¼ 60
Now that we have the electric field distribution in the slot aperture we can
derive the electric vector potential in the far field
eejkr
Fz ¼ ½E ½hðkx ; kz Þ (9.21)
2pr
where [h(kx, kz)] is the same as the excitation vector, cf. (9.19), but with (kx, kz) now
representing the outgoing direction. The far field is easily obtained since
Hq ; Hj ¼ jw Fq ; Fj (9.22)
The bistatic scattering cross section is defined as s ¼ limr!1 4pr2 Pscat =Pinc ,
where Pinc and Pscat are the incident and scattered power densities, respectively,
yielding
k 2 kz2
2
s¼ ½E ½h
(9.23)
pZ0
In this expression kz represents the outgoing direction, which is also contained
in [h]. The incidence direction is implicit in the solution for [E].
Figure 9.42 shows a scattering diagram for a case with a waveguide slot
illuminated from (q, j) ¼ (60 , 0 ). The slot length is 16 mm and thus close to
resonance for the frequency 9.375 GHz. At the higher frequency an asymmetric
slot field results due to the presence of higher slot modes. This is illustrated in
Figure 9.43 showing the electric field distribution in the slot [83].
0.8
Electric field amplitude
0.6
0.4
0.2
0
0 0.2 0.4 0.6 0.8 1
Position along aperture
Figure 9.43 Slot electric field along the slot for the same case as in
Figure 9.42. Solid line is 14 GHz, and broken line is 9.375 GHz.
IEEE 1990. Adapted from [83], with permission
ky/k
l/dy
Visible region
dy
gl inc n r 1 kx/k
l/dx
sin θi dx
inc n
r
gl
Figure 9.44 The grating lobe lattice. The incidence wave (‘inc’) excites the slots
with a phase slope causing a reflected wave (‘r’). In this example the
grating lobe (‘gl’) is close to radiating back towards the illuminator
slots. At oblique incidence from the illuminator the slots are excited with a phase
slope and radiate like a phase steered array. The grating lobe lattice, the reciprocal
of the slot element lattice [84], is shown in Figure 9.44 for an array with slots in a
rectangular grid.
254 Slotted waveguide array antennas: theory, analysis and design
l
sin qinc þ sin qgl ¼ m ; (9.24)
d
l
sin qinc ¼ sin qgl ¼ m (9.25)
2d
The array ground plane scatters like a metal plate of the same size. Based on a
physical optics approximation (neglecting edges) the monostatic cross section for a
2
plate with area A is s ¼ 4pA
l2
in the normal direction. The approximate angular
dependence for a square plate with side L is
sinðkL sin qÞ 2 2
sðqÞ ¼ sð0Þ cos q (9.26)
kL sin q
see [85,86].
The simple expressions such as (9.26) for the rectangular plate work well near
the normal direction. An analysis based on equivalent currents along the plate
edges can predict the scattering more accurately outside the principal symmetry
planes of rectangular plates, as demonstrated by Polka and Balanis [87].
A circular plate with 40 cm radius has about 3,000 m2 radar cross section at
X-band. This is quite a large cross section. The contribution from the slot array can
be of the same order of magnitude [88].
It is evident that bending of the array defocuses the returns, that is, conformal
arrays have an advantage here. A curved array of slotted waveguides was shown in
Figure 9.26. We compare in Figure 9.45 the scattering from a planar and a
cylindrical array of waveguide apertures, in both cases 40 32 apertures covering
about 120 of the 600 mm diameter (¼13.6 l0) cylinder.
The example in Figure 9.45 is for the H-plane. In the E-plane the comparison is
not so favourable for the conformal case due to the specular return for all j values.
See also [62].
Antenna systems and special requirements 255
40 z
30 ∞ ∞
20
<σTETE> (dBsm)
10
H-plane
0
–10
–20 E-plane
–30
R = 0.3 m x
R = 100.0 m
–40
0 20 40 60 80 100 120 140 160 180 ∞ ∞
(a) q (degrees) (b)
Figure 9.45 (a) Monostatic radar cross section for a waveguide array with 40
32 slots on a cylinder (solid line) compared with a corresponding
array on a planar surface (dashed line). H-plane, (b) shows the
geometry. IEEE 2003, reprinted from [81], with permission
2
Y0
2
Y2 P2
1 Y0 1 q
Y3 3
P3
Y0
3
(a) (b)
The impact of the feeding system on the scattering from slotted waveguide
array antennas is not much discussed in the literature with few exceptions; see, for
example, Tittensor and Newton [88]. There are also various ways of modelling feed
networks [89].
–10 Total
Field amplitude in dB
PO
–20
–30
–40
–50
–90 –60 –30 0 30 60 90
Theta in degrees
Figure 9.47 Scattered field in the H-plane for a rectangular slotted waveguide
array antenna. Incidence is from qinc ¼ 45 . IEEE 2007,
reprinted from [92], with permission
slotted waveguide array antenna has been described by Stanek and Johansson [94]
(see Figure 9.48).
The FSS radome in Figure 9.48 is built up of two metal skins and a dielectric
support. The 2,000 þ slots in each skin are laid out in a quasi-periodic pattern due
258 Slotted waveguide array antennas: theory, analysis and design
to the curvature, and designed for equal phase delay within 5 and minimum loss
within the operating frequency band. Out-of-band (and for the orthogonal polar-
isation) the radome scatters as a solid spherical surface. It can be viewed as a large
conformal slotted array (but with no waveguides).
9.9.1 Introduction
The typical linear slotted waveguide array has longitudinal slots separated at about
one-half waveguide wavelength (0.5 lg along the waveguide). The slot excitation is
governed by the amount of slot offset from the waveguide centreline, with equal
phase radiation obtained by alternating the offset directions (Figure 9.49).
Since the radiator spacing of 0.5 lg corresponds to about 0.7 l0 in free space
the first grating lobe appears at an angle corresponding to sin qg ¼ 1/0.7 1.4, that
is, outside visible space (cf. Figure 9.11). However, considering the alternating slot
arrangement, the array period is actually twice that of the single slot, that is, the
array element is strictly speaking the slot pair. Thus, a grating lobe will appear at
sin q ¼ 0.7 or circa 45 from broadside. However, in the vertical plane containing
the array symmetry line such a grating lobe will not appear. But outside this plane
the phase differences due to the slot offsets will result in increased sidelobes,
or ‘butterfly lobes’ [95,96] (see Figure 9.50). This phenomenon, also named
‘Gruenberg lobes’, has been studied by several authors [97–101].
9.9.2 Analysis
For the simple case shown in Figure 9.49 the radiation patterns with and without
slot offsets are easily calculated. Assuming the slots to be radiating isotropically
dy
(a)
(b)
Figure 9.49 (a) A linear slot array, with the slot offset from the centreline
indicated (dy). (b) A corresponding real antenna (with tapered
excitation)
Antenna systems and special requirements 259
z 21
y
19
18 16
12 11 12 11
x
17
20
Figure 9.50 Butterfly lobes (numbers 18, 19, 20 and 21 in the figure) due to slot
offsets showing up outside the cardinal planes. Reprinted from [96]
with identical polarisation we only need to examine the array factor. Two examples
are shown in Figure 9.51.
As predicted, the secondary lobe appears at the grating lobe location kx =k ¼ 2dx l
where dx is the slot spacing along the waveguide. The magnitude of the disturbance
is proportional to the slot offset from the centreline; it follows the cone kx ¼ constant
and increases linearly with the transverse coordinate ky.
It is possible to reduce the secondary lobe amplitude somewhat by breaking up
the symmetry of the array; see Section 7.6. A possible layout from Ahlbom et al.
[18] is shown in Figure 9.52. This type of layout applied to our eight-slot example
(Figure 9.49) given earlier gives the result shown in Figure 9.53. See also [97,101].
A phase steered array with slotted waveguides as radiating elements (as in
Figure 9.1) will also suffer from secondary lobes due to the slot offsets. As the main
beam is steered the secondary lobe will be steered as well. It may then be moved
into a region where the sidelobes otherwise are quite low (see Figure 9.54).
In Figure 9.55 the radiation over the full kx/ky-plane is shown for the same
array, now with a –26 dB Dolph–Chebyshev taper in the E-plane and a slot offset of
0.1 wavelengths. The waveguide width is 0.77 l and the main beam is steered
11.5 in the E-plane (about the maximum possible for normal waveguides; cf.
Figure 9.2).
–5
–10
–15
dB
–20
–25
1
–30
0
0.5
0.5 kkyy/k
/k
(a) kx/k
1 0
–5
–10
–15
dB
–20
–25
1
–30
0
0.5
0.5 ky/k
(b) kx/k
1 0
Figure 9.51 Radiation patterns in one quadrant with (a) no slot offset,
(b) slot offset dy ¼ 0.1 wavelengths. Eight slots, spacing
along waveguide dx ¼ 0.7 wavelengths
a straight line (Figure 9.56). Another possibility is to use post or iris excited slots
(Figures 9.57 and 9.58).
The iris can be inductive or inductive-capacitive (compound iris) as in
Figure 9.58 (see Section 8.4). The iris can be tuned together with the slot to improve
the overall bandwidth of the radiator [103,104]. The ridge waveguide with alternating
depths discussed in connection with Figure 9.4 represents still another solution.
Antenna systems and special requirements 261
Figure 9.52 Slot panel. The upper half is a mirror image of the lower half.
Courtesy of Saab AB
–5
–10
dB
–15
–20
–25
1
–30
0 0.8
0.2 0.6
0.4 0.4
0.6
kx/k 0.8 0.2 ky/k
1 0
Figure 9.53 Computed pattern for a single waveguide with mirrored slot
layout. Cf. Figure 9.51(b)
All these examples refer to shunt slots in the broad wall of the waveguide. The
secondary beam problem occurs also in shunt slots in the narrow wall of
the waveguide. In those, instead of controlling the excitation with slot offsets,
the excitation is controlled by rotating the narrow wall slots. Second-order beams
262 Slotted waveguide array antennas: theory, analysis and design
0.8
0.6
Abs
0.4
0.2
0 1
0
0.2 0.5
0.4
0.6 ky/k
(a) kx/k 0.8
1 0
0.8
0.6
Abs
0.4
0.2
1
0
0
0.2 0.5
0.4
0.6 ky/k
(b) kx/k 0.8
1 0
Figure 9.54 Array factor for a planar array of eight slotted waveguides. The
beam is steered 17.5 in the E-plane. (a) No slot offset. (b) With
a slot offset ¼ 0.3 wavelengths. Note the vertical axis scale is
here in linear voltage (magnitude)
0 G
–5
–10
dB
–15
–20
–25
–30 1
–1 0.5
–0.5 0
0
(a) kx/k 0.5 –0.5 ky/k
1 –1
M
B
0
B
–5
–10
–15
dB
–20
–25
–30 1
–1 0.5
–0.5 0
0
(b) kx/k 0.5
–0.5 ky/k
1 –1
Figure 9.55 (a) Pattern with no slot offset; (b) pattern with slot offset.
M denotes the main beam, steered 11.5 in the y/z plane. G is
the corresponding (part of the) grating lobe. B indicates the
four butterfly lobes
Figure 9.56 Meandering the waveguide keeps the slots on a straight line.
Reprinted from Kaminow and Stegen [102]
264 Slotted waveguide array antennas: theory, analysis and design
lg
2
l
2
polarised waveguide arrays). Baffles for a longitudinal slot array were described in
a patent by Gruenberg [106,107] (Figure 9.60).
For planar arrays with rotated edge wall slots the external space between
adjacent waveguides can act as a choke which reduces the currents associated with
cross-polarisation radiation. An optimum choke depth is about a quarter of a
wavelength [10].
Antenna systems and special requirements 265
Figure 9.59 Non-tilted edge slots with wire excitation. IEEE 1990,
reprinted from [105], with permission
(a) (b)
Figure 9.62 (a) The lowest order rotationally symmetric mode in a circular
waveguide is TM01. (b) The lowest order propagating mode in
a circular waveguide is TE11. Solid lines indicate electric field;
broken lines are magnetic field lines [120]
268 Slotted waveguide array antennas: theory, analysis and design
frequency than the TM01 [119,120]. A proper symmetric feeding system is there-
fore required to prevent excitation of this mode.
With TM01 (or a coaxial waveguide) the currents flow along the length of the
guide. For exciting axial slots probes can be used as shown in the example in
Figure 9.63. Here additional screws/probes are also seen in between each set of
radiating slots, claimed to improve impedance matching. One can also use inclined
or crossed slots as was shown by Grabherr and Huder [121] and others.
A beacon antenna [122] for navigational purposes is shown in Figure 9.64. It
has an upper circular slotted array for horizontal polarisation (X-band). The lower
circular array for S-band has just one row of inclined slots covering both horizontal
and vertical polarisations. The biconical horn arrangement provides beam shar-
pening in elevation; the vertical coverage for both antennas is about 20 .
(a) (b)
Figure 9.63 (a) Slotted cylindrical waveguide array; (b) cross section
showing probes for exciting slots [34]
Figure 9.64 A radar beacon antenna system using circular waveguide arrays and
an electronics package and radome for X- and S-band operation.
Courtesy of Telefonaktiebolaget L M Ericsson, reprinted from [122]
Antenna systems and special requirements 269
The coefficients Cm are phase mode amplitudes; the mth mode has m times 2p
phase variation along the circumference. The radiated far field can similarly be
written as a Fourier series:
X
1
EðfÞ ¼ Am ejmf (9.30)
1
R y
R
x φ
2L
(a) (b)
Figure 9.65 (a) An array of slots in a conducting cylinder; (b) one slot
with coordinate definitions
270 Slotted waveguide array antennas: theory, analysis and design
The radiated pattern from an N-element circular array can also be written as a
summation of the element contributions:
X
EðfÞ ¼ Vn ELðf nDjÞejkR cosðfnDjÞ (9.31) = (9.14)
n
where EL(f) is the element pattern and Dj ¼ 2p/N is the element angular spacing
around the circle. The element pattern is the same for all elements although the
elements point in different directions. Since it also is periodic, period 2p, the ele-
ment pattern can be written in Fourier series form as
X
1
ELðaÞ ¼ Dp ejpa (9.32)
p¼1
Inserting also the excitation V(j) according to (9.29) and taking out just the
contribution from phase mode number m yields the expression
X1 ð
1 p jðmpÞðjfÞ kR cosðjfÞ
Em ðfÞ ¼ Cm ejmf Dp e e dj (9.34)
p¼1
2p p
or
" #
X
1
Em ðfÞ ¼ Cm Dp j mp
Jmp ðkRÞ ejmf ¼ Am ejmf (9.35)
p¼1
2LV0 ejkr X
1
jn enf
Ef ¼ 0 (9.36)
p3 Rr 1 Hnð2Þ ðkRÞ
where V0 is the peak of the sinusoidal voltage across the slot with length 2L, and
0
Hnð2Þ is the derivative of the Hankel function of the second kind and order n.
Antenna systems and special requirements 271
90 40 dB
120 60
30
150 20 30
10
180 0 deg
210 330
240 300
(a) 270
1 1
0.75 0.75
Magnitude
Magnitude
0.5 0.5
0.25 0.25
0 0
–2 –1 0 1 2 –2 –1 0 1 2
(b) Spectral order m/2 (c) Spectral order m/2
Figure 9.66 (a) Axial slot pattern (solid line) and function (1 þ cos f) (dashed
line). Mode spectra for slot (b) and (1 þ cos f) function (c).
Wiley 2006, reprinted from [63], with permission
272 Slotted waveguide array antennas: theory, analysis and design
10
8
4
5
6
Ripple dB
7
4 6
8
9
0
2 3 4 5 6 7 8 9
kR = N*d/λ
Figure 9.67 Maximum ripple versus cylinder size kR for several numbers N of
slots with assumed (1 þ cos j) element patterns
15
10
No of elements
1 dB
3 dB
10 dB
0
0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
d/λ
Figure 9.68 Ripple levels indicated by contour lines: solid curve 1 dB, dashed
3 dB and dotted 10 dB. (1 þ cos j) element patterns assumed
always be a problem for applications where bandwidths of the order of 10% or less
are required. In any case, the bandwidth ‘problem’ deserves our attention and many
of the chapters in this book have already addressed this issue.
In this last section of the chapter we discuss bandwidth limiting factors from an
antenna system point of view, without going into the details of specific designs.
Design rules are addressed in a broad sense, so that the reader can get a feeling for
what is important and what phenomena are involved. Cross-references to relevant
parts of the book and references to the literature are given. The index at the end of
the book also provides aid to find more information.
(a) (b)
Figure 9.69 (a) Stripline feed (hatched) with multiple slots for increased
bandwidth; (b) dumbbell slot
0.9
0.8
0.7
Conductance G/Y0
1%
4
Wavelengths*103
–4
–1%
–8
0 2 4 6 8 10
Offset (mm)
Figure 9.71 Calculated change in resonant length when the slot width is changed
from 3 mm to 1.6 mm, for different slot offsets. Standard X-band
waveguide with wall thickness 1.27 mm
φ1 φ2 φN
y1 y2 y3 yN
Figure 9.72 Equivalent network for an end-fed linear shunt slot array antenna
c1+ c2+
W
c1– c2–
e jφ 0
φ W=
0 e–jφ
1 2+y y
y W=
2 –y 2 – y
z 1 2 + z –z
W=
2 z 2–z
We can write
! !
cþ cþ
1
¼W 2
(9.37)
c
1 c
2
or in short form
c1 ¼ W c2 (9.38)
The total wave matrix for the complete ladder network such as shown in
Figure 9.72 is found by multiplying the matrices of the individual blocks (9.39).
Y
N
tot ¼
W i
W (9.39)
i¼1
From the total wave matrix the total transmission and reflection are obtained
according to
(
T ¼ 1=W11
(9.40)
G ¼ W21 =W11
The wave matrices for some elementary networks are shown in Figure 9.74.
Antenna systems and special requirements 277
2ð3 þ q2 Þ
a¼ where q ¼ pNdf =100 (9.41)
qð2 þ q2 Þ
The return loss (reflection factor) according to [135] is then obtained as
pffiffiffiffiffiffiffiffiffiffiffiffiffi
1 þ 1 þ a2
G¼ pffiffiffiffiffiffiffiffiffiffiffiffiffi (9.42)
1 þ a2 þ 1 þ a2
The two approaches agree quite well; Figure 9.75, cf. also Figure 6.2.
It must be emphasised that these calculations are approximate. However, they
indicate the general behaviour in terms of frequency sensitivity and number of slots
fed in series. The following effects are not accounted for
● slot admittance is frequency sensitive
● the dispersion of the waveguide itself
● mutual coupling among slots
● impact on radiation patterns
A commonly used expression for the relative bandwidth in % is 50/N, where N
is the number of slots fed in series. Examination of Figure 9.75 reveals that 60/N is
a good measure for the (total) bandwidth if 10 dB return loss is postulated. If
15 dB is required 30/N gives a good match.
9.11.5 Overloading
The term ‘overloading’ refers to choosing the slot conductance values higher than
G0/N where G0 is the waveguide admittance, for instance R G0/N, with an over-
load factor of R [34]. The input admittance of the array will then be increased and
278 Slotted waveguide array antennas: theory, analysis and design
–5
–10
N=8
Return loss dB
–15
N=6
–20 N=4
–25 N=2
–30
0 1 2 3 4 5
Frequency df %
Figure 9.75 The nominal (half) bandwidth in per cent is obtained from the return
loss versus frequency for arrays with two, four, six and eight slots.
The solid line is according to the wave matrix analysis; the broken
line according to the formula (9.42). No overloading
an impedance transformer is required on the input in order to match the array. The
transformer pcould
ffiffiffi be a waveguide section with a quarter wave length and admit-
tance Gt ¼ RG0 [131]. A similar overloading scheme can be implemented for
arrays with series slots where the slots are represented by series elements.
Results obtained by overloading by a factor of R ¼ 2 are shown in Figure 9.76
and compared with the previous case with no overloading.
Overloading improves the impedance match also for an amplitude tapered
array. The amplitude tapering is usually introduced in order to improve the pattern
characteristics and reduce the sidelobe level. However, in a study by Derneryd and
Petersson [137] regarding a typical planar array antenna it was concluded that
overloading should be avoided when maximum bandwidth from a sidelobe point of
view is desired. In the practical case a careful trade off will have to be made.
–5
–10 8
6
Return loss dB
–15 4
–20
8
6
–25 4
–30
0 1 2 3 4 5 6 7 8 9 10
Frequency df %
Figure 9.76 The return loss versus frequency for arrays with four, six and
eight longitudinal slots. The wave matrix analysis was used;
the solid line is without overloading, the dashed line with an
overloading factor of 2
It has been found that breaking up the periodicity of a planar slot array can
reduce the bandwidth limitation caused by the mutual coupling. Instead of having
identical radiating waveguides with alternating slot offsets, adjacent waveguides
can be designed to be mirror images of each other. In one example [139,140] the
improvement was about a doubling of the reflection bandwidth, from 1.2% to 2.5%
at 15 dB return loss (at the expense of some grating lobes in the diagonal planes).
0.2
Slot in array
0.1
0.05
0
8.5 9 9.5 10
(a) Frequency GHz
0.1
Slot in array
–0.05
–0.1
8.5 9 9.5 10
(b) Frequency GHz
Figure 9.77 (a) Normalised conductance and (b) normalised susceptance versus
frequency for an isolated longitudinal slot compared to a slot in an
infinite array. Standard X-band (R 100) waveguide, slot width
1.59 mm, total slot length 14.76 mm, offset 2 mm from centreline. Slot
spacing in the array 0.7 l0 along waveguides, 0.75 l0 across [138]
the array diameter, provided that there is only one feeding point for the complete
circular array. If there are more feeding points we can write the bandwidth in GHz as
pffiffiffiffiffiffi
BðGHzÞ ¼ 250 Np =D (9.44)
where Np is the number of feeding points per quadrant and D is the diameter in
millimetre (see Figure 9.78).
In this analysis we have not included the radiation pattern performance,
tapering effects, overloading or the precise shape of the array. In a detailed
Antenna systems and special requirements 281
Bandwidth (MHz)
1,400
1,200
1,000
800
600
400
200
0
200 300 400 500 600 700 800
Diameter (mm)
investigation the influence from coupling slots, mutual coupling, waveguide short
design, impedance levels and transformers, etc., will have to be included [141].
The ideal solution could be to use a corporate feed broadband power divider
to feed all radiating waveguides in parallel, or even to feed all individual slots by
such a power divider. In one solution by Zhou et al. [142] a 1:16 waveguide
power divider feeds 2 2 sub-arrays with radiating wide slot apertures. Better
than 30% bandwidth (return loss < 10 dB) was obtained at 12 GHz. Another
interesting feed network technology is based on barline transmission lines which
essentially is a coaxial line technology. It is more broadband and offers better
design flexibility than the traditional waveguide feed using rotated coupling
slots. Dual barline power dividers 1:8 were used to feed the dual polarised
slotted waveguides that make up the panels of a large C-band scatterometer
instrument [47].
References
[1] Amitay N., Galindo V., Wu C. P. Theory and Analysis of Phased Array
Antennas. Wiley Interscience, New York, 1972.
[2] Josefsson L., Derneryd A., Karlsson E. R. ‘A phase steered slotted wave-
guide array with non-squinting beam’. Proceedings Military Microwaves.
London, July, 1990, pp. 251–256.
[3] Sen B., Ranga Rao K. S. ‘Impedance properties of narrow radiating slots in
broadface of dielectric loaded rectangular waveguide’. Proceedings of the
IEEE Antennas and Propagation Symposium, Vancouver, Canada, 1985,
pp. 191–194.
[4] Rengarajan S. R., Steinbeck M. ‘Longitudinal slots in dielectric filled rec-
tangular waveguides’. Proceedings of the IEEE Antennas and Propagation
Symposium, London, Ontario, Canada, 1991, pp. 1276–1279.
[5] Joubert J., Coetzee J. C. ‘The use of inhomogeneously loaded rectangular
waveguide to overcome grating lobes and distribution taper limitations in
transverse slot arrays’. Proceedings European Microwave Conference. 1992,
pp. 870–875.
[6] Cohn S. B. ‘Properties of ridge wave guide’. Proceedings of the IRE. 1947;
35(8):783–788.
[7] Hopfer S. ‘The design of ridged waveguides’. IRE Transactions on Micro-
wave Theory and Techniques. 1955;3(5):20–29.
[8] Hu A. Y., Lunden C. D. ‘Rectangular-ridge waveguide slot array’. IRE
Transactions on Antennas and Propagation. 1961;9(1):102–105.
[9] Shnitkin H. ‘Unique joint stars phased-array antenna’. Proceedings of the IEEE
Antennas and Propagation Symposium, Dallas, USA, 1990, pp. 678–681.
[10] Hilburn J. L., Kinney R. A., Emmett R. W., Prestwood F. H. ‘Frequency-
scanned X-band waveguide array’. IEEE Transactions on Antennas and
Propagation. 1972;20(4):506–509.
[11] Bostrom D. E., Castaneda J. A., Hodges R. E., Fujii P. ‘Large X-band
electronically scanned array’. Proceedings of the IEEE Antennas and
Propagation Symposium, Los Angeles, CA, USA, 1981, pp. 213–216.
[12] Elsallal M. W., Herting B. J., West B. J. ‘Planar edge slot waveguide antenna
array design using COTS EM tools’. Proceedings of the 2007 Antenna
Applications Symposium, Allerton Park, IL, USA, 2007.
[13] Conciauro G., Bressan M., Zuffada C. ‘Waveguide modes via an integral
equation leading to a linear matrix eigenvalue problem’. IEEE Transactions
on Microwave Theory and Techniques. 1984;32(11):1495–1504.
[14] Falk K. ‘Admittance of a longitudinal slot in a ridge waveguide’. IEE Pro-
ceedings H – Microwaves, Antennas and Propagation. 1988;135(4):263–268.
[15] Garb K., Kastner R., Meyerova R. ‘Analysis of planar arrays of shunt slots in
ridged waveguides’. Electronics Letters. 1994;30(7):533–534.
[16] Spradley J. L. ‘A volumetric electrically scanned two-dimensional microwave
antenna array’. IRE International Convention Record. 1958;6:204–212.
Antenna systems and special requirements 283
[104] Dudley D. G. ‘An iris-excited slot radiator in the narrow wall of rectangular
waveguide’. IRE Transactions on Antennas and Propagation. 1961;9(4):
361–364.
[105] Hashemi-Yeganeh S., Elliott R. S. ‘Analysis of untilted edge slots excited
by tilted wires’. IEEE Transactions on Antennas and Propagation. 1990;
38(11):1737–1745.
[106] Gruenberg H. ‘Theory of waveguide-fed slots radiating into parallel-plate
regions’. IRE Transactions on Antennas and Propagation. 1952;3(1):63–66.
[107] Gruenberg H. Microwave Antenna. U.S. patent 2 659 005, Patented
Nov. 10, 1953.
[108] Roberts A. (ed.) Radar Beacons. MIT. Rad. Lab Series, Vol. 3, 1947.
[109] Galindo V., Green K. ‘A near-isotropic circularly polarized antenna for
space vehicles’. IEEE Transactions on Antennas and Propagation. 1965;
13(6):872–877.
[110] Scott W. G., Soo Hoo K. M. ‘A theorem on the polarization of null-free
antennas’. IEEE Transactions on Antennas and Propagation. 1966;14(3):
587–590.
[111] Kraus J. D. Antennas, 2nd edition, McGraw-Hill, New York, 1988, p. 89.
[112] Wait J. R. ‘On the conductance of slots’. IRE Transactions on Antennas
and Propagation. 1956;4(2):124–127.
[113] Wait J. R. Electromagnetic Radiation from Cylindrical Structures.
Pergamon Press, London, 1959.
[114] Gayen R. K., Das S. D., Das A. C. ‘A high gain-bandwidth, nearly omni-
directional waveguide slot array antenna’. Proceedings of the IEEE Antennas
and Propagation Symposium, Vancouver, BC, Canada. 2015, pp. 222–223.
[115] Li J-Y., Li L-W. ‘Analysis of omnidirectional waveguide slots array
antennas’. Proceedings of the 6th International Symposium on Antennas,
Propagation and EM Theory, ISAPE’03, Beijing, October 28–November 1,
2003, pp. 38–41.
[116] Solbach K. ‘Simulation of slot array antennas using standard network analysis
program’. IEEE Antennas and Propagation Magazine. 1991;33(6):45–47.
[117] Croswell W. F., Knop C. M., Hatcher D. M. ‘A dielectric-coated
circumferential slot array for omnidirectional coverage at microwave
frequencies’. IEEE Transactions on Antennas and Propagation. 1967;15(6):
722–727.
[118] Fiet O. O., Scudder R. M. Ultrahigh-Frequency Broadcast Antenna System.
US patent 2 658 143, November 1953.
[119] Lee C. S., Lee S. W., Chuang S. L. ‘Plot of the modal field distribution in
rectangular and circular waveguides’. IEEE Transactions on Microwave
Theory and Techniques. 1985;33(3):271–274.
[120] Marcuvitz N. Waveguide Handbook. MIT. Rad. Lab Series, Vol. 10, 1951.
[121] Grabherr W., Huder B. ‘Omnidirectional slotted-waveguide array antenna’.
Proceedings of the 29th European Microwave Conference, Munich, Germany.
1999, pp. 134–137.
Antenna systems and special requirements 289
In this chapter, we discuss slot arrays in two types of parallel plate waveguides
propagating the TEM mode. In the first an electromagnetic wave propagates in the
radial direction inward or outward in a circular parallel plate waveguide, while slots
cut in one of the parallel plates radiate. Such an array is convenient for large
apertures requiring high gain and mass production. The second type has a rectan-
gular shape and propagates the TEM mode. Slots cut in one of the plates radiate.
We then present slot arrays in substrate integrated waveguides (SIW) which have a
great potential for easy integration with planar devices. Finally slot arrays in gap
waveguides are discussed.
Figure 10.1 A typical radial line slot antenna. IEEE 1961, reprinted from [2],
with permission
Figure 10.2 Monopulse radial line slot array. IEEE 1964, reprinted from [3],
with permission
Slot arrays in special waveguide technologies 293
nearly uniform aperture distributions required for high gain applications. In this struc-
ture the increase in power density of the wave as it propagates inward is offset by the
power radiated by the slots. In a single-layer radial waveguide antenna power flows
radially outward. It is better suited to the design of a tapered amplitude distribution.
Since a systematic design of the two-layer radial line slot antenna was presented first,
we start with a discussion of that structure. Figure 10.3 shows a radial line slot antenna
consisting of a two-layer circular parallel plate waveguide [4]. The lower parallel plate
waveguide is excited by an outward travelling radial TEM mode. In the upper parallel
plate waveguide consisting of radiating slots on the upper conductor, the TEM mode
propagates inward. It is then terminated in an absorber at the centre.
Figure 10.3(a) shows the two-layer antenna structure. Figure 10.3(b) illustrates
the power flow direction in both parallel plate waveguides. The antenna has a
coaxial cable to radial waveguide adaptor and a 180 E-bend between lower and
upper radial lines. In the upper waveguide orthogonal slot pairs couple power from
the radial waveguide and radiate a circularly polarised broadside beam. An
absorbing material placed near the centre at the radial value rm of the upper
Y D2
Ψ ρ X
O
Sρ
Sφ
ρM
Z
ρm ρ
d
(b)
Figure 10.3 A two-layer radial line slot array: (a) complete antenna structure;
(b) direction of power flow in the radial waveguides. IEEE 1985,
reprinted from [4], with permission
294 Slotted waveguide array antennas: theory, analysis and design
#1
#1
β1
ρ1 Θ1 2L ρ1 Θ P1
O X O
1
X
β1 ρ2
P1 Φ3
β3 ρ3 Θ1 #2
Sφ
ρ4 Θ3 P2
ρ2 δ #3
Θ1 Θ3
P3
P2
#4
#2 P4
–Y
(a) (b)
Figure 10.4 Slot arrangements in the upper wall: (a) first slot pair; (b) first two
slot pairs. IEEE 1985, reprinted from [4], with permission
waveguide dissipates the remaining power in the travelling wave. In order to pro-
duce a good match at the coaxial cable input, the coaxial cable to radial waveguide
adaptor and the 180 E-bend have to be designed to have very little reflection.
In addition, reflection from the slot pairs shall be minimised as well.
The arrangement of slot pairs on the top plate of the upper parallel plate
waveguide is illustrated in Figure 10.4. The two slots of each slot pair are ortho-
gonal to each other with nearly the same amplitude of excitation and 90 phase
difference so as to produce a circular polarisation. The current distribution in the
plate for an inward travelling wave will be in the form of a Hankel function of the
first kind and order zero, whose phase term may be approximated by a cylindrical
phase function. Therefore the radial coordinates of the centres of the slot pair are
spaced a quarter guide wavelength so that their excitation phases are 90 apart as
shown by (10.1) and (10.2) [4].
radial spacing Sr. The first slot of each slot pair makes the same angle with respect
to the radial line from the origin to the slot centre (e.g. Q1 ¼ Q3). In order to
produce a broadside beam, the two slot pairs in Figure 10.4(b) should radiate in
phase in the direction normal to the aperture. This requires that the tilt angle of the
second slot pair with respect to the first compensate the excitation phase difference
as specified by (10.3).
Lower
radial waveguide W1
θ
C
A
B
Coaxial cable
εr
r W0
Upper θ2
radial W.G. W2
h1 h2
D
Lower
radial W.G. W1 θ1
Figure 10.6 180 E-bend between lower and upper radial lines (D ¼ 12 mm, q1 ¼
43.8 , q2 ¼ 38.7 , W1 ¼ 12.5 mm, W2 ¼ 15 mm, h1 ¼ h2 ¼ 4 mm).
IEEE 1987, reprinted from [5], with permission
296 Slotted waveguide array antennas: theory, analysis and design
symmetric with respect to a vertical axis. The radii of the inner and outer conductors of
the coaxial cable are r and r þ W0, respectively.
Ando et al. [5] used the commercial finite element analysis code HFSS to
design the parameters A, B, C, q, W0, W1 and r of the adaptor and D, q1, q2, W1, W2,
h1 and h2 of the bend at 6 GHz. The reflection coefficients of these devices were
found to be less than 0.1 in the frequency range of 5.5–6.5 GHz. One could use any
computational electromagnetic code to design these devices with a starting value
obtained by scaling these parameters for the frequency range of interest.
α = 0.15
–4.0
–6.0
–8.0
–10.0
–12.0
0 2 4 6 8 10
Radial value in wavelengths
x
Periodic
b wall
y
Figure 10.8 Waveguide model for designing the coupling from a slot pair. IEE
1990, reprinted from [6], with permission
pffiffiffi
coupling factor and the aperture taper. The factor 1= r and the attenuation term in
(10.4) will not appear in the analysis model as they are not relevant in the rectan-
gular waveguide. Let the coupling coefficient of the nth slot pair be cn and the field
incident at the nth slot pair be En. The power radiated by the nth slot is (Encn)2. It is
easy to show that the coupling coefficient cn is related to the attenuation coefficient
by ð1 c2n Þ1=2 ¼ eaðrn rn1 Þ . If the radial spacing between adjacent slot pairs is the
same, the coupling coefficients are equal.
f X
O
Sρ
Sf
ρM Z
ρm ρ
du
dl
Slow wave Absorber
structure Coaxial cable
Figure 10.9 The slot pairs cut on concentric circles. IEEE 1988, reprinted from
[7], with permission
Current-flow line
ρ1 E
θ1
1
λg/2 E
ρ2 Desired
2
θ2 Polarisation
f 3
E
4
Figure 10.10 First two slot pairs. IEEE 1988, reprinted from [7], with
permission
polarised slot array also. The co-polarised and cross-polarised radiations for the slot
pair are given by (10.5) and (10.6).
sin q1 sinðq1 þ fÞ sin q2 sinðq2 þ fÞ ¼ 1 (10.5)
sin q1 cosðq1 þ fÞ þ sin q2 cosðq2 þ fÞ ¼ 0 (10.6)
Slot arrays in special waveguide technologies 299
and
q2 ¼ p f=2 (10.8)
The radial distance between adjacent slot pairs is kept one guide wavelength. If
the guide wavelength is kept less than the free space wavelength by dielectric
loading, grating lobes are suppressed. The azimuthal distance Sf between adjacent
slots is kept one guide wavelength, just like that of the radial line slot antenna for
circular polarisation (see Figure 10.3).
If there are MN slot pairs in this ring and if the coupling coefficient, defined in
Section 10.1.2, of each slot is bN, the incident field in the (N 1)st ring is
1=2 pffiffiffiffiffiffiffiffiffiffi
AN 1 ¼ exp jkr rN 1 1 MN b2N = rN 1 (10.10)
The number of slots in a ring is proportional to the radius of the ring. The
reflections from the slots are ignored in this analysis. The coupling coefficient bN1
of each slot in this ring is related to CN and CN1, the aperture excitation coeffi-
cients of Nth and (N 1)st rings, respectively, as given in (10.11).
pffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi 1=2
bN = rN ¼ bN 1 = rN 1 1 MN b2N ðCN =CN 1 Þ (10.11)
Recurrence relations like this are written for all rings up to the first or the
innermost ring. The wave going past the innermost ring of slots will be dissipated
in the absorber.
The value of bN is initially estimated and subsequently updated in an iterative
fashion until the power dissipated in the absorber becomes a small value such as
5% of the power incident in the Nth ring of slots.
Figure 10.11 shows the values of coupling coefficients, bi, where i ¼ 120
denote the ring number. The phase of the aperture distribution is assumed to be
300 Slotted waveguide array antennas: theory, analysis and design
0.10
0.09
0.08
Uniform
0.07 10 dB edge taper
0.06
Coupling
0.05
0.04
0.03
0.02
0.01
0.00
0 2 4 6 8 10 12 14 16 18 20
Radial value in wavelengths
Figure 10.11 Slot coupling coefficient as a function of the radial value for an
inward travelling wave in a 20 wavelength radius double-layer
radial waveguide
uniform whereas two amplitude distributions, uniform and 10 dB taper, have been
considered for the circular aperture [8]. For the outermost slots, coupling increases
as the radius decreases for both uniform and tapered apertures due to decreasing
power in the travelling wave as it propagates inward. For tapered apertures, the
coupling coefficient starts with a smaller value for the outermost ring but the rate of
increase is greater than that of the uniform distribution since the aperture excitation
increases as the radius decreases. The coupling coefficient reaches a peak and starts
decreasing for innermost rings since the ratio rn1/rn decreases rapidly for small
values of n.
The incident travelling wave amplitude in dB in each ring is shown in
Figure 10.12. For both cases the incident power in the outermost ring is the same. For
uniform distribution, the incident wave power decreases more rapidly than for the
tapered distribution as greater power is radiated by the outermost rings in the former.
Both cases were designed to dissipate 14 dB power at the centre relative to the
incident wave power at the 20th ring. Therefore both plots coalesce at the centre.
Since the slot pairs are spaced one wavelength radially, the reflections from
them add up. Therefore the reflection coefficient of this antenna type is high. It is
possible to reduce the reflection coefficient at the input port by introducing addi-
tional pairs of slots called reflection cancelling slots on the radiating wall or on the
opposite wall in the case of a single-layer parallel plate waveguide or by changing
the slot spacing in the radial direction as a function of the azimuthal angle, thereby
producing a squinted beam off broadside. In one case the squinted beam design
produced 22 dB return loss compared to 4 dB of the corresponding broadside beam
design [9]. The improved return loss caused 1 dB increase in gain for the squinted
beam design. It is convenient to produce the squinted beam in the polarisation
Slot arrays in special waveguide technologies 301
0.00
–2.00
Incident wave power in dB
–4.00
–6.00
–8.00
–10.00 Uniform
–14.00
0 2 4 6 8 10 12 14 16 18 20
Radial value in wavelengths
Figure 10.12 Incident wave power in dB as a function of the radial value for an
inward travelling wave in a 20 wavelength radius double-layer
radial waveguide
plane, for example, for an x-polarised beam pointing at an angle (qt, ft ¼ 0). The
spacing, Sr, between two adjacent slot pairs in the radial direction is given next.
lg
Sr ¼ (10.13)
1 p1ffiffiffi sin qt cos f
er
0.09
0.08
0.07
Uniform
0.06 10 dB edge taper
Coupling
0.05
0.04
0.03
0.02
0.01
0.00
0 2 4 6 8 10 12 14 16 18 20
Radial value in wavelengths
Figure 10.13 Coupling coefficient as a function of the radial value for an outward
travelling wave in a 20 wavelength radius single-layer radial
waveguide
0.00
–2.00
Incident wave power in dB
–4.00
–6.00 Uniform
–10.00
–12.00
–14.00
0 2 4 6 8 10 12 14 16 18 20
Radial value in wavelengths
Figure 10.14 Incident wave power in dB as a function of the radial value for an
outward travelling wave in a 20 wavelength radius single-layer
radial waveguide
increases, larger amounts of coupling coefficient are needed to produce the desired
aperture distribution even though there are greater number of slots per ring as the
radius increases. For the tapered distributions, the coupling coefficient is larger
than that of the uniform distribution for small radial values while it is less for large
values of the radius.
Figure 10.14 illustrates the incident wave power as a function of the radial
value. For small radial values the slot coupling coefficient and the radiated power
Slot arrays in special waveguide technologies 303
are small and hence the incident wave power is nearly constant whereas for large
radial values the incident wave power drops rapidly. For tapered distributions, the
power level is lower since greater power is radiated from the centre of the aperture.
Both plots coalesce at N ¼ 20 since they are designed to have the same amount of
power absorbed at the outer end of the parallel plate waveguide.
2=Z0
C0 ¼ (10.15)
ð1 þ Zii =Z0 Þ 1 þ Zjj =Z0
Z0 is the characteristic impedance of each feed line, and Zii and Zjj are the self-
impedances of the input ports i and j, respectively.
For the sum pattern with all four feeds excited equally (A1 ¼ A2 ¼ A3 ¼ A4 ¼ A),
the incident travelling wave is equivalent to that produced by a single centred probe
4 feeding probes at
radius = b
Figure 10.15 Feed probes for monopulse application. IEEE 1996, reprinted
from [11], with permission
304 Slotted waveguide array antennas: theory, analysis and design
feed. For the azimuth difference pattern in the xz plane (A1 ¼ A2 ¼ A3 ¼ A4 ¼ A)
the incident travelling wave field is equivalent to that produced by a centred probe
feed weighted by a cos f term. The pattern in the xz plane produces a null in the
broadside direction with a major lobe on each side. Similarly for the elevation dif-
ference pattern in the yz plane (A1 ¼ A2 ¼ A3 ¼ A4 ¼ A) the incident travelling
wave field is equivalent to that produced by a centred probe feed weighted by a sin f
term. The pattern in the yz plane produces a null in the broadside direction with a
major lobe on each side. The design of the radiating slots may use any of the methods
discussed in the previous sections.
y y
z ar x
Square plate
Feed rectangular
waveguide
Z Radiating slot
X
Parallel plate waveguide
Absorber
Feed waveguide
Dielectric Feed slot
Coaxial line
Figure 10.16 Slot array fed by a parallel plate waveguide. IEEE 1992,
reprinted from [15], with permission
l0 l0
pffiffiffiffi < af < pffiffiffiffi (10.16)
2 ef ef
In order to have a nearly uniform TEM mode to propagate in the parallel plate
waveguide, the spacing between the coupling slot pairs should be less than the
TEM mode wavelength in the parallel plate waveguide, that is, lgf < pl0ffiffiffi
er . Typically
l0 pffiffiffi
lgf ¼ pffiffiffiffi 2 (10.17)
ef
Therefore,
ef
er < (10.18)
2
The design of coupling slot pairs in the feed waveguide is carried out using the
procedure described in Chapter 6. Generally, a uniform aperture distribution is
306 Slotted waveguide array antennas: theory, analysis and design
desired in many applications. The scattering parameters of a given slot pair may be
obtained from a computational electromagnetic code such as HFSS as a function of
slot lengths and spacings. The tilt angle is kept at 45 . The spacing between slots in
a pair is nominally a quarter guide wavelength while the slots are parallel. The
lengths (non-resonant) and spacings are adjusted so that the reflection from the slot
pair is zero in the feed waveguide. For uniform aperture distribution, the coupling
into the parallel plate waveguide is the same from every slot pair and the power
going into the absorber in the feed waveguide should be small.
Figures 10.17 and 10.18 display a four-port coupler and the analysis model of a
feed slot pair, in which the parallel plate waveguide is replaced by a rectangular
waveguide with two periodic side walls. Such a model can also be used for the case
z
y Port 3
x Periodic wall
Port 1 Port 2
af
Figure 10.17 Analysis model for the feed slot pair. IEEE 1992, reprinted from
[15], with permission
Port 3
Periodic wall y
Electric wall x
#2
lf2 df
Port 1 Port 2
lf1
#1
Magnetic field
Magnetic field (TEM mode)
(TE10 mode)
Port 4
Figure 10.18 Feed slot pair in the common wall of a four-port coupler. IEEE
1992, reprinted from [15], with permission
Slot arrays in special waveguide technologies 307
y
ar
x
Port 2
Periodic
Adjacent #1
boundary
lr1
#1
#2 dr
lr2
Adjacent #2
Port 1
Figure 10.19 Analysis model for a radiating slot pair. IEEE 1992, reprinted
from [15], with permission
rectangular waveguides can be designed with a very small b dimension and are
easily transitioned to microstrip transmission lines. If the number of metallised vias
is greater than 10 per wavelength, such waveguides essentially behave like
dielectric-filled rectangular waveguides with metallic side walls. Based on a study
of dispersion characteristics of the substrate integrated rectangular waveguide, the
cut-off frequencies of the first two modes were obtained as a function of the
structural parameters, as shown in (10.19) and (10.20) [18].
1
c D2
fc10 ¼ pffiffiffiffi W (10.19)
2 er 0:95b
1
c D2 D3
fc20 ¼ pffiffiffiffi W (10.20)
er 1:1b 6:6b2
where c is the velocity of light in free space, D is the diameter and b is the spacing
of adjacent vias. W is the distance between the two arrays of vias, that is, nominally
the a dimension of the corresponding dielectric-filled waveguide. If the spacing
between vias is sufficiently small, that is, greater than 10 vias per wavelength, the
effective value of the width of the waveguide is given by
D2
Wef f ¼ W (10.21)
0:95b
Thus an SIW is equivalent to a dielectric-filled metallic rectangular waveguide
with width Weff. Then the values of offsets and lengths of the radiating slots in a
dielectric waveguide can be determined using the design procedures discussed in
Chapters 6 and 7.
An example of a planar array of slots in a two-layer SIW configuration is
shown in Figure 10.20 [19]. They used the conventional design of a slot array in
equivalent dielectric-filled rectangular waveguides. Transitions from microstrip
line to SIW using tapered lines were found to work well. In the lower layer a feed
waveguide is formed using the SIW technology with centred-inclined coupling
slots. Coupling coefficients for such slots are found using (7.2) in Chapter 7.
A single-layer SIW slot array for 60 GHz is shown in Figure 10.21 [20]. The
feed waveguide couples into radiating waveguides through coupling windows at
one end. Even though such an end-fed arrangement has bandwidth limitations, it
has the advantage of being a low-cost device because of the ease of fabrication of
the single-layer device. A conductor backed coplanar waveguide (CBCPW) is used
to feed the antenna, transitioning into a feed SIW.
SIW technology, that is, arrays of metallised vias in place of conducting walls,
has also been employed in parallel plate waveguide slot arrays [21]. The low-
temperature co-fired ceramics (LTCC) manufacturing technique is suited to make
SIW slot arrays, especially at millimetre-wave frequencies and even higher fre-
quencies in the range of hundreds of GHz, with the availability of low-loss
dielectric materials [22,23].
Slot arrays in special waveguide technologies 309
Top layer
COL6
(slot array)
Coupling
slots
Input
port
Bottom layer
(feed waveguide)
COL1 z^
x^
ROW1 ROW6
Figure 10.20 A two-layer SIW slot array at 60 GHz. Reproduced from [19],
courtesy of The Electromagnetics Academy
Power
divider
50 Ω CBCPW Radiating
slots
Substrate
CBCPW to SIW
transition Metalised
via holes
Figure 10.21 A single-layer SIW slot array at 60 GHz. IEEE 2010, reprinted
from [20], with permission
310 Slotted waveguide array antennas: theory, analysis and design
h x a
d z
w a
Figure 10.22 The geometry of the gap waveguide. IEEE 2011, reprinted from
[25], with permission
Slot arrays in special waveguide technologies 311
Figure 10.23 A linear array of four slots excited by a ridge gap waveguide with a
corporate feed. IEEE 2014, reprinted from [28], with permission
Rectangular
waveguide opening Bottom plate
with radiating slots
Corporate feed
network
Figure 10.24 A planar array of 2 2 slots excited by a ridge gap waveguide with
a corporate feed. IEEE 2014, reprinted from [28], with
permission
In the design of the bends of the T-junctions, a few pins were removed or
relocated locally. This does not affect the performance of the gap waveguide
structures. For designing the antennas without grating lobes, one row of pins has
been used at some locations. Even with one row of pins the leakage of energy to
neighbouring elements is 20 dB down.
Figure 10.24 shows a 2 2 planar array in the ridge gap waveguides. It has a
90 transition from ridge gap waveguide to rectangular waveguide. Since the ridge
312 Slotted waveguide array antennas: theory, analysis and design
z
y
x
ls
Radiating slot
ws
Top
Coupling slot layer
lc
Cavity ws
Bottom
lT layer
Ridge feed line lm
wm wT
Figure 10.25 The 2 2 array unit cell excited by a cavity. IEEE 2014,
reprinted from [29], with permission
gap waveguide propagates a quasi-TEM mode just like a microstrip line, the ridge
gap waveguide to rectangular waveguide transition is designed similar to a
microstrip line to waveguide transition.
The four-way power divider is designed by tapering the ridge sections gradually
for the impedance match. This results in 20 dB return loss over 1215 GHz band.
The 2 2 element array is excited in phase and with equal amplitude by a corporate
feed network. The element spacing is about 0.875l at 15 GHz. Both the 1 4 and
2 2 arrays produced about 20% bandwidth for better than 10 dB return loss.
Figure 10.25 shows a unit cell comprising 2 2 slots excited by a cavity [29].
The bottom layer of Figure 10.25 has a ridge waveguide terminated in a T section.
A coupling slot in the middle excites the cavity which in turn excites the four
radiating slots in the top plate. Sixteen such unit cells make up an 8 8 array of
slots in ridge gap waveguide technology with a corporate power divider as illu-
strated in Figure 10.26 [29].
The 8 8 array discussed earlier has uniform aperture distribution and pro-
duces 13 dB sidelobe levels in both the E- and H-planes. It is possible to design
tapered distributions using non-uniform power dividers and non-resonant slots of
different lengths for low sidelobe arrays. The slot arrays in ridge waveguide tech-
nology are expected to find many applications in the future, such as integration of
planar antennas, filters, active devices, etc., up to terahertz frequencies.
Slot arrays in special waveguide technologies 313
lt wq
Radiating slots
wt lq
Feeding
network
Transition to
WR-15
Figure 10.26 An 8 8 slot array in ridge gap waveguides with a corporate power
divider. IEEE 2014, reprinted from [29], with permission
References
[1] Kelly K. C. ‘Recent annular slot array experiments’. IRE National Conven-
tion Record, pt. 1, 1957, pp. 144–151.
[2] Goebels F. J., Kelly K. C. ‘Arbitrary polarisation from annular slot
planar antennas’. IRE Transactions on Antennas and Propagation. 1961;9(4):
342–349.
[3] Kelly K. C., Goebels F. J. ‘Annular slot monopulse antenna arrays’. IEEE
Transactions on Antennas and Propagation. 1964;12(4):391–403.
[4] Ando M., Sakura K., Goto N., Arimura K., Ito Y. ‘A radial line slot antenna
for 12 GHz satellite TV reception’. IEEE Transactions on Antennas and
Propagation. 1985;33(12):1347–1353.
[5] Ando M., Ito S., Sakurai K., Goto N. ‘Suppression of reflection in a radial
line slot antenna for 12 GHZ band satellite reception’. IEEE Antennas and
Propagation Symposium Digest, 1987, pp. 898–901.
[6] Hirokawa J., Ando M., Goto N. ‘Analysis of slot coupling in radial
line slot antenna for DBS reception’. IEE Proceeding H. 1990;137(5):
249–254.
[7] Ando M., Numata T., Takada J.-I., Goto N. ‘A linearly polarized radial line
slot antenna’. IEEE Transactions on Antennas and Propagation. 1988;36(12):
1675–1680.
[8] Stutzman W. A., Thiele G. A. Antenna Theory and Design, 3rd edn. John
Wiley and Sons, Hoboken, NJ, 2013.
314 Slotted waveguide array antennas: theory, analysis and design
In this chapter, we will take a look at manufacturing aspects that are important for
high-quality production of slotted waveguide array antennas. It is a wide subject so
we will concentrate on critical parameters such as mechanical tolerances that relate
to electrical performance. We will also discuss joining methods for metal materials,
for example, dip brazing of aluminium. Applications at high frequencies (milli-
metre waves) in particular call for very high precision and special methods. An
important area is the use of slotted carbon fibre reinforced plastic (CFRP) wave-
guides that offer light weight and thermal stability, typically required for large
antenna systems for space applications. The technology is also used in some mili-
tary ground and airborne radar systems.
Some of the exciting developments in low-cost fabrication using plastic
materials and metallisation techniques are discussed. This area is related to the
microelectronic area, particularly for high frequencies and highly integrated
antenna/microwave assemblies.
V0 sV0
Figure 11.1 The nominal element voltage V0 and the error voltage sV0
while the array pattern function has a modified Rayleigh distribution; see, for
example, Rondinelli [3]. The error model is represented graphically in Figure 11.1.
Assuming small errors the errors expressed in phase and amplitude become
sphase ¼ s radians
(11.2)
sdB ¼ 20logð1 þ sÞ dB
As shown here the amplitude error and the phase error are both given by the
same parameter s. This can of course be questioned since in many cases the two
types of error have different sources. Still, the model illustrates the level of influ-
ence from typical errors. For given or assumed phase and amplitude errors we can
in practical cases write
s2 ¼ s2tot ¼ s2phase þ s2ampl (11.3)
The effect on the array gain can be estimated as
G ¼ G0 = 1 þ s2 (11.4)
Normally this effect is quite small. The effect on the sidelobe level is usually
much more serious. The main beam power level can be written as
!2
XN
Gmax / Vn (11.5)
1
Example:
Number of radiating elements: N ¼ 64
Element error: 0.5 dB and 5 , uncorrelated => stot ¼ 0.11
Pattern function: SL0 ¼ 50 dB sidelobes, h ¼ 0.72 (error-free)
This gives the average sidelobe level due to the errors alone:
s2
SLerr ð hN Þ ¼ 36 dB. Sidelobe peaks will be perhaps 68 dB higher. Clearly,
this is an unbalanced design; the tapering is too strong considering the error level
(too much overdesign).
In this example the 64 elements could be the 64 slots in a single linear wave-
guide array. If we now take 32 identical slotted waveguide arrays of this kind and
combine them into a 32 64 planar array, we would still get an error sidelobe level
of about 36 dB. However, if the 32-element slotted arrays had completely inde-
pendent errors (but at the same average level as before), the sidelobe average level
due to errors becomes 15 dB lower, that is, at about the same level as SL0, a much
more reasonable design.
Figure 11.2 shows the calculated sidelobe level SL versus the error-free side-
lobe level SL0 for different numbers of uncorrelated elements N.
We have here discussed array element errors in dBs and degrees, while the real
error character is slot positions, orientations and lengths, etc., in inches and milli-
metres. Furthermore, the average sidelobe level as discussed earlier is not always
the most suitable parameter. A more relevant measure can be the maximum side-
lobe level that is expected with, say, 95% confidence. This approach was studied by
SL [dB]
–20
s = 2 dB
hN = 37
–30
s = 1 dB
hN = 70
–40
s = 1 dB
hN = 700
–50
SL0 [dB]
–60
–20 –30 –40 –50 –60
Figure 11.2 Average sidelobe level SL (achieved) versus sidelobe level SL0
(ideal) with error s as parameter. N is the number of uncorrelated
radiating elements, and h is the aperture efficiency
320 Slotted waveguide array antennas: theory, analysis and design
1.0
σ = .1 .2 .3 .4
.5 .6
0.8 .7 .8
.9 1.
0.6
Probability
0.4
0.2
Figure 11.3 Normalised probability function for sidelobe level. Both the
amplitude and the error parameter s are normalised to the design
sidelobe level. 1982 IEEE. Adapted from [4], with permission
Hsiao [4], who derived the probability density function of the sidelobe level (see
Figure 11.3).
The error tolerance problem can also be studied by direct simulation of the
antenna array in question. Lee et al. [5] used the Monte Carlo method for deter-
mining acceptable error levels. The tolerance for each element was set inversely
proportional to the nominal element weight. A Monte Carlo simulation was also
used by Rengarajan et al. [6]. For their simulation random element error levels of
0.5 dB and 5 were assumed. The authors also made a Method of Moments analysis
with assumed machining tolerances of 1 mil for slot positions, lengths and
widths. Finally, a finite element analysis using HFSS code was made before
building and testing an experimental antenna. The results from all these efforts
showed good agreement.
In two early papers [7,8], the authors treated the tolerance problem for linear
waveguide arrays with longitudinal slots. The mechanical errors considered were
slot position errors: along waveguide, across waveguide (i.e. offset errors), and slot
length errors. The errors were assumed to be independent, random and normally
distributed. The error in slot position along the waveguide was found to be of minor
importance compared to the errors in slot offset and slot length. Bailin and Ehrlich
[7] quoted a result by Stegen [9] where the change of phase of the radiated field was
measured as a function of a change in slot length off resonance. Stegen’s curve is
reproduced in Figure 11.4.
We find from Stegen’s graph that 5 degrees of phase error results from a
slot length error of about 0.5%, or less than one-tenth of one mm (X-band).
Manufacturing aspects 321
80
60
40
Relative phase/degrees
20
–20
–40
Figure 11.4 Radiated phase as a function of change in slot length off resonance.
The points on the curve indicate slot offsets, from about 0.7 to 3.9
mm. Standard X-band waveguide, slot width 1.6 mm, frequency 9.375
GHz. Adapted from Kaminov and Stegen [10]
(We know from Chapter 5 that an error in slot offset also changes the slot resonant
length as well as the slot admittance.)
The offset of a longitudinal slot is referenced to the waveguide centreline,
that is, the line halfway between the inner side walls of the waveguide. This line is
difficult to establish with high precision from the outside of the waveguide,
especially for long waveguides. If in error the result is a systematic error, that is,
a periodic offset error with a period of about one guide wavelength, giving rise
to grating lobes. The artillery locating radar described in Chapter 3 (Figure 3.5)
has a radiating aperture of frequency scanned non-resonant waveguides, each
with 64 longitudinal slots. It has been estimated that a slight offset error in the
waveguide centreline by as little as 0.04 mm would result in a 20 dB grating lobe
level [11].
Rectangular waveguide sizes are standardised (EIA, IEC)1 including toler-
ances and flange types [12]. Also ridge waveguide dimensions are standardised to
some extent. However, for slotted waveguide arrays using ridge waveguides (e.g. to
reduce the width of the waveguide) non-standard cross sections may be required.
The tolerances for the critical dimensions can be analysed theoretically using HFSS
or special software [13,14]. Figure 11.5 shows a simplified cross section of a ridge
waveguide which is characterised by four main dimensions: width (A), height (B),
1
EIA ¼ Electronic Industries Association, IEC = International Electrotechnical Commission. See also
Appendix.
322 Slotted waveguide array antennas: theory, analysis and design
ridge width (S) and the separation between the upper wall and the ridge (D). From a
theoretical analysis it was found that the separation (D in the figure) is the most
critical and should be given the tightest tolerance. This is also the region where the
electric field is concentrated.
The traditional assembly technique for joining aluminium parts is dip brazing.
The parts are kept together by fixtures, if necessary, and submerged in a bath of
special salt heated to about 600 C. This is close to the melting temperature of the
aluminium alloy. A filler material is added in the joints where it will melt and join
the assembly. After the brazing the unit is hardened and surface treated. Anderson
et al. [17] describe the design and fabrication of X- and Ku-band array antennas
using numerically controlled machining and dip brazing. Joining the various parts
by epoxy bonding has also been reported [18].
A 94 GHz linear array of 21 longitudinal slots was made in copper by photo-
etching the slots on a 0.1 mm copper sheet [19]. The sheet was then brazed onto a
copper block with the waveguide channel. The theoretical analysis indicated a
required positional accuracy for the slot positions of 0.0004 in. (0.01 mm) and even
tighter for the slot length. An array of series inclined slots was also fabricated.
In another study [20] it was shown that a linear slotted waveguide array with
ten slots can be built for 94 GHz from standard WR-10 waveguide. However, the
wall thickness was reduced to facilitate scaling from X-band data. The slots were
milled with a machine tolerance of typically 0.02 mm ( 0.001 in.).
For the machining of rectangular slots the slot ends are often rounded and a
correction for this can be applied as discussed in Chapter 5, Section 5.3.3; cf. also
Section 8.10.1.
Low-cost mass production of 20 and 60 GHz planar arrays was aimed for in an
experimental work by Ando and Hirokawa [21]. Their findings include die-casting
of waveguide channels showed promise. Laser-welding of a slotted plate to the
waveguide channels caused too much distortion while bonding using a conducting
adhesive worked well mechanically, but losses were too high. However, feeding
adjacent waveguides in antiphase (compensated with slot offset directions) was
promising since the currents in the waveguide vertical walls almost vanished.
A crucial step in the fabrication is the joining of a slot plate to the waveguide
channel(s). Screws can only be considered for low frequencies. For high fre-
quencies (millimetre waves) diffusion bonding is an interesting technique where
the parts are pressed together at a high temperature causing the parts to join (no
adhesive). Stainless steel is often used but it has high conduction loss. Copper is
preferred for antennas and was tried by [22,23]. Also, 60% efficiency was obtained
for a 94 GHz 18 18 slotted waveguide array. Alignment of the layers during
bonding is critical. See also reference [24] and Chapter 10.
Solbach [25] reported work on linear and planar arrays at 40 and 60 GHz. The
slots were photoetched on a thin (0.1 mm) copper plate. The slot lengths and widths
showed a spread of about 50100 mm, mainly due to underetching. A thinner plate
or thin metallisation on a dielectric carrier was suggested to improve this. Such a
dielectric carrier was used by Zhao et al. [26] for a ten-element slotted waveguide
array at 93 GHz. They used a teflon substrate (Rogers 5880) with a thickness of
0.254 mm. The slot plate was photoetched and bonded to the milled waveguide
channel using a conductive resin. A related technology is also used in stripline fed
slotted arrays; cf. Section 9.7.
324 Slotted waveguide array antennas: theory, analysis and design
(a) (b)
Figure 11.7 Mobile radar system: (a) radar unit; (b) array antenna with
metallised CFRP slotted waveguides [28]. Courtesy of Saab AB
Manufacturing aspects 325
steered in elevation. The horizontal slotted waveguide runs are fed in the centre
with different slot spacings in the two halves in order to maximise the frequency
bandwidth [29]. The centre feeding technique is discussed in more detail in
Chapters 7 and 9.
With composites it is possible to form several waveguides in one operation, for
example, to produce integrated subpanels; an example is shown in Figure 11.8.
When metallised this design performs electrically as its aluminium counterpart, but
with improved characteristics in terms of weight and thermal stability. The same
technique has also been used for airborne applications [30].
Low weight and thermal stability are critical parameters for space applications.
A metallised CFRP composite material was therefore chosen for the SAR antennas
for the European Earth Resource Satellites ERS-1 and ERS-2. The C-band SAR
antenna is 10 m 1 m when deployed (Figure 3.4(b)). The antennas were manu-
factured by Dornier Satellitensysteme in Germany [32]. Metallised carbon fibre
arrays have also been used for X-band active SAR systems, for example, in the
TerraSAR-X Radar satellite [33]. The active radiating element with the dual CFRP
waveguides (for dual polarisations) is shown in Figure 11.9.
Large parts of modern aircraft structures are made of CFRP panels. It has been
suggested that such panels could be designed to function as, for example, radar
antennas and loadbearing structures at the same time. Several studies along these
lines have been published [34,35], addressing RF performance and structural
performance of test panels: SWASS = Slotted Waveguide Antenna Stiffened
Figure 11.9 The active radiating element of the TerraSAR array with a
combination of edge slots in the rectangular waveguide and
longitudinal slots in the ridge waveguide, both in CFRP technology.
2003 IEEE. Reprinted from [33], with permission
326 Slotted waveguide array antennas: theory, analysis and design
Structure. The tested panels were not metallised. The reported transmission losses
were somewhat high and probably not acceptable for most applications. However,
losses depend on the orientation of the carbon fibres in the waveguides. If the
technology is successful the result could be large high-performance conformal
slotted waveguide array antennas with minimum impact on drag and weight.
11.2.3 Microfabrication
Microelectronics packaging is a key element in the production of devices and
systems that are becoming indispensable in our modern life. Miniaturisation is
taken for granted in portable electronic devices such as mobile phones, Wi-Fi
networks, automotive radar, etc. The trend towards higher frequencies makes the
design and fabrication a challenging task. Circuits on ceramic substrates have been
used extensively and efforts to combine this technology with antennas have
been reported. The low-temperature co-fired ceramics (LTCC) technology has been
studied for slotted waveguide array antennas at 79 GHz [36] and even at 140 and
270 GHz [37]. One problem is high transmission losses in the substrate material.
More about the LTCC process and the materials used in [38,39].
An eight-element slotted waveguide array antenna was developed for 300 GHz
using four silicon layers and a UV-sensitive epoxy-based photoresist (SU-8). The
contact between layers proved to be a critical factor [40]. The slot length at this
frequency is indeed small, in fact only 0.54 mm.
Photoetching of slots in a thin copper plate and brazing the plate onto a milled
waveguide channel at 94 GHz has been successful [19]. It is a less complicated process,
but lacks the possibility to integrate active and passive circuits in the same structure.
The PolyStrata (literally ‘many layers’) process (http://nuvotronics.com) has
been used for fabrication of micro-coaxial lines and components [41] and recently
also for slotted waveguide arrays at millimetre-wave frequencies, 130180 GHz
[42]. According to this process a photo resist pattern is applied onto a silicon wafer.
Next, a copper layer is electroplated on the wafer. Additional photoresist is applied
(defining the waveguide walls) and a copper layer is again deposited, etc. About 15
layers can be built up this way. The final layer includes the slot elements. At the
last step the photoresist is rinsed away through release holes. For corrosion pro-
tection the copper surface can be gold plated.
The PolyStrata process is illustrated schematically in Figure 11.10 for fabrica-
tion of a micro-coaxial line using five layers [43]. (For a slotted waveguide array no
centre conductor is needed.) Two fabricated slotted waveguide arrays for 130180
GHz with 10 and 20 elements, respectively, are shown in Figure 11.11. The technol-
ogy has been applied to slotted waveguide arrays operating up to 300 GHz [44].
The PolyStrata process is very promising for low-cost fabrication of arrays
and components at very high frequencies. Another possible fabrication method in
the future might be the 3D printing technique, so-called ‘Additive Manufacturing’
[45,46].
Manufacturing aspects 327
S1 S2 S3 S4
S5 S6 S7 S8
Photoresist
Dielectric
Copper
Silicon
Figure 11.11 The ends of two PolyStrata waveguide arrays (130180 GHz).
Note the small holes that, together with the slots, are used for
releasing the lossy photo resist. IEEE 2012. Reprinted from [42],
with permission
328 Slotted waveguide array antennas: theory, analysis and design
References
[1] Ashmead D. ‘Optimum design of linear arrays in the presence of random
errors’. IRE Transactions on Antennas and Propagation. 1952;4(1):81–92.
[2] Elliott R. S. ‘Mechanical and electrical tolerances for two-dimensional
scanning antenna arrays’. IEEE Transactions on Antennas and Propagation.
1958;6(1):114–120.
[3] Rondinelli L. A. ‘Effects of random errors on the performance of antenna
arrays of many elements’. IRE International Convention Record. 1959;
1(Part 1):174–189.
[4] Hsiao J. K. ‘Constraints on low sidelobe phased array’. IEEEAntennas and
Propagation International Symposium, Albuquerque, NM, USA, 1982,
pp. 687–690.
[5] Lee J., Lee Y., Kim H. ‘Decision of error tolerance in array element by the
Monte Carlo method’. IEEE Transactions on Antennas and Propagation.
2005;53(4):1325–1331.
[6] Rengarajan S., Zawadski M. S., Hodges R. E. ‘Waveguide-slot array antenna
designs for low-average-sidelobe specifications’. IEEE Antennas and
Propagation Magazine. 2010;52(6):89–98.
[7] Bailin L. L., Ehrlich M. J. ‘Factors affecting the performance of linear arrays’.
IRE Transactions on Antennas and Propagation. 1952;1(1):85–106.
[8] O’Neill H. F., Bailin L. L. ‘Further effects of manufacturing tolerances on
the performance of linear shunt slot arrays’. IRE Transactions on Antennas
and Propagation. 1952;4(1):93–102.
[9] Stegen R. J. ‘ Slot radiators and arrays at X-band’. IRE Transactions on
Antennas and Propagation. 1952;1(1):62–84.
[10] Kaminov I. P., Stegen R. J. Waveguide slot array design, Hughes Aircraft
Company, Techn. Memo No. 348, July 1954 (available as NTIS Report AD
63600).
[11] Karlsson I. ‘Applications of waveguide arrays in commercial and military
radars’. Antenna Applications Symposium, Allerton Park, IL, USA, 1993,
pp. 1–22.
[12] Brady M. M. ‘A plea for clarity in waveguide designation’. IEEE Transac-
tions on Microwaves, Theory and Techniques. 1965;13(2):247–249.
[13] Conciauro G., Bressan M., Zuffada C. ‘Waveguide modes via an integral
equation leading to a linear matrix eigenvalue problem’. IEEE Transactions
on Microwaves, Theory and Techniques. 1984;32(11):1495–1504.
[14] Falk K. ‘Admittance of a longitudinal slot in a ridge waveguide’. IEE Pro-
ceedings H - Microwaves, Antennas and Propagation. 1988;135(4):263–268.
[15] Sikora L., Womack J. ‘The art and science of manufacturing waveguide
slot-array antennas’. Microwave Journal. 1988;31(6):157–162.
[16] Muhs H. P. ‘Mm-wave antenna’. Microwave Journal. 1985;28(7):91–194.
[17] Anderson T., Yun-Li-Ho, Michalski J. ‘Design and manufacturing tech-
niques for planar slot array antennas for a variety of radar application’.
Proceedings of the IEEE International Radar Conference Radarcon, 1998,
pp. 337–342.
Manufacturing aspects 329
[18] Singh A. K. ‘A low cost, low side lobe and high efficiency non-orthogonally
coupled slotted waveguide array antenna for monopulse radar tracking’.
IEEE Antennas and Propagation International Symposium, Washington DC,
USA, 2005, pp. 732–735.
[19] Rao B. R. ‘94 Gigahertz slotted waveguide array fabricated by photolitho-
graphic techniques’. IEEE Antennas and Propagation International Sympo-
sium, Houston, TX, USA, 1983, pp. 688–691.
[20] Farrar F. G. ‘Millimeter wave W-band slotted waveguide antennas’. IEEE
Antennas and Propagarion International Symposium, Los Angeles, CA,
USA, 1981, pp. 436–439.
[21] Ando M., Hirokawa J. ‘Single-layer slotted waveguide arrays for DBS
reception and higher frequency applications’. Electromagnetics. 1999;19(1):
23–48.
[22] Hirokawa J., Zhang M., Ando M. ‘Millimeter waveguide fabrication to reduce
transmission loss by diffusion bonding, light-curing resin or dielectric par-
tially-filling’. Asia-Pacific Microwave Conference, APMC. 2008, pp. 1–4.
[23] Zhang M., Hirokawa J., Ando M. ‘Fabrication of a slotted waveguide array
at 94 GHz by diffusion bonding of laminated thin plates’. IEEE Antennas
and Propagation Society International Symposium, Charleston, SC, USA,
June 2009, pp. 1–4.
[24] Ando M., Zhang M., Lee J., Hirokawa J. ‘Design and fabrication of milli-
meter wave slotted waveguide arrays’. 4th European Conference on Anten-
nas and Propagation (EuCAP), Barcelona, Spain, 2010, pp. 1–6.
[25] Solbach K. ‘Some millimeter-wave slotted array antennas’. 14th European
Microwave Conference, Liege, Belgium, 1984, pp. 181–186.
[26] Zhao G., Xu H., Chen Z., Sun H., Lv X. ‘A W-band waveguide slot array
antenna based on print technology’. International Conference on Microwave
and Millimeter Wave Technology (ICMMT), 2012, Vol. 3, pp. 1–4.
[27] Jönsson D. ‘Metallisation methods & applications’. Proceedings of the ESA
Symposium on Space Applications of Advanced Structural Materials,
ESTEC, Noordwijk, March 1990, pp. 35–38.
[28] Dahlsjö O., Ljungström B., Magnusson H. ‘Fibre-reinforced plastic compo-
sites in sophisticated antenna designs’. Ericsson Review. 1987;64(2):50–57.
[29] Karlsson E. R. ‘Waveguide element for an electronically controlled radar
antenna’. US Pat. 4788 552, 1988.
[30] Brunzell S., Magnusson H. ‘Lightweight CFRP waveguide array’. Pro-
ceedings Military Microwaves, London, UK, October 1984, pp. 305–309.
[31] Josefsson L., Derneryd A., Lagerlöf R. ‘Electronic scanning with slotted
waveguide arrays’. Proceedings of the Radarcon 90 Conference, Adelaide,
Australia, April 1990.
[32] Petersson R., Bonnedal M. ‘The slotted waveguide arrays of the European
remote sensing satellites ERS-1 and ERS-2’. Electromagnetics. 1999;19(1):
77–89.
[33] Stangl M., Werninghaus R., Zahn R. ‘The TerraSAR-X active phased array
antenna’, IEEE International Symposium on Phased Array Systems and
Technology, Boston, MA, USA, October 2003, pp. 70–75.
330 Slotted waveguide array antennas: theory, analysis and design
In this book we have analysed the design of several types of slotted waveguide
array antennas, from theories and optimisation to applications and manufacturing
techniques.
In this last chapter we will discuss the current status in the field and look at
new technologies and new applications recently presented and researched. Still, in
this short overview it is not possible to mention all the details of the evolving
field; the reader is referred to the respective chapters and the references for more
information.
SAR stands for synthetic aperture radar, a technology used in airborne and space-
borne systems. By coherent signal processing along the flight path a large
(synthetic) aperture is created in this direction. Together with a large physical
aperture across the flight path very high angular resolution is possible. In
one application [26] scatterometer antennas provide ocean wind data and soil
moisture content. The technology is used in many remote sensing satellites such as
the C-band instruments ERS-1 and ERS-2; cf. Figure 3.4(b). The ERS satellites
were operational almost 20 years [27]. The major antenna technology for these
systems is slotted waveguide arrays. New developments include dual polarisation,
dual and triple beams, beam scanning, X-band and higher frequencies, and also
Outlook for the future 333
millimetre waves. The new spaceborne SAR systems will provide improved high-
resolution images on a global scale for monitoring the environment, wind speed,
ocean currents, land mapping, climate change, etc. Together with advanced signal
processing techniques, digital beam forming and interferometry, the available
information content will increase at least one order of magnitude [28].
12.4 Communication
12.5 Manufacturing
The challenges are to reduce cost and complexity. We have come a long way, for
example, in integrating the feeding waveguides in the same plane as the planar
slotted waveguide array [32]. Traditionally, the antenna used to be a separate unit,
connected to the transmit/receive front end via coaxial cables or waveguides. At
higher frequencies the active front end can be integrated with the slotted waveguide
antenna using the same materials and processes, for example, LTCC, the SU-8
process or the PolyStrata process [22,33]. One example is the LTCC substrate-
integrated slotted waveguide array antenna. It is essential to master the tolerance
requirements and possibly modify the design to match the available fabrication
capability, especially at millimetre-wave bands [34].
12.6.1 HFSS
HFSS stands for high-frequency structural simulator, originally developed by
Professor Z. Cendes around 1980 [39]. The code was further developed by Ansoft
Corporation and is today marketed by ANSYS [40]. It is a FEM and includes an
integral equation solver (MoM) which is convenient for slot array problems.
12.6.2 XFdtd
The XFdtd by Remcom [41] uses the FDTD technique, first proposed in 1966 [42].
The problem geometry is simulated by a mesh which can be directly connected to
lumped circuit elements. This has become attractive for simulations of sub-
assemblies that include both MMIC circuits and antenna parts.
12.6.4 WASP-NET
WASP-NET [44] combines several methods: MM, the FEM and the MoM for
design of slotted waveguide arrays including feed networks.
Outlook for the future 335
12.6.5 SWANTM
The software SWANTM [45] is a powerful CAD tool for the design and analysis of
very large slotted waveguide arrays. It is based on Elliott’s theory with a number of
significant extensions and improvements. It has a graphical user interface so that
the user can input design and analysis parameters easily.
References
[1] Grabherr W., Huder B. ‘Omni-directional slotted waveguide array antenna’.
29th European Microwave Conference, Munich 1999, pp. 134–137.
[2] Ekengren B. ‘Ericsson SLAR – airborne radar for maritime surveillance’.
Ericsson Review. 1979;56(4):151–157.
[3] Alves A. M. P., Duplat D. N., de Oliveira L. P., Hernández-Figueroa H. E.
‘Aerotransported radar antenna for oil spills monitoring antennas and
propagating’. 4th European Conference on Antennas and Propagation
(EuCAP), Barcelona, April 2010, pp. 1–3.
[4] Morwing B. ‘AGA-ERICON – a marine radar beacon’. Ericsson Review.
1981;58(4):180–187.
[5] Anderson T., Yun-Li-Hou, Michalski J. ‘Design and manufacturing techni-
ques for planar slot array antennas for a variety of radar application’.
Proceedings of the IEEE Radar Conference, Radarcon 98, Dallas, TX, USA,
May 1998, pp. 337–342.
[6] Brookner E. ‘Phased arrays and radars – past, present and future’. Microwave
Journal. 2006;49(1):1–12.
[7] Kopp C. ‘Evolution of AESA radar technology’. Microwave Journal.
2012;55(8):3–12.
[8] Pendergast S. L. ‘Recent advances in radar technology’. Microwave Journal.
2015;58(9):1–14.
[9] Karlsson I. ‘Applications of waveguide arrays in commercial and military
radars’. Proceedings of the 1993 Antenna Applications Symposium, Allerton
House, IL, USA. pp. 1–22, 1993.
[10] EMS White Paper Passive phased arrays for radar antennas. EMS Tech-
nologies Inc., Norcross, GA, USA, December 2005.
[11] Lewis D. J., Lee J. R., McCarty D. K. ‘A single-plane electronically scanned
antenna for airborne radar applications’. Phased-Array Antenna Symposium,
Polytechnic Institute of Brooklyn, Farmingdale NY, 2–5 June 1970.
[12] Coburn W., Litz M., Miletta J., et al. A Slotted-Waveguide Array for
High-Power Microwave Transmission. Army Research Laboratory, Rpt. MD
20783-1197, January 2001.
[13] Baum C. E. Sidewall Waveguide Slot Antenna for High Power. Sensor and
Simulation Note No. 503, Univ. New Mexico. August 2005.
[14] Callus P. J., Nicholson K. J., Bojovschi A., Ghorbani K., Baron W., Tuss J.
‘A planar antenna array manufactured from carbon fibre reinforced plastic’.
336 Slotted waveguide array antennas: theory, analysis and design
Standard rectangular waveguides are listed with their common designations and
useful frequency ranges. The waveguide inner dimensions are given in mm.
References
[1] Brady M. M. ‘A plea for clarity in waveguide designations’. IEEE Transac-
tions on Microwave Theory and Techniques. 1965;13(2):247–249.
[2] Lewin L. ‘Letter symbols to designate microwave bands’. IEEE Transactions
on Microwave Theory and Techniques. 1964;12(5):551.
[3] https://en.wikipedia.org/wiki/waveguide_(electromagnetism)#References.
Retrieved 25 Oct 2017.
Index