Corrugated Horns For Microwave Antennas

Download as pdf or txt
Download as pdf or txt
You are on page 1of 243

ELECTROMAGNETIC WAVES SERIES 18

Corrugated horns
for microwave
antennas
P. J. B. Clarricoats and A. D. Olver

Peter Peregrinus Ltd.


on behalf of the Institution of Electrical Engineers
IEE Electromagnetic Waves Series 18

Series Editors: Professor P.J.B. Clarricoats


E.D.R. Shearman and J.R. Wait

Corrugated horns
for microwave
antennas
Previous volumes in this series

Volume 1 Geometrical theory of diffraction for


electromagnetic waves
Graeme L. James
Volume 2 Electromagnetic waves and curved
structures
Leonard Lewin, David C. Chang and
Edward F. Kuester
Volume 3 Microwave homodyne systems
Ray J. King
Volume 4 Radio direction-finding
P J . D . Gething
Volume 5 ELF communications antennas
Michael L. Burrows
Volume 6 Waveguide tapers, transitions
and couplers
F. Sporleder and H.G. Unger
Volume 7 Reflector antenna analysis
and design
P J . Wood
Volume 8 Effects of the troposphere on radio
communications
Martin P.M. Hall
Volume 9 Schuman resonances in the
earth-ionosphere cavity
P.V. Bliokh, A.P. Nikolaenko and
Y . F . Filippov
Volume 10 Aperture antennas and diffraction
theory
E.V. Jull
Volume 11 Adaptive array principles
J.E. Hudson
Volume 12 Microstrip antenna theory and design
J.R. James, P.S. Hall and C. Wood
Volume 13 Energy in Electromagnetism
H.G. Booker
Volume 14 Leaky feeders and subsurface radio
communications
P. Delogne
Volume 15 The Handbook of Antenna Design
Volume 1
Editors; A.W. Rudge, K. Milne,
A . D . Olver, P. Knight
Volume 16 The Handbook of Antenna Design
Volume 2
Editors: A.W. Rudge, K. Milne,
A.D. Olver, P. Knight
Volume 17 Surveillance Radar Performance
Prediction
P. Rohan
Corrugated horns
for microwave
antennas
RJ. B. Clarricoats and A. D. Olver

Peter Peregrinus Ltd


On behalf of The Institution of Electrical Engineers
Published by: Peter Peregrinus Ltd., London, UK.

© 1984: Peter Peregrinus Ltd.

All rights reserved. No part of this publication may be reproduced,


stored in a retrieval system or transmitted in any form or by any
means — electronic, mechanical, photocopying, recording or otherwise
without the prior written permission of the publisher.

While the author and the publishers believe that the information and
guidance given in this work is correct, all parties must rely upon their own
skill and judgment when making use of it. Neither the author nor the
publishers assume any liability to anyone for any loss or damage caused
by any error or omission in the work, whether such error or omission is
the result of negligence or any other cause. Any and all such liability is
disclaimed.

Clarricoats, P. J. B.
Corrugated horns for microwave antennas.
—(IEE electromagnetic waves series; 18)
1. Microwave antennas 2. Antennas, Reflector
3. Antenna feeds
I. Title li.Olver, A. D. III. Series
621.381'33 TK7871.6

ISBN 0 86341 003 0

Printed in England by Short Run Press Ltd., Exeter


Contents

Preface viii

Acknowledgments ix

1 Introduction 1
1.1 Historical observations 1
1.2 Organisation of the book 3

2 Introduction to hybrid-mode feeds 5


2.1 Illumination of a reflector antenna 5
2.2 Focal field of a reflector 10

3 Propagation and radiation characteristics of cylindrical corrugated


waveguides 20
3.1 Introduction 20
3.2 Propagation in cylindrical corrugated waveguide 22
3.3 Fields in cylindrical corrugated waveguide 35
3.4 Attenuation in corrugated waveguide 40
3.5 Radiation from corrugated waveguides 43
3.5.1 Introduction 43
3.5.2 Radiation from a circular aperture 45
3.6 Junctions in corrugated waveguides 50
3.7 Multimode corrugated waveguides 54

4 Propagation and radiation characteristics of conical corrugated wave-


guides 58
4.1 Introduction 58
4.2 Modal characteristics of corrugated conical horns 58
4.3 Radiation from corrugated conical horns by the spherical wave
expansion method 65
4.3.1 The spherical wave expansion method 65
4.3.2 Frequency dependence of crosspolar radiation field 68
vi Contents

4.4 Generation of higher order modes 69


4.4.1 Scattering at an abrupt discontinuity 71
4.4.2 Higher order mode generation at the throat 73
4.4.3 Mode conversion along a horn with constant flare angle 73
4.4.4 Mode conversion along a horn with variable flare angle 78
4.5 Radiation from corrugated horns by the Kirchhoff—Huygen
method 86
4.6 Comparison of predicted and measured radiation patterns 90
4.7 Radiation from corrugated horns by cylindrical waveguide
approximation 90
4.8 Radiation from conical corrugated horns by means of the
Laguerre— Gaussian expansion method 95

5 Design of cylindrical and conical corrugated horns 97


5.1 Introduction 97
5.2 Copolar radiation characteristics 100
5.3 Crosspolar radiation characteristics 114
5.4 Flare section of horn 122
5.5 Throat region of horn 128
5.6 Gain and directivity 135
5.7 Efficiency when used as a feed for a reflector 137
5.8 Horn design 141
5.8.1 Large aperture, narrow flare angle horns 141
5.8.2 Small aperture horns 143
5.8.3 Wide flare angle horns 145
5.8.4 Multimode horns 150
5.8.5 Broadband horns 153
5.8.6 Multifrequency horns 154
5.8.7 Compact horns 155
5.9 Numerical prediction of performance 156
5.9.1 Propagation characteristics 157
5.9.2 Radiation characteristics 160

6 Manufacture and testing of corrugated horns 163


6.1 Manufacture of corrugated horns 163
6.2 Testing of corrugated horns 169
6.2.1 Introduction 169
6.2.2 Types of test range 171
6.2.3 Anechoic chamber design for low crosspolarisation
horn measurement 172
6.2.4 Assessment of test range performance 175
6.2.5 Measurement of copolar and crosspolar patterns 177
6.2.6 Measurement of phase patterns and gain 180
Contents vii

7 Rectangular and elliptical corrugated horns 181


7.1 Introduction 181
7.2 Rectangular corrugated horns 181
7.2.1 Background 181
7.2.2 Two-walled rectangular corrugated waveguide 183
7.2.3 Four-wall rectangular corrugated waveguide 188
7.2.4 Radiation characteristics of rectangular corrugated
horns 190
7.3 Elliptical corrugated horns 193

Appendix 199

Bibliography 213

Programs 222

Index 227
Preface

The idea for a text on the theory and design of corrugated horns became apparent
to us during the late 1970s, as this type of feed became more and more widely
used in antennas. Generally, the corrugated horn offers the reflector antenna
designer the potential for producing antennas with higher efficiency, lower cross-
polarisation and lower sidelobes, and it can be fairly said to lie at the heart of most
optimally designed reflector antennas.
The book contains material which should be of value both to the designer and
the research worker. Parts have been presented previously by us, at various vacation
schools and in invited papers at conferences, over the last ten years, but the text
also contains many new results. Some of these are drawn from the theses of our
research students and from the work of others in our Electromagnetics Applications
Group at Queen Mary College in the University of London. To these people we
offer our warmest thanks; they are separately identified in the acknowledgments
which follows. A description of the organisation of the book is to be found in
Chapter 1, alongside a brief history of the subject.
Acknowledgments

The authors are indebted to many who have helped in the production of this text.
They would especially like to thank the following members of the Queen Mary
College Electromagnetics Applications Group, whose theoretical and experimental
contributions over the period 1968-83, have proved so valuable: Drs. Al-Hariri,
Chan, Chong, Elliott, Hockham, Mahmoud, Parini, Poulton, Saha, Salema and
Seng. They also thank Mr. Kolb who undertook many measurements in our Antenna
Laboratory; Messrs Ede, Fairbrass and Goose, whose excellent craftsmanship
produced many intricate feeds, and Miss Sandra O'Callaghan, who typed the
manuscript with great care and attention to detail. One of us (PJBC) is grateful
for the encouragement of his wife Phyl during the preparation of the text.
Chapter 1

Introduction

1.1 Historical observations

There are two main reasons for the existence of corrugated horns as feeds for
reflector antennas. First, they exhibit radiation pattern symmetry, which offers
the potential for producing antennas with high gain and low spillover; secondly,
they radiate with very low crosspolarisation, which is essential in dual-polarisation
systems. The former property provided the motivation for Kay [88, 89] who
conceived the wide-angle corrugated horn in 1962 while working at TRG in the
USA. It is said that he was studying the effect of quarter-wavelength chokes at
the horn aperture and found that, by adding more than one, the pattern symmetry
improved. He noted invariance of the chokes to field orientation and thus coined
the term scalar feed to describe his horn. At almost the same time, and quite
independently, Minnett and Thomas [101], at CSIRO in Australia, pursued the
radiation properties of corrugated waveguide feeds for use in radiotelescopes.
They were motivated by efficiency and polarisation considerations. In this latter
respect, Rumsey [134] contributed during a stay at CSIRO in 1966, by observing
general properties of hybrid modes in respect of polarisation purity.
Slightly later another but related, hybrid-feed, the dielectric cone or dielguide,
was devised by Barlett and Mosely [179, 180] at Radiation Incorporated in USA.
This feed, which has some properties similar to optical waveguides, radiates after
the manner of corrugated horns.
It is said that Barlett and Mosely's discovery was made while experimenting with
a fibre glass tube in order to support a sub-reflector in a Cassegrain antenna.
Apparently, the introduction of expanded foam into the tube was found to improve
the feed performance. A qualitative description of the operation was provided in
their patent applications which followed. A quantitative description came from
Clarricoats and Salema [182], who made a comprehensive study in the early
seventies. The authors' group at Queen Mary College was to the fore in analysing
the corrugated horn and waveguide, beginning their investigations in 1968 and
continuing them to the present day. They were first to publish a theory for the
wide-angle corrugated horn, although Viggh [154] at TRG had made a contemporary
2 Introduction
study which appeared as an internal report in 1969. Minnett and his colleagues
[102, 145] at CSIRO were also working on the wide-angle horn at about the
same time as was Jeuken and his colleagues [80, 82, 85, 86], at the Technical
University of Eindhoven.
In India, Narasimhan [107, 121] produced useful approximate methods of
design suitable for those with limited computer resources. It appears that Bryant
[19, 20] was the first to publish an account of rectangular corrugated horns,
describing work undertaken by him at the Plessey Company in England, in the
late sixties. Peters and his colleagues [57, 58, 92, 100, 143] at Ohio State Uni-
versity analysed relevant corrugated surfaces but in general the rectangular
corrugated horn is not susceptible to closed-form analysis in the way that horns
of circular cross-section can be treated. Thus, there is a great paucity in the literature
of descriptions of feeds of this class and all early designs were empirical. Bryant
commented in his paper that corrugated horns might have high attenuation, and
this remark prompted Clarricoats and Saha [29] to examine the attenuation
properties of circular corrugated waveguides. Contrary to Bryant's suggestion,
these waveguides exhibit, for the lowest order hybrid HEu mode, an attenuation
even lower than that of the TEOi mode in a smooth-wall waveguide of comparable
size. During the period 1970-75, a very detailed theoretical and experimental
study of the attenuation properties of corrugated waveguides was conducted by
the present authors and their reserach students. Although, all the basic properties
were confirmed, it proved very difficult to manufacture corrugated waveguides
in long lengths and the project was terminated around 1976. However, the very
detailed investigation which included full space-harmonics analysis of the periodic
waveguide, provided an excellent base for subsequent investigations of both the
propagating and radiation properties of corrugated horns used as feeds.
By 1975, interest in dual-polarised feeds was growing under the influence of
the satellite communication market. This provided a major impetus for further
work on corrugated feeds. Tracking requirements also prompted work on dual-
mode feeds and mention should be made in this selective history of the substantial
contribution from workers at ERA Technology Limited. Two important feed
designs were developed within their group; the matched-feed, for use with offset
reflectors, conceived by Rudge and Adatia [189, 190], and the dual-depth cor-
rugated horn investigated by Ghosh [63, 64].
Around 1975, satellite broadcasting led to shaped beam requirements which
stimulated research on elliptical corrugated waveguides. There the work of Jeuken
and Vokurka [78, 83] is noteworthy. However, the difficulties of analysis and
manufacture have limited application of this work.
We are now close to the present day, and many names have been missed
from this brief account, although their valuable contributions will be identified
as the reader delves further into our text. However, before concluding this section
we should comment on the numerous contributions from Thomas, James and their
colleagues [74-77, 101, 102, 144-152], at CSIRO, whose interest in the subject
of corrugated horns has traversed the same period as the present authors, and whose
Introduction 3

work has been widely reported in the literature. Also Dragone [53—56] at Bell
Telephone Laboratories, whose 1977 papers provided substantial added insight.

1.2 Organisation of the book

Chapter 2 lays the foundation for the chapters which follow, by exploring the
desired properties of reflector antennas, both from the standpoint of radiation
patterns and focal fields. These studies reveal the optimum qualities of hybrid
mode feeds of which the corrugated horn provides the most important example.
Chapters 3 and 4 present in some detail the theory underlying the propagation
and radiation characteristics of, respectively, cylindrical and conical corrugated
horns. The former is amenable to exact analysis although in many cases it is
sufficient to use the surface-impedance approximations to describe the boundary
conditions at the corrugated waveguide wall. By contrast, there is no exact
formulation for the conical corrugated horn, but, notwithstanding, very accurate
prediction methods have been developed using, in particular, a spherical mode
treatment which extends to hybrid modes, the pure-mode descriptions found for
example, in the text by Harrington [183J. Here the authors have drawn heavily
on the work of their colleague Mahmoud [95].
Chapter 5 leads us in to design. It builds on the work of Chapters 2, 3 and 4,
but, if desired, it can be read almost independently of them. Mathematics appears
sparingly so this chapter should appeal strongly to the person who has to build
horns and does not have the time to cross all the theoretical bridges of earlier
chapters. Chapter 5 draws heavily on experimental results and computer
programs from QMC (two of which are included). The authors recognise that
many others will have obtained similar results. There are obvious advantages to be
gained in a text when results are drawn mainly from one source.
Chapter 6 addresses the problem of horn manufacture. It is to be hoped that
our observations will prove helpful but we shall be glad to learn from readers of
any techniques which they have found satisfactory and which we have not reported.
Chapter 6 also explores methods for the measurement of corrugated horns.
As stated there, these are essentially the methods for the precision measurement
of any small antenna except that the corrugated horn does exhibit some rather
special properties. As these properties are what the customer pays for, e.g. low
cross-polarisation, low sidelobes and pattern symmetry, it is understandable that
he will want to know rather precisely how well his horn performs. Chapter 6
provides answers which will also serve the interests of those seeking a general
introduction to microwave antenna measurements. More information is to be
found, for example, in The Handbook of Antenna Design.
Chapter 7 explores the characteristics of corrugated horns of rectangular and
elliptical cross-section. In contrast with the circular case the non-circular corrugated
waveguide is not amenable to exact analysis. However, approximate methods have
been tried with limited success and the chapter presents a summary with pertinent
results. These feeds are difficult to manufacture and their performance is generally
4 Introduction
inferior to their circular counterpart. They are used only in those applications
where a primary feed pattern is needed with differing beamwidths in the principal
planes.
The book has an appendix in which the space-harmonic analysis of the circular
corrugated waveguide is presented in detail.
Before concluding this introduction we must comment on certain feeds which
have not been included in our text. First, the Potter horn [188] in which a hybrid
field is synthesised from pure modes in a multimode circular waveguide. The
prototype employed two modes, the TE U and TM U , but subsequent versions
were considered in which additional modes were employed. By the nature of
its operation the structure is narrow band but it has the merit of relative simplicity
of construction. It is sometimes used as an alternative to corrugated horns where
bandwidth and beamwidth considerations allow. An account of the Potter horn
is also to be found in papers collected in the reprint volume edited by Love [94].
Another group of feeds which are not considered are those in which the
corrugations form chokes exterior to the primary pure-mode circular waveguide.
Feeds of this kind have been described by Koch [91], WohUeben, Mattes and
Lochner [164], and also Scheffer [136]. A prototype structure, the 90° flare-
angle corrugated horn, was analysed by Hockham [71], but he considered only
the case where the plane of the primary waveguide coincided with that of the
corrugations. Some features present in the other feeds were apparent but not all
the details which made them of practical importance as primary feeds for
paraboloid antennas. Unfortunately, the general case in which the chokes do not
lie in the plane of the primary waveguide represents a most difficult electro-
magnetic problem that awaits solution. These structures are only related to the
hybrid feeds discussed in this text because they excite spherical hybrid modes
in the vicinity of the radiator. The feeds described herein all support hybrid modes
as propagating fields before radiating at their apertures.
Finally, our text ends with a bibliography which attempts to be comprehensive
at the time of writing.
Chapter 2

Introduction to hybrid-mode feeds

2.1 Illumination of a reflector antenna

The most important use in antennas for a hybrid-mode waveguide or horn is as a


feed for a reflector, so we begin this chapter by considering, in general terms,
how the feed influences antenna performance. Fig. 2.1 shows a photograph of a
horn feed in a Cassegrain antenna forming part of a satellite-communication earth-
station while Fig. 2.2 shows a dielectric-cone feed in an antenna for a similar
application. Although different in appearance and construction, the corrugated
horn and dielectric-cone feed support nearly identical hybrid fields and in this
chapter we shall use the term hybrid-mode feed to cover both types.
The antennas shown in Figs. 2.1 and 2.2 produce pencil beams, and to be
effective, their gain should be as high as possible, consistent with an acceptable
sidelobe envelope. Furthermore, if, as is now usual, the antenna is required to
transmit and receive orthogonally polarised waves, there will also be an exacting
specification on the crosspolar radiation pattern. The antenna gain in the boresight
direction G, is related to the aperture area ,4 and wavelength through

G =f , (2.1)
where t}, the antenna efficiency, is the product of a number of terms identified
in Table 2.1. The table shows how the first four terms depend on the feed radiation
pattern. If the feed radiation pattern is symmetric, i.e. the £-plane and if-plane
patterns are identical, 7?x = 1 and furthermore, if the feed has a common phase
centre for both planes and is properly located, 7jp = 1. This leaves the illumination
and spillover losses as the principal causes of gain degradation due to the feed and
even in an ideal antenna there must be a compromise between these last two
factors. If the illumination is uniform 7?j = 1 but then the spillover efficiency r?s,
for any realisable feed, would be too low. Not only would this reduce the antenna
gain but the antenna sidelobe levels and antenna noise-temperature would be too
high giving rise to problems both in earth-station and microwave-relay applications.
In passing we note that by using a dual-shaped reflector design, it is possible to
Introduction to hybrid-mode feeds

Fig. 2.1 Intelsat type B earth-station antenna with a narrow flare-angle corrugated horn
(Courtesy Andrew Antennas)

Fig. 2.2 Reflector with dielectric cone feed


Introduction to hybrid-mode feeds 7

achieve the best possible compromise between these two factors since by controlling
the subreflector shape, the illumination can be made nearly uniform over most
of the main reflector while being heavily tapered near to the reflector edge.
We can see from the above that an ideal feed for a symmetric reflector antenna
is one which has a symmetric radiation pattern and this also ensures zero cross-
polarisation. Many different reflector geometries exist and since systems operate
over quite wide bandwidths, sometimes in two or more frequency bands, we
require our feeds also to have wide bandwidths, as well as a variety of beamwidths,
corresponding to different reflector f/D ratios. Only the hybrid-mode feed offers
a means to approach all these objectives simultaneously.

Fig. 2.3 Transverse field patterns for dominant HEn mode in cylindrical waveguide of radius

To emphasise this statement let us examine, for the dominant mode, the electric
field in the aperture of a hybrid-mode waveguide, such as that of Fig. 2.3. The
derivation of the fields will come later. The representation is appropriate to a
corrugated waveguide of internal radius rx. It is also a good approximation for a
dielectric waveguide of radius ru if krx > 1 (aperture large compared to a wave-
length) for which the field is then mainly confined to the dielectric region r < rx.

E = J2(Kr)(cos2(Pix + sin 2cpiy) (2.2)

Jn(Kr), n = 0, 2 is a Bessel function of the first kind and order n, ixy are unit
vectors in the x and y directions, Uo is a normalised transverse wavenumber, K
and k are transverse and free-space wavenumbers, respectively, X and Y are the
normalised reactance and admittance of the boundary at r — rx. In Fig. 2.3, the
boundary is that of a circular cylinder so,
1/2

X = - j - ~ j (2.3)

(2.4)

and the fields in eqns. 2.3 and 2.4 are to be evaluated at the boundary at r =
8 Introduction to hybrid-mode feeds

Table 2.1 Components of efficiency of reflector antennas


Illumination efficiency

2 fdo°(\FE\+ \FH\2)sin8dd
Spillover efficiency

Phase error efficiency

•nP =

Crosspolarisation efficiency

Blockage efficiency
+ \FH\) tan (6/2) d6\2
B

Surface error efficiency


7?E = exp [— (4ne/X)2] e = RMS surface error
Total efficiency

In the above, the feed radiation pattern is represented by


e-ikp

E = — — (F E sin <t> + F H cos (p)

FE and F H are the E plane and H plane amplitude patterns, respectively. P is the
distance from the feed to the reflector surface.

From eqn. 2.2 we see that if the term (X — Y) vanishes, the aperture field is
independent of the angular variable 0 and is also free of crosspolarisation.
Furthermore, because the radiated field is the Fourier Transform of the aperture
field, these properties are transferred to the radiation field. The condition (X—Y)
= 0 can be realised either with X and Y finite and equal or with X and Y both
zero. We now examine the condition in two types of hybrid-mode waveguide,
Introduction to hybrid-mode feeds 9

noting in passing that the condition is never satisfied in a conventional pure-mode


waveguide.
In a corrugated waveguide X is generally zero, or nearly so, and

(2.5)

where Af measures the frequency deviation from the resonant value / 0 , at which
Y vanishes when m is an odd integer. Thus a corrugated waveguide does satisfy the
constraint required of an ideal hybrid-mode feed but only at certain frequencies.
Usually the corrugation depth is made approximately X/4 at the centre of the
operating band, then m = 1 in the expression for Y. It is an unfortunate property
of corrugated waveguides that the condition for pattern symmetry and zero cross-
polarisation occurs only at specific frequencies, but eqn. 2.2 does show that the
relative effect of the (p independent terms, which are non-zero away from resonance,
decrease as 1/ATV This is a general property of cylindrical hybrid-mode feeds for
which it can be stated that one with a large aperture gives an inherently better
performance than a feed of small aperture, but it will of course produce a narrow
beamwidth pattern unsuitable for prime-focus applications.
For a dielectric waveguide, if krx > 1 and the refractive index differs by only
a small amount An from unity, X and Y are given by:

(2 6)
h -
Thus,
(X-Y) = -(2Anfn (2.8)
and we see that the optimum condition is approached when An is made very
small. The condition is also frequency independent if, as is reasonable, the
dielectric is free from dispersion over the operating frequency band. However,
unless An is sufficiently large it is impossible to efficiently launch the dominant
mode of the dielectric waveguide and a significant amount of energy will escape
directly from the launcher. In practice a compromise must be reached. As an
example, with Anjn = 0*025, the performance of the two types of hybrid-mode
feed become equal when (A/// o ) is 13%. Nearer to resonance the corrugated wave-
guide is superior. Other things being equal, the designer can choose, for example,
between a lower crosspolarisation over a narrower band with a corrugated feed,
or a somewhat higher crosspolarisation over a much wider band with the dielectric
feed.
10 Introduction to hybrid-mode feeds

Some radar- and most satellite-broadcast antennas have non-circular cross-


section in order that a shaped beam can be generated. A feed with a non-circular
cross-section is required to generate an asymmetric pattern on transmit, or match
an asymmetric focal field on receive. In spite of the asymmetry, orthogonal states
of polarisation can exist and are frequently demanded with comparable specifi-
cations as for spot beams.
Both rectangular and elliptical cross-section corrugated horns have been used
as feeds and rectangular cross-section dielectric waveguide feeds have also been
studied. These structures can all be made to exhibit linear polarisation for the
dominant mode. This property has been shown by Dragone [55] to be quite
generally approached as the waveguide aperture dimensions become very large,
provided there exists an axis of symmetry in the transverse plane.

2.2 Focal field of a reflector

The desirable features of hybrid-mode waveguides as feeds can be illustrated


alternatively by examining the focal fields of a reflector. When a plane wave
illuminates the reflector, as under normal receiving conditions, the field in the
focal region can be predicted by means of physical optics. Fig. 2.4 shows the
reflector configuration.

Fig. 2.4 Reflector configuration

The incident plane wave is assumed to have its electric field linearly polarised
in the x direction, then the focal region electric field obtained by integrating the
surface current K over the reflector surface, is

4TT
f [K-(Kir)fr]—dS (2.9)
Introduction to hybrid-mode feeds 11

Here, in the physical optics approximation,


K = 2/7 x Hi (2.10)
and
^ = -H0ye-jkr«cosd (2.11)

//i is the incident magnetic field of the plane wave and the integration is over the
reflector surface. Following Minnett and Thomas [186], it is possible to show
that the focal-region transverse electric and magnetic fields at z = 0 have the
form

E = JGJMO#O (ixh +/ 2 (cos20 f /* + sin20 f ? y )} (2.12)

H = JCOMO ~- (iyh + / 2 (sin 20fiJC - cos 20f L)} (2.13)


z0
The quantities h and / 2 , which are, respectively, proportional to the copolar
and crosspolar fields, are given by

h = J ° sin 6J0(krf sin d)dd [fe'j2hf] (2.14)

h = J*° sin0 tan2 - J2(krf sin 6)d6 \fe'j2kf] (2.15)

Later we shall show that eqn. 2.2 for the transverse electric field in a hybrid-mode
waveguide can be recast in the form
E = A{TJX + 72(co$ 2 fix + sin20* y )} (2.16)
and we can thus identify Ix and I2 as the kernels of the integrals I\ and / 2 . The
field in the focal region is now recognisable as an angular spectrum of cylindrical
hybrid waves. We gain insight into the meaning of this statement when we examine
the form of the transverse-electric field in the focal region of a paraboloidal
reflector possessing a representative f/D ratio as shown in Fig. 2.5. There the
direction of the field is indicated while Figs. 2.6(a) and (b) show for various values
of f/D the copolar and crosspolar intensity of the field. For a long focal-length
reflector (0O small), the field is almost entirely linearly polarised everywhere
although the direction of the field alternates as the radial distance from the focus
rf is increased. For this case / 2 = 0 and the focal field has the form

The above equation corresponds to the scalar solution obtained by the 19th-
century physicist George Airy, for the distribution of light in the focal region of
12 Introduction to hybrid-mode feeds

is ; ; ;
I i > i !I » !

:::\:::;

A(UH) ;;',-
1;E X V V W X

BCUHJ -'( A "*-*-4..L< 1•1

10
!!; . X x v s V s v

I 111

;;', f;v-
~V»^ v ^ v vN v
v x \ \ \ \ \ \ ! !
A(UHJ -;-|-; ~t- -t-j s -- ^T^>V^V N V \ \ \ \ \ \ V \ > \ » v
•-•-i.i i / / > * / - \ v v v ! * ! ! ! ! • ! <

C\ ^*; 11 ; * ! J | 1
OS j j ^ V ' '
1 | » | > I < » J

:H
^ \ \ \\V

' ' 'V v S \ 1 \ I « i I » » I 1

; ; ; ; ; ^ O v l \ » ' 11 J ' ' ^v ,\ »\ \ l t > t l 1 1 1

1
' ' ' ' ' < ~ \ V \\ \ \ »

•»• • • ' i l u S ! ! ! ! ! ! ! !
| 1J J ' ' • » li M » i i » • < i i
1 1 1 ! I ! ' i c *•' *\* •
nl ' ' ' r»!• ' i1 • ' 1 • • ' i 1 ; '.' ' ' ' ' ' ' '
0-5

E(U) - mox.

Fig. 2.5 Field distribution in focal plane of paraboloid with d0 — 63° [176]
a Polarisation of E field
b Contours of amplitude E(u)
Introduction to hybrid-mode feeds 13

a lens. For that case, observation showed the light intensity to comprise annular
bands of light separated by dark rings with the intensity decreasing away from the
lens axis. Even if we relax the constraint on 0O being small, the field in the central
region remains nearly linearly polarised, and, as we shall see later, it matches well
the field of the dominant HE n hybrid mode. To effect a complete match to the
focal field it would be necessary to synthesise all the cylindrical hybrid waves
passing through the focal plane z = 0. This would require an infinitely overmoded
waveguide which is practically unrealisable. For certain special applications where
bandwidth is not a problem, a feed with several higher modes in addition to the
dominant mode can be used to produce very high efficiency.

10

08
\

0-6

0-2
\J
Y \
-0-2

-0 4 1

0-2

3
CD 0
5° T
-0-2
6 8 10 18 20
U

Fig. 2.6 Functions A (U) and B(U) for paraboloid subtending semiangle d0 at focus [176]
AfU) — a/j and B(U) = a/ 2 are respectively proportional to the copolar and cross-
polar focal field

To better understand the relation between antenna efficiency and the focal field
captured by the feed we will next examine the power flowing through the focal
region. From eqns. 2.12 and 2.13 we can obtain the average rate of energy flow
through an elemental area of the focal plane, Sz, as

(2.18)

where Ix and I2 are as in eqns. 2.14 and 2.15.


14 Introduction to hybrid-mode feeds

We next integrate Sz over an aperture in the focal plane of radius a, and introduce
the parameter Ua = ka sin 60. Let the total power incident on the paraboloid be Po>

Po = 2n(ft,nd-A2 f (2.19)

P/Po is then given by


P j .va
r?o = — = - {A\U)-B2(U)}UdU (2.20)

where

A{U) = -cosec 2 -^/! (2.21)

= | cosec2 ^ / 2 (2.22)

and Ua=ka sin 0O- -4(17) and KU) are plotted in Fig. 2.6.
For small 0O, i-e. for a reflector with long focal length, B(U) = 0 and

£ = l-Zg^WHtfa) (2.23)
which is a result derived originally by Lord Rayleigh.
Fig. 2.7 shows P/Po as a function of £/a. When 60 is small, the first maximum
is obtained when Ua = 3.8, a value which corresponds to the first null of A(U).
For large 60, as Ua increases beyond this value, P/Po declines initially but this
is not to say that the power coupled into an optimum hybrid-mode feed (shown
as a chain curve in Fig. 2.7) will decline, this quantity in fact always increases
monotonically. The explanation for this apparent paradox is that there are local
regions off-axis where the direction of the Poynting vector is opposite to that on
axis and these regions are larger for larger QQ, However, as we now show, the fields
that cause these regions of negative Poynting vector, do not couple to the field
of a balanced hybrid mode and so the efficiency of an antenna fed by a waveguide
supporting this dominant-mode field increases monotonically.
Let us rewrite eqns. 2.12 and 2.13 as
(2.24)
Hx = HA+HB (2.25)
where
£ A = CA(U)ix (2.26)

HA = -~-A(lf)iy (2.27)
Introduction to hybrid-mode feeds 15

Fig. 2.7 Normalised power flow in the focal plane of a paraboloidal reflector
aperture efficiency
maximum aperture efficiency

EB = CB(U)(co$ 2<j>tix + sin 2 0 f iy (2.28)

= 7~ B(U)(sin 2<ptix - cos 20 f iy (2.29)

and C is a constant. Now consider a feed of aperture S carrying unit power, placed
in the focal region with axis concident with that of the reflector and with fields
E2 and H2 in its aperture. The power coupled into the feed by the focal fields
E\ and H\ is given by

•dS (2.30)

where it has been assumed that E\ and H\ arise from a plane wave with unit power
incident on the reflector aperture. Now if

E2 = JFfa,0)/y (2.31)
16 Introduction to hybrid-mode feeds

and
F
H
2 = - J (rf,4>)iy (2J2)

r? = C \\2n\aA(U)F(rf,<l>ytdrfd<t> (2.33)
J
Ho o

Notice that in contrast with eqn. 2.20 a term in B(lf) does not arise and r\ increases
monotonically with feed radius a. Maximum efficiency occurs when F(rfi<p)ix
matches A(U) then

o A(U)UdU (2.34)

Eqn. 2.34 is plotted in Fig. 2.7 where it can be seen that the corresponding value
of rjo is always greater than that of 7]0 given by eqn. 2.20.
From eqn. 2.2 we see that when X = Y (balanced-hybrid condition) the trans-
verse field of the feed has a/ 0 (A>) dependance whereas A{U) has the form {/i(0j/
U for small 0O. However, if we arrange that the feed just encompasses the first
null of A(U), i.e. if we choose the radius rx such that

r
3.83 X
i = —f ~ (2.35)

we will obtain maximum efficiency. The value for small 60 is

Vo = l - / g ( 3 . 8 3 ) (2.36)
= 0.84

Thus the maximum antenna efficiency of a front-fed paraboloid with a single-


mode feed is 84%. Should we wish to improve on this value then we must improve
the match between the focal field and the hybrid-mode field of the feed. We can
do this either by using a dual-shaped reflector configuration or by a multimode
feed. The former solution offers potentially a much wider bandwidth and is the
solution chosen in, for example, all large earth-station antennas, whereas in radio-
astronomical telescopes operating in the prime-focus mode of operation, the
latter solution is sometimes chosen.
Although the symmetrical reflector of paraboloidal shape is most frequently
used as an antenna, the offset paraboloid is gaining in importance. Offset con-
figurations offer the advantage that the main reflector aperture is unblocked,
the antenna efficiency is thus potentially higher, but, more important, there is
minimal scattering of energy into the far-out sidelobes of the antenna pattern.
Introduction to hybrid-mode feeds 17
This makes the offset paraboloid of particular value as a space craft antenna and as
an antenna for ultra-low sidelobe radars. We now ask whether the hybrid-mode
feed remains the best choice for this type of antenna. The answer is 'yes' but with
a reservation. Fig. 2.8 shows the focal field of an offset paraboloidal antenna with
a long fjD. In contrast with the field of the symmetrical reflector shown in Fig.
2.5, we find that the offset reflector has two regions of high cross-polarisation
falling within the central region of the focal field, consequently a conventional
dominant mode hybrid feed will couple strongly to these fields and the cross-
polar performance of the antenna is generally poor. All is not lost, however, for,
as will be explained in more detail later, by adding an additional asymmetric
hybrid mode to that of the HE n mode, it is again possible to create a match to
the focal field, restoring excellent cross-polar performance.

copolar contours
decibels

cross-polar
contour decibels

Fig. 2.8 Approximate contour plot of typical focal-plane field distribution of an offset
parabolic reflector uniformly illuminated from a distant linearly polarised source

To complete this introduction, we show in Fig. 2.9 the field along the axis
of a spherical reflector, as obtained by Thomas, Minnett and Vu [192]. The
spherical reflector is useful in beam-scanning or multiple-beam applications and
when the angle subtended at the paraxial focus is small (large //£>), there is
appreciable energy confinement near the paraxial focus. Then a multimode hybrid
feed placed at that position can couple efficiently to the focal field. Alternatively,
by use of a Gregorian subreflector acting as an aberration corrector, see Fig. 2.10,
Phillips and Clarricoats [187] have shown that efficient coupling to a hybrid-
mode feed is possible over a wide range of subtended angles (f/D ratios). Although
18 Introduction to hybrid-mode feeds

200 SchelTs solution


with obliquity factor

Hybrid-wave
solution
100

0.50 0.65

200 -

0.60

Fig. 2.9 Normalised field along axis of a spherical reflector. R/k — 400 for various values of
d0. P.F. -paraxial focus. M.F. = marginal focus Thomas, Minnett and Vu [192]

/ J\ spherical
*-*-*• reflector

centre of
curvature
0 Gregorian
subreflectoi

(or focal curv

Fig. 2.10 Spherical reflector with Gregorian corrector


Introduction to hybrid-mode feeds 19

the spherical reflector offers advantages for beam scanning, the symmetric version
of Fig. 2.10 suffers from serious blockage problems. An offset parabolic torus
reflector, as investigated and produced by Hyde, overcomes that difficulty and it
can be a very useful design for an earth station when communication with several
satellites is required.
Propagation and radiation Chapter3
characteristics of cylindrical
corrugated waveguides

3.1 Introduction

The sucessful design of a feed with corrugated walls depends on the synergism of a
number of related elements of theory. Fig. 3.1 identifies three main types of feed
while Table 3.1 identifies the main elements of the theory.
In this chapter we are concerned with feeds in the class of Fig. 3.1 (a), i.e. the
cylindrical corrugated waveguide with a radiating aperture. Feeds of this kind are
used at the prime focus of a reflector for, as they can be made with aperture
diameters as small as one wavelength or less, they provide for efficient illumination.
The feed of Fig. 3.1 (a) also represents a good first approximation to the feed of
Fig. 3.1 (b) provided that the horn flare angle is small and that a correction is made
for the phase curvature of the aperture fields. Feeds of this kind are frequently used
at the secondary focus of a Cassegrain antenna or in beam-waveguide fed antennas.
Fig. 3.1 (c) shows a feed that combines features of both of the preceding types and
draws on knowledge relevant to both. It is used when a narrow beamwidth
radiation pattern is desired in a feed of short length.
Table 3.1 identifies the main elements of the theory required to determine the
radiation pattern of a corrugated feed guide. First it is necessary to represent the
boundary, we have two options, we can assume a uniform surface impedance for
the wall or we can solve exactly the boundary-value problem. Then the field in the
interior of the waveguide is represented by a space-harmonic series and the field
within each corrugation by an infinite set of resonant modes. The first method is
simple and quite accurate, however, when the waveguide radius is of order one
wavelength, space harmonics must be included if an accurate crosspolar pattern is
to be obtained. Additional computational time is involved in the space-harmonic
solution. Next, in step two, an application of the boundary conditions leads to the
modal propagation coefficients and transverse wavenumbers. In most cases we may
assume the feed to be constructed from a perfect conductor and then the propa-
gation coefficients are purely imaginary and the transverse wavenumbers purely
real. Actually, the corrugated waveguide has interesting attenuation properties.
When the walls have finite conductivity it transpires that the dominant HE n mode
Cylindrical corrugated waveguides 21

exhibits lower attenuation than any mode of a uniform waveguide of comparable


radius. We shall return to this feature later noting for the moment that an assump-
tion of infinite conductivity has no effect on the radiation properties of the feed.

—rr\nnruinnnnr\}

Fig. 3.1 Types of cylindrical corrugated waveguide feeds


a Open-ended waveguide
b Narrow flare angle horn
c Profiled horn

In step three, the transverse fields of the modes are expressed, and generally
these are assumed to be the fields in the aperture. Actually at the aperture we are
confronted with a complicated boundary value problem which has yet to be solved
exactly. A complete solution would require the determination of the amplitudes
and phases of all modes excited at the aperture together with a description of the
currents which flow over the waveguide flange and along the outer wall of the wave-
guide. Fortunately the absence of a complete theory does not constitute a difficulty
in design, for so far as the far-field radiation pattern is concerned, the finite flange
can be treated as if it were infinite. This allows us in step four to determine the
radiation pattern of a given mode most accurately by use of the Fourier Transform
22 Cylindrical corrugated waveguides

Table 3.1 Stages in theoretical determination of radiation pattern of corrugated


feed

1 Characterisation of feed boundary


(a) Surface impedance
(b) Inclusion of higher-order modes in slots and space harmonics in waveguide
interior
2 Determination of modal propagation coefficients
3 Determination of fields in feed aperture
4 Determination of radiation fields by application of either:
(a) Kirchhoff—Huygen method
(b) Fourier—Transform method
(c) Spherical wave expansion (conical horn)
(d) Aperture integration with phase correction (conical horn)
5 Determination of transmission and reflection coefficient at horn mouth and (in
case of conical horn) mode conversion along horn.

method. The Kirchhoff—Huygens method is also satisfactory for large apertures. If


we desire we can correct for the finite size of the flange by using the Geometrical
Theory of Diffraction, but the additional effort is seldom justified. Finally, in step
five, we determine the amplitude and phase of the propagating mode or modes of
the corrugated waveguide, together with the reflection coefficient of the incident
mode in the input smooth-wall waveguide. This problem is solved by a straight-
forward modal expansion method.
We now develop the propagation equation and consider the mode characteristics
combining, in effect, steps one to three above.

slot outer region


depth
T
r0 inner region

, ridge
slot
77//////{//A//\///)///////7777
z=0

Fig. 3.2 Geometry of corrugated waveguide

3.2 Propagation in a cylindrical corrugated waveguide

Fig. 3.2 shows a cylindrical corrugated waveguide and defines the notation. In
general, both longitudinal components of electric and magnetic field are required
Cylindrical corrugated waveguides 23

to satisfy the boundary conditions, the exceptions occurring at cut-off for all
modes and for all values of the propagation coefficient, when the modes are
azimuthally independent (TEOm and TMOm nodes). In the interior r<rx the
longitudinal components of field Ez and Hz satisfy
V2tEz+K&Ez =0 (3.1)
V2HZ+K2NHZ =0 (3.2)
where
Kjj = k2-$2N (3.3)

ft* =ft>+ — (3.4)

k2 = co2e0/io = (2TT/X 0 ) 2 (3.5)


JUJt
and a time dependence e is assumed.
Solutions of eqns. 3.1 and 3.2 in cylindrical co-ordinates are

Ez = t ANmJm(KNr) e-ifo* (3.6)

Hz = Nf BNmJm{KNr) e-M»"* (3.7)

The form of the equations shows the fields to be expanded in space-harmonic


travelling waves with propagation coefficients flN related through the spatial
period p, which is the separation distance of the corrugations.
On the perfectly conducting ridges between the annular slots Ez must vanish.
Within the slots, with r > rx,

E
z = Z [Cp$m cos npZ-j'CpSD sin ripz] cosm<£ (3.8)

P==
sin T]Pz + DPR^ cos n'Pz] sin m<t> (3.9)

The form of the equations shows the fields within the slots to be standing waves
and to satisfy the boundary conditions on the side walls of the slots,

VP = 2P~l (3-10)

Vp = (2P-1)~ (3.11)
b
In the above equations Jm(x), Sm(x, y), Rm(x, y) are, respectively, Bessel and
24 Cylindrical corrugated waveguides

cylindrical functions of order m. The cylindrical functions are defined as follows:

RmQcy) - * Jm(x) Y M - J M Ymix) <3-12)

- * f§^3 (3,3)
S (

(3.15)

To determine the propagation coefficient (3N and the form of the transverse fields
required for the radiation pattern, boundary conditions at r = rx must be applied
to the tangential components of the electric and magnetic fields, viz, Ez, E^, Hz and
HQ. In the Appendix, the propagation equation is formulated using space harmonics
but here we shall proceed with a much simpler method, usually referred to as the
surface-impedance approach. Notwithstanding this departure from rigour, from
time to time, in discussion and in further developments, we shall draw upon the
results of the space-harmonic formulation.
In the surface-impedance approach, we neglect the effect of the ridges and take
only the lowest order TMm standing wave in the slot which has components^, H^
and Hr. This is a reasonable assumption provided the lowest order TE m wave
cannot be supported in the slot, i.e. the slot width b<\/2. The approximation
improves as the number of slots per wavelength increases and the thickness of the
ridge decreases. The greatest effect is upon the ratio of HJEZ and this has the effect
of altering the optimum design frequency as we shall see.
When space harmonics are neglected
(3.16)
Hz = bmn Jm (Kr) e-^smm<p (3.17)
For convenience, we drop the z dependence, and write eqns. 3.16 and 3.17 in the
form
Ez =amJm(x)e*m+ (3.18)
m
Hz =-amjy0AJm{x)e* + (3.19)
z
1/2 is tne
where y0 = (eo/juo) free-space wave admittance. A is the so-called
normalised hybrid factor, a most important quantity in determining the crosspolar
radiation characteristics of feeds. From Maxwell's equations we can express the $
components of the electric and magnetic fields.
For r < rx
E0 = am j 2 /m(*){mJ3 + AF m (x)}eJ'»* (3.20)
Cylindrical corrugated waveguides 25

,3.21)

<3 22)
«•« - * JtM B
-
r
x = Kr, x = kr, p = —

For r 0 > r > ru assuming only the TMm standing wave in the slot,
a
E = "» oD/y x'% (3.23)
2
r(4r m ( *' X o )
r
m\xo)
XQ = kv$

As there are no TE standing waves in the slot EQ — 0 at r = r l 5 thus, from eqn. 3.20,
(3.25)
Eqn. 3.26 is just a special case of the general characteristic equation (29 in the
Appendix). In eqn. 3.26, the fundamental space harmonic N= 0 is taken together
with the first standing wave. Solutions of the exact characteristic equation differ
The characteristic equation for |3 is obtained on applying continuity conditions to
the surface admittance H$\EZ at r = rx.

mQc[,x'o) (3.26)

from those of eqn. 3.26 in ways we shall discuss but many general features of
the exact solutions of Fig. 3.3 are the same. The sketches of Fig. 3.4, which
correspond to different values of rtlr0, and Figs. 3.5—3.7, help us to identify these
features as follows.

The cut-off condition Q = 0


In eqn. 3.26 the condition @ = 0 yields either
*"m(*i) = 0 (3.27)
or
Fm(xi) = Sm(xi,xb (3.28)
Eqn. 3.27 is satisfied when
4 = 0 (3.29)
then Hz is finite and Ez = 0. The cut-off is therefore an //-type and in the limit as
tends to be zero, the mode exhibits pure TE properties with A tending to infinity.
26 Cylindrical corrugated waveguides

Eqn. 3.28 is satisfied when


Jm(x0) = 0 (3.30)
then Ez is finite and E2 = 0. The cut-off is therefore an E-type in the limit as 0
tends to zero, the mode exhibits pure TM properties with A tending to zero. Notice
that for H-mode cut-off the boundary condition corresponds to that of a TE mode
in a uniform waveguide of radius rx while for #-mode cut-off the boundary condition
corresponds to that of a TM mode in a uniform waveguide of radius r0. The differ-
ence arises because Hz is always zero for r > rx subject to our assumption about the
absence of TE waves in the slot, while by contrast Ez can penetrate the slot to
vanish on the perfectly conducting base of the slot at r — r0.

'//?/»/ //''
10

8 -
\

6 --

'/ ''//,/IJfi, / / /
/

7 /
/

2 -
(//:ll //
/ '

i l l i // I ,
/] 1 \ \\X \
H HE
01/ 21\ HE
\ 31 \ 02 H
8
kr,
10

HE,, EH 3 1 E 0 2 EH 12 HE 12

Fig, 3.3 Dispersion characteristics of corrugated waveguide. rx/r0 — 0m8

Following logically from the above observations, the sketches of Fig. 3.4 show
that the //-type cut-off frequency depends on the ratio r1/r0 whereas the Z?-type
does not. There exists, therefore, conditions where the cut-off frequencies of the
HE l m and EH l m modes coincide. The condition for coincidence occurs when
= Sl(x'l9x'o) = 0 (3.31)
For higher order n = 1 modes, this occurs for higher values of the ratio rxjr0 and
then, for very large values of krx, from the properties of Bessel functions, k(rQ —
ri) = n in the limit. Thus the degeneracy conditions occurs for high-order modes
when the corrugation depth is approximately \ / 2 .
When designing corrugated horns, especially those with narrow flare angle, we
are interested in the way the special conditions vary with the normalised corrugation
depth (r0 —r^/X as krt varies. This information is plotted in Fig. 3.6.
Cylindrical corrugated waveguides 27

The balanced-hybrid conditions Sm(x[t x'Q) = 0


When the surface admittance Y = 0, Sm (x[f x'o) = 0 and we have from eqn. 3.26
Fl{xx) = (m$)2 (3.32)
which together with eqn. 3.25 leads to
5
S=oo I
EHu HEii '

i
XHE 1 2

\T/ A
5
S=0 S=oo

HE 12

kr,
c
Fig. 3.4 Dispersion diagrams for corrugated waveguide
arjro = 0"4
b rjr0 = 0*63 (typical of the throat of a horn)
c rx /r0 = 0*8 (typical of the aperture of a horn)
Points 1 and 2 represent cut-off for modes with Ez = 0 and Hz = 0, respectively, at
(3 = 0
P o i n t s 3 a n d 4 r e p r e s e n t c o n d i t i o n @ = ( 3 / k = 1 , c o r r e s p o n d i n g t o S l (k, rlf r 0 ) = — \
(kr,)2
Point 5 represents balanced hybrid or 'open-circuit' boundary condition corresponding

Point 6 represents short-circuit boundary condition corresponding to Sx (kfrlfrQ) —


28 Cylindrical corrugated waveguides

(Af = 1 (3.33)
We refer to this as the balanced-hybrid condition since eqns. 3.18 and 3.19 give
when A = 1
1/2
/ • < / • , (3.34)
H,
Eqn 3.33 has two roots,
A = +1 (3.35)
corresponding to HE modes and
A = -1 (3.36)
corresponding to EH modes.

1-0

0-8

0-6

0-4

0-2

Fig. 3.5 Parametric dependence of special points as function of rx /r0 chain curves denote two
examples of loci for HExi mode in a flared horn.

The condition is marked as 5 in Figs. 3.4 and 3.7. Fig. 3.7 shows A as a function
of krx for values of r1/r0 corresponding to Fig. 3.4. We have seen previously that A
equals 0 or °° at cut-off depending on the type of cut-off. In Fig. 3.7(c), r1/r0 is
such that all the HE l m modes shown have E-type cut-off for which A = 0. Then as
kr1 increases, A increases to reach the balanced condition A = + 1 before tending
to infinity at the 'short-circuit' condition where A = SA = «> a n ( j r 0 — rx ^ A/2. We
note that for asymptotically large values of krx, not represented in Figs. 3.4-3.7,
]3 tends to unity at the condition
Jq(xi) = 0 (3.37)
Cylindrical corrugated waveguides 29

where q = 0 for HE l m modes and q = 2 for EH i m modes. The lowest order solution
for the HE n mode occurs when
xx = krx = 2-405 (3.38)
Dragone [56] has pointed out that, within the surface-impedance model, for larger

o0 2 4 6 8 10

Fig. 3.6 Parametric dependence of special points as function of normalised slot depth

Table 3.2
For HE l m modes
x = x
Y
\ 0m I 1 ~~
4° J
(kr )"'
1
kri
J(
Y
+ i- (3.39)
4 l)3 " ' j

where xnm is the mth root of/n (*) = 0


Y Y2 1
A - 1 xSJ- krx 8
- ( 4 - +" ^ O m )
(kr,)2
x
om Y2 „Y -1 1
-4- (3.40)
2 4
For EH l m modes
v' v )1 Y (3.41)
x
m X
2m \L 2krx !
Y
A- 1 *?„ 2kr (3.42)
x
30 Cylindrical corrugated waveguides

and larger values of krl, the actual value of the surface impedance has less and less
effect on the propagation behaviour of a fast hybrid mode. Eqn. 3.38 generally
represents the lowest order asymptotic solution of eqn. 3.26 for HE l m modes.
Table 3.2, containing eqns. 3.39—3.42 due to Dragone [56], gives general
asymptotic series forXj and A Fis the normalised surface admittance of eqn. 2.4.

The short-circuit condition Sm(x[, x'Q) = °°


The next special condition with increasing krx occurs for fixed r1/r0 when S = °°
and the corrugation depth is approximately one half-wavelength. The surface
impedance is then zero. For this case,
Y = Sm(xfhxr0) = - (343)
corresponding to the label 6 in Figs. 3.4 and 3.7. Under this condition the wave-
guide behaves as if it had a continuous perfectly conducting boundary at r = rx.
Pure mode solutions are obtained with

Ez = 0 for TM modes at r = rx
or
—- = 0 for TE modes at r = rx
dr
The condition is important in the design of the throat of a corrugated horn.

The fast-wave to slow-wave transition point 0 = 1


Next, for a fixed value of rjro, as 0 is increased we encounter the fast-wave to
slow-wave transition at 0 = 1. When 0 = 1, K = 0 and Fm(x) = ±m. A = - 1 the
point marked (3) while A = + 1 the point marked @t From an expansion of eqn.
3.26 about X! = 0, we have for m = 1
(3.44)

The high-frequency cut-off $ = °°


Subject to the surface-impedance approximation, as frequency increases within the
slow-wave domain, the condition is eventually reached where 0 tends to infinity.
This is known as high-frequency cut-off. It corresponds to the condition 0p = n for
a periodic structure, with p, the period, tending to zero. Consider m = 1 in eqn.
3.26. The asymptotic form of Fj (ja) with (a) real is
F1(]a) = a (3.45)
then in the limit the condition requires
Si(x'l9xi) = 0 (3.46)
Thus as Fig. 3.4(c) shows, as the H E a reaches a balanced hybrid condition with
Cylindrical corrugated waveguides 31

Fig. 3.7 Hybrid factor diagrams corresponding to dispersion diagrams of Fig. 3.4
a rjro = 04
b rjro =0-63
crjro =0-8
32 Cylindrical corrugated waveguides

A = + 1 the slow EH n mode escapes to high-frequency cut-off. We also note that


at high-frequency cut-off the value of A is given by

A = - — (3-47)
krl
Summary
Let us summarise the behaviour as krx increases from cut-off with ri/r0 constant
at a value greater than 0-75, using Figs. 3.4(c) and 3.7(c) as a guide. For HE l m
modes, cut-off has 0 = 0 and A = 0. The balanced-hybrid condition has A = + 1,
and for large krx the corrugation depth approaches X/4. Beyond this as we approach
the short-circuit condition (corrugation depth A/2), the mode tends to a pure TE
mode and A = + °°.
The mode continues with A = — °° and as we approach the fast-wave to slow-
wave transition, point /3 = 1, we have now A = — 1. Finally, as frequency increases
further the mode terminates at high-frequency cut-off with j3 = °° and the cor-
rugation depth is 3A/4, then A = — \\krx which tends to zero as krx tends to
infinity. We see that as 0 progresses from 0 to °°, A progresses from 0 to + «> and
from — °° back to near zero again.
For EH l m modes, we begin at |3 = 0 with A = — °° and pure TE conditions. For
the lowest order EH l m mode, the pass-band is narrow and the wave terminates with
0 = oo at the same point that the HEn mode reaches the balanced-hybrid con-
dition. We have termed this first EH l m mode the EH n mode, but some authors, for
example, Dragone [56], simply refer to this mode as 4the surface wave mode'. Such
nomenclature is ambiguous for obvious reasons. The next EH l m mode, our EH 12 ,
remains a fast wave through the first balanced-hybrid condition where the corru-
gation depth is approximately A/4, it has A = 0 and behaves as a pure TM mode
when the corrugation depth is A/2, it becomes a slow wave when A = + 1 and mode
terminates when the corrugation depth is 3A/4. We now see that in general, within
the surface-impedance model, a mode is sustained with increasing frequency over
the range in which the corrugation depth increases by slightly more than A/2. The
behaviour of 0 and A for rxjr0 near and below the degenerate cut-off is also shown
in Figs. 3.4 and 3.7. A detailed account will not be presented but it should be
recognised that such values of rl/rQ can prevail in the throat of a horn.

Effect of space harmonics


Fig. 3.3 shows fir% as a function of krx with inclusion of the N=±l space
harmonics in addition to the fundamental; higher harmonics beyond the N= ± 1
pair have negligible effect on fir. Inclusion of harmonics has an influence on all of
the special conditions discussed above to an extent which depends on the mode and
the waveguide parameters.
Chong [23] has made a detailed comparison for the TE 01 , TM01, EH U and
HE n modes but because the latter mode is the most important in feed applications
we concentrate on his results for the HE n mode. Table 3.3 shows a comparison of
Cylindrical corrugated waveguides 33

Table 3.3 Comparison of low frequency cut-off values (krx) of the HEn mode

ri/r0 0-8 0-7 0-6 0-5 0- 4


SI model 3-0654 2-6822 2-2990 1-9159 1-•5327
SH model
at j30/*i = 0- 1 3-1108 2-7215 2-3202 1-9221 1•7742
% error 1-46 1-44 0-91 0-33 13 •6
A at J V I = 0-1 0-00233 0-0124 0-0372 0-239 31 •7

the surface impedance (SI) and space-harmonic (SH) models, with rxlr0 as para-
meter, of the normalised frequency at low-frequency cut-off (jS = O). Actually,
because of a singularity at j3 = 0, the comparison is made for a value of $rx = 0-1
rather than $rx = 0 ; this small difference has a quite negligible effect on the
normalised frequency because d$jdod is very large near cut-off. The percentage error
in using the SI model is seen to be less than 2% in the range 0-5 <ri/r0 < 0-8. For
the other modes referred to above, the maximum error in the same range is about
4%, thus we can use the SI cut-off mode chart of Fig. 3.8 with confidence. Table

•-H02-
--EH 22 -

EH'41

EH 3 1 -

02 06 08 10

Fig. 3.8 Low-frequency cut-off mode chart


Because of the degeneracy of cut-off discussed in the text the nomenclature
refers to cut-off for r1/r0 values greater than the degeneracy point. Below this the
nomenclature is reversed.

3.4 makes a similar comparison for the high-frequency cut-off. Here the agreement
between the models is weaker because, at the high-frequency cut-off, there is equal
power in the fundamental and N= — I harmonics so neglect of space harmonics
34 Cylindrical corrugated waveguides

Table 3.4 Comparison of high frequency cut-off values (krj of the HEn mode
ri/r0 0-7 0-6 0-5 0-4
SI model 11-23 7-32 4-85 3-29
SH model — 6-5 4-5 3-17
% error - 12-6 7-8 3-8

in the SI model is clearly less valid. Nevertheless, the mode chart of Fig. 3.9 is a
useful guide to design.
In subsequent discussion of the radiation characteristics of corrugated waveguides
and horns we shall frequently make use of the SI model in the vicinity of A = 1.

0-4

Fig. 3.9 High-frequency cut-off mode chart

Fig. 3.10 compares for the HE n mode, the normalised wave number Krt as a
function of the normalised frequency krx using the two models. Corresponding
values of A are shown in the vicinity of A = 1. The value of krx at A = 1 differs by
10% between the two models for this value of ri/r0. Fig. 3.10 also allows us to
check the assumption / 0 (Kri) = 0 which is sometimes used as an approximation to
characterise fields in and radiation from a corrugated waveguide. In Fig. 3.3 we
include points on the propagation curve obtained by making this simple assumption
where it can be seen that over a useful range of values away from cut-off, the
approximation is quite good. Fig. 3.10 does show, however, that the approximation
is rather worse than the SI model might lead one to suppose.
To complete this discussion of the propagation characteristics of corrugated wave-
guides Fig. 3.11 shows for the case rl/r0 = 0-6, jJ and A for the four lowest HE l m
modes. The influence of the finite period is manifest among the higher order
modes where the high-frequency cut-off occurs for a value krx close to that at
which j3 is unity. The condition A = 1 occurs close to the frequency at which the
corrugation depth is a multiple of X/4.
Cylindrical corrugated waveguides 35

Fig. 3.10 Comparison of wavenumbers and hybrid factor in space harmonic and surface im-
pedance calculations: rx /rQ = 0-8, s/rx = 0-2, t/b = 03
spaed harmonic calculation
surface impedance calculation
Krx =2-405

HEn HE 12 XHE 13 \ \HE

10 12

Fig. 3.11 Dispersion characteristics and hybrid factor for HExm modes: rx/r0 = 0'6, s/rx
02, t/b = 0-3,K = L = \
36 Cylindrical corrugated waveguides

3.3 Fields in a cylindrical corrugated waveguide

Our main objective, the radiation characteristics of hybrid-mode feeds, requires as


an intermediate goal the fields in the feed aperture. For the cylindrical waveguide,
subject to the use of the SI model, the electric and magnetic fields are obtained
using eqns. 3.18—3.21. Table 3.5 presents these equations in polar and cartesian
form, numbered 3.48-3.62.
Table 3.5
Ez = amJm(x)cosm(t> (3.48)
Hz = am y0AJm(x) sin mcj) (3.49)
k ,
K

Hr = —\amyo—2Jm(x)[^AFm{x) + m] %inm(S) (3.51)

(3.52)

(3.53)
A ~
= xJ'm(x)IJm(x),x = Kr. (3.54)

= ~ j T * [6 + A) ^o (x) + (A - 0) Jj (JC)] cos0 (3.55)


2 A^
= j ~~ K ? + A ) / 0 ( * ) - ( A - ? ) / 2 ( x ) ] sin0 (3.56)
^/ A.

y 0 7 KPA ) O W ( 0 ) 2 ( ) ] in0 (3.57)


2 A

% P+1)JW(1 J3A)/()] s^ (3.58)

Ex = - J ? l [(?+A)yo(*) + (A-^)/ 2 (*)cos20] (3.59)


2 A

^ (3.60)

(3.61)

// y = - -]yo a-± i [(1 +13 A)/ o (*) - (1 - ? A ) / 2 (x) cos 20] (3.62)
22 A
A
Cylindrical corrugated waveguides 37

The case m = 1 is of particular interest and in equations 3.59-3.62 we present


the corresponding Cartesian components of the tranverse electric and magnetic
fields. Two features are evident; first, the condition A = |3 renders the transverse
electric field Ex independent of 0 and the component Ey identically zero. This

1 1 l
-in r

I I
-11
' I ' space harmonic number

Fig. 3.12 Amplitude of space harmonic coefficients for HExl mode: K = L = 11, rx/r0 — 0'8,
rx = 40 mm, s - 10 mm, t/b = 0' 1, krx - 84
a Electric field coefficient A
b Magnetic field coefficient B

condition is important because when the Fourier Transform method is used to


obtain the radiation field from the aperture electric field, the same condition yields
zero radiated crosspolarisation when the radiating aperture is large in terms of wave-
length. Secondly, we note that the condition which renders the transverse magnetic
field to be linearly polarised is not A = J3, but A/3 = 1. The two conditions occur at
nearly the same frequency when fi -* 1 which it does for a large aperture waveguide
in the SI modes.
38 Cylindrical corrugated waveguides

Fig. 3.13 Transverse electric-field patterns of HEn mode at various frequencies above cut-off
a fifc = 1 0 b flfc = 1 "23 c flfc = 1 -89 d f/fc = 242 e f/f = 2 8 0
Cylindrical corrugated waveguides 39

We know from our prior discussion that space-harmonics are required for an
accurate representation of the fields in a periodic waveguide. The space-harmonic
components of the fields are confined near the boundary r — rx where they modify
the field matching conditions between the interior of the waveguide and the slot
region. Chong [23] has made a detailed study of these features and below we
present just a few of his results. First, in Figs. 3.12(a) and (b) we show how the
amplitudes of the space-harmonics of the longitudinal electric and magnetic field
vary with the order N of the harmonic. The oscillatory nature is evident together
with the slow convergence. This behaviour explains the small change in the
propagation coefficient between K = 1 and K = 11. In passing, we note that field
matching demands the same number of slot modes as harmonics in the interior,
thus we always chose K = L in computations involving the space-harmonic model.

\h
/////////////jx '/////////(M

i Aft A

i 0
H
Fig. 3.14 Longitudinal electric field pattern of HEU mode: rJrQ = 0 8, rx = 40 mm, Xg =
30 mm at 10389 GHz

Fig. 3.13(a)-~(e) shows the transverse electric field pattern of the HE U mode as
the frequency / is increased above the cut-off value / c . As indicated before, for
these parameters, at cut-off the field of the HEn mode exhibits pure TMn pro-
perties but as frequency increases the mode changes becoming nearly a pure linearly
polarised wave well above cut-off.
To complete our discussion Fig. 3.14 shows the electric field in a longitudinal
section of the waveguide at an instant of time while Fig. 3.15 shows corresponding
currents on a developed section of the corrugated wall. Here it must be said that
discontinuities in current paths do arise because of the limitations imposed by
inaccuracies in field matching, a consequence of limited computer resources.
40 Cylindrical corrugated waveguides

e=o c

base side ridge side base side ridge side


top top

Fig. 3.15 Current on the surface of the corrugations for HExl mode (slots have been "opened
out' to display the currents), parameters as Fig. 3.14

3.4 Attenuation in corrugated waveguides

Introduction
Although the attenuation of waveguides and horns used as a feed is so low as to be
a relatively unimportant factor in design at present, the same does not necessarily
hold for waveguides used as feeders, especially if the antenna is mounted remote
from the transmitter. Furthermore both feed and feeder attenuation contributes
directly to the system noise-temperature in satellite receiving terminals and this
consitutes another reason for interest. Finally, with a progressive move towards
higher microwave frequencies for telecommunications, attenuation will become an
increasingly significant feature in design in the future.
A very detailed study of the attenuation of corrugated waveguides was made
during the early 1970s by the authors and their co-workers at Queen Mary College,
following a discovery made by one of them in 1969. Clarricoats observed that the
HE n mode of a corrugated waveguide had a lower value than that of the TEOi
mode in a smooth-wall waveguide of comparable size. It will be recalled that at this
time considerable commercial exploitation of the TE01 mode long-haul waveguide
was expected so the discovery attracted some interest.
In simple terms the explanation for the phenomena is easy; the HEn mode in a
corrugated circular waveguide operating near the balanced-hybrid condition has
only an Hz component of magnetic field tangential to the wall surface, i.e. H^^Q.
In common with the TEOi mode, t\mHz component is very small compared to the
transverse components of electric and magnetic field which support the power
flowing through the waveguide and furthermore, it decreases with increasing
normalised waveguide radius. Since attenuation is proportional to the strength of
the field tangential to the metal surface the attenuation is correspondingly low. All
fields in the slot region of the corrugated waveguide are proportional to the strength
Cylindrical corrugated waveguides 41

of the Hz field at the wall so, in spite of the increased wall area compared to a
smooth wall waveguide, the attenuation is much lower. The fact that the HE U
mode has lower attenuation than the TE01 mode in a comparable size of waveguide
can be explained by the presence of two loops of magnetic field for the TE01 mode
compared to only one for the H E n mode. The Hz field in the H E n mode is thus
lower for a given power flow. Compared to the dominant T E n mode of smooth
wall waveguide, the HE U mode in a corrugated waveguide has an attenuation which
is significantly lower, typically by one or two orders of magnitude.

Evaluation
The evaluation of the attenuation coefficient of a corrugated waveguide is presented
below using the SI model. The derivation is straightforward and physical insight can
be obtained from SI results. However, in contrast with the radiation characteristics
of a corrugated waveguide, inclusion of space harmonics is vital to an accurate result
for attenuation. The space harmonic formulation has been developed by us and is
given in Reference 39. In this section our numerical results will draw on that theory.
We begin by deriving an expression for the Poynting vector Sz for m = 1 modes.
In general,
Sz = ExHy -HxEy (3.63)
For the m = 1 modes, we have on substitution from eqns. 3.59—3.62,
a
? - h I*'
{Jl(x)-Ji(x)}-2$(1 - A2)JQ(x)J2(x) co$2<p} (3.64)
Eqn. 3.64 shows incidentally that when A = ± l the energy distribution is
circularly symmetric. Further if A — $ — 1 we have

- I Jl(x) (3.65)
To obtain the total power flowing through the region r < rt PT, we must integrate
Sz over 0 < r < rt. Then

T — 9 l I •/ 0 */ j \X i j J^ jj \^ 1 i IV j ^2 •*• 1 \X 2 J ~r~ r j \Jt \ j

+ £(*i ~ 1 ) } + A(l + ^ 2 ] (3.66)


In the same approximation as above, noting Ft (xt) ^ — 1,

(3.67)

The above results are for circularly polarised waves; for linearly polarised waves the
power flow is halved.
42 Cylindrical corrugated waveguides

If space harmonics and higher-order slot modes are neglected, a simple expression
for the attenuation coefficient a in a corrugated circular waveguide can be obtained
following lengthy algebraic manipulations.

a = (3.68)
2PTb

s
VY
10-*
: \V \ \ " " " " - ^

V \

\\
TEn(R s ro) - -
. \

CD

c
\
o

•\TE 01 (R = r i )
a 1 0" 3

-
-
10" ._ i t 1 J —1 - 1 —

6 10
radius r<],cm

Fig. 3.16 Comparison of HEXI mode attenuation as a function of radius in brass corrugated
waveguide and TEn mode and TE01 mode in brass smooth-wall waveguide. Frequency
= 9 GHz. sA = 0 25. b/K = 0 3. t/b = 0-1

where
(3.69)
and JPLI *S t n e power lost respectively on the side-walls and base of the corru-
gations, o is the slot width and the ridge thickness is assumed negligible.
Cylindrical corrugated waveguides 43

p — n2 _

(3.70)

03

C
o

001

0001
tur-j/c

Fig. 3.17 Attenuation characteristics of corrugated circular waveguide. Conductivity — 5m8 X


10''S/m, rx = 25mm, s = fOmm, b = 10mm, t/b = 0 2

A, = k2r0 (3.71)
[S?(x'uxf0)]2
Rs is the resistivity of the metal from which the waveguide is constructed. The
functionsSf (x,y) andSifx,^) are defined in eqns. 3.14-3.15.
Under balanced hybrid conditions eqns. 3.70 and 3.71 simplify so that when
combined with eqn. 3.67 for P T , the expression for a becomes, when ri/r0 ^ 1

(3.72)
44 Cylindrical corrugated waveguides

whenxi > 1

Eqn. 3.73 shows that the attenuation at the balanced-hybrid condition falls as
k/(x[)3 which implies a dependence upon rx of r^3 and upon frequency / o f f'2.
The resistivity depends u p o n / a s / 1 7 2 leading to an overall frequency dependence
of/" 372 . These are the same characteristics as are exhibited by the TEOi mode in a
smooth wall waveguide. As mentioned previously, however, the results of exact
calculations shown in Fig. 3.16 show that the HE n mode in a corrugated waveguide
has lower attenuation than the TE01 mode in a waveguide of comparable size. Fig.
3.17 shows the attenuation for higher order modes in addition to the HE n mode.
An extensive programme of measurements made by Clarricoats, Olver and Chong
[39], also by Parini [129] have supported the theoretical results presented above.
The interested reader is referred to these for further details.

3.5 Radiation from corrugated waveguides

3.5.1 Introduction
Two methods have been used to determine the radiation pattern of cylindrical
corrugated waveguides: the Kirchhoff—Huygen (KH) method and the Fourier
Transform (FT) method. For a complete description of the radiated field, both
methods require a description of transverse fields over the waveguide aperture, over
the waveguide flange and over the remainder of the aperture plane. The KH method
requires both the transverse electric and transverse magnetic field to be specified,
whereas the FT method only requires the transverse electric field. Since the trans-
verse electric field vanishes over the assumed perfectly conducting flange and is
very small beyond, the FT method yields accurate results when the field beyond
the radiating aperture is assumed to be zero. This assumption is equivalent to an
infinite flange. To obtain comparable accuracy with the KH method would require
a knowledge of the currents over the flange in order to describe the transverse
magnetic field there and this information is not readily available.
When the aperture is large compared to a wavelength the currents over the flange
are very small and then the two methods yield very similar patterns. To improve on
the approximate FT method a correction for the finite size of the flange can be
made using the GTD method.
From the above it is clear that the FT method offers both accuracy and simplicity
and it is used in the treatment of radiation from the corrugated waveguide
described below. Readers interested in details of the KH method are referred to
Section 4.5 where results for conical horns are given.
Cylindrical corrugated waveguides 45

3.5.2 Radiation from a circular aperture


If, as in Fig. 3.18(a)? the Cartesian components of electric field Ex, Ey are specified
over the aperture 0 < r < rx, 0 < <p' < 2TT, the radiated components of electric field
at R, d, (j> are given by
/•27T r ' i
Ex = A\ Ex exp (j (3.73)
Jo J0 y

Fig. 3.18 a Geometry for radiation from corrugated waveguides


b Copolar radiation patterns for a corrugated waveguide of aperture diameter — 4\
eqn. 3.84
exact solution using space harmonics
c Crosspolar radiation patterns corresponding to Fig. 3.18 (b)
eqn. 3.85,
exact solution using space harmonics
46 Cylindrical corrugated waveguides

where

A = JL *COS#

-30

-50

d Peak crosspolar power as a function of frequency corresponding to Fig. 3.18 (c)


(ridge width /slot width = 0"2 for space harmonic model)

Initially we neglect the effect of space harmonics and insert £* and£ y from eqns.
3.59—3.60 into eqn. 3.73 so as to obtain for n = 1 modes
& xp m J.IB {{A 0 +(A-]3)iV 2 cos20} (3.74)
Eyp IB{(A (3.75)
where
k
B = —ja 1 - 7 T

Nk = = o,2
z;2
*i == A:/- 1
v = sin 6
To obtain the copolar and crosspolar fields (the latter in the third definition given
by Ludwig, see Section 5.2) we use
ft 0 \ f)
- - sin2 - cos20 \-Eyp sin2 - sin20 (3.76)
Cylindrical corrugated waveguides 47

Ecx = -Exp sin2 - sin20 + Eyp(cos2 - + sin2 — cos20 | (3.77)


2 \ 2 2 )
On substitution for Exp and Eyp
1 I ft f) \
Eco = AB [(A + p)N0 + (A - J3)W 2 cos20] cos 2 sin2 - cos20
_ _ e 1
— (A -P)N2sin2 - cos20sin20 (3.78)
In the principal planes where 0 = n — and n = 0 , 1 , 2,3

° S flljn = K' 3 (3-79)


In the 45° planes where <p = n — andn = 1, 3, 5, 7

£•c o = AB (A + ^)JV 0 COS2 I (3.80)

The crosspolar field vanishes in the principal planes while in the 45° planes, where
it attains a maximum value off-axis,

ECI = AB - (A + p)N0 sin2 | + (A - ^)A^2 cos2 M (3.81)

cr
The ratio s C is given by

c = _sin2i+|z|l^ cos2 i (3 82) .


2 (A + 0) ^V o9=o 2
For the HE n mode in a large aperture waveguide J3 « 1, JC2 « 2.405, and following
Dragone [56], we can show that the maximum in the cross polar pattern occurs

when v = 3.67, then if sin2 — < 1,

(3 83)
-
Recalling that v = krxsinS, the above expression is valid when krx is greater than
about 8.
Let us now examine the patterns predicted by the above equations. Fig. 3.18(b)
shows for 2rxl\ = 4, the normalised copolar power pattern

obtained by referring the radiated power to that in the boresight direction and
48 Cylindrical corrugated waveguides

,^7\ -30

,f \ dB (> A
N
J' r // / v
\
/ \ All
¥ \ Lo
\K
— / / v^lj
\\ A
\
1 X
• ' / ' '

\\\
1

'' I
! /

I
- I

1? !

ii

I1
1 I i i
-90 -60 -30 0 30 60 90
degrees
a

-20-

-40

Fig. 3.19 a Crosspolar radiation patterns of open-ended corrugated waveguide


measured computed with Fourier Transform (FT) theory com-
puted with Kirchhoff—Huygen (KH) theory Geometric parameters as Fig. 3.17.
b Peak crosspolar power as a function of frequency for waveguide of Fig. 3.19 (a)
K = L = 1 K = L=0 HE12 mode crosspolarisation. Vertical
bars indicate measured values
Cylindrical corrugated waveguides 49
n
assuming that Xi = 2-405, Fig. 3.19(c) shows for the angles where sin2 — < 1

(3.85)

In both cases the results are compared with predictions based on FT theory applied
to the aperture fields including N=±\ space harmonics in addition to the
fundamental. The agreement is excellent for the coplar pattern and satisfactory for
the cross-polar pattern. However, the absolute value of Pcx requires an accurate
estimate of the factor [(A — J3)/(A 4- J3)]2. For other than small aperture waveguides
we can obtain |3 with considerable accuracy from the expression

^ (3.86)

with*! =2-405.
A depends much more critically on the corrugation parameters, as discussed in
Chapter 5, and to obtain a useful estimate of the crosspolar level the value of A
must be obtained from a solution of the characteristic equation including the ± 1
space harmonics in addition to the fundamental, as Fig. 3.18(d) shows.
In practice, feeds whose circular aperture is preceded by a section of cylindrical
corrugated waveguide are of two kinds; those of small diameter, where to obtain
accurate predictions eqns. 3.78 and 3.81 must be used without approximations and
a correction made for the finite size of the flange, and feeds of large diameter
forming part of a so-called profiled horn. In the first category, Fig. 3.19(a) shows
results for a corrugated waveguide of 5 cm internal radius measured at wavelengths
around 3 cm. The patterns compare well with theory, provided space harmonics are
included, and the graph of maximum crosspolar intensity as a function of
frequency, Fig. 3.19(b), also agrees well, although the effect of higher mode
excitation is in evidence at higher frequencies.
For large diameter corrugated feeds the intrinsic radiation from the HE n mode
can be predicted with accuracy using the approximate eqns. of 3.84 and 3.85. To
achieve these aperture sizes of greater than about 5X in a reasonably short length,
the aperture must be connected to the input waveguide by means of a profiled
section and in this section higher mode conversion can occur unless the design is
optimum. Thus, in general, the patterns will be disturbed by the presence of higher
modes in the radiating aperture. These must be included for accurate prediction, but
the main design objective is to minimise their excitation. Further discussion of the
profiled horn will be deferred until mode excitation has been discussed in Section
4.4. Additional design information can be found in Chapter 5, Section 5.8.7.
A large aperture can also be achieved with a narrow flare-angle conical horn.
Results for such horns, where higher modes are minimal, approximate well with
those of large aperture corrugated waveguides. The agreement between experimental
50 Cylindrical corrugated waveguides

and theoretical results is very good, but to obtain precise correlation account must
be taken of the spherical wavefront. Discussion of this well be taken up again in
Chapter 4, see Figs. 4.21 to 4.23.

3.6 Junctions in corrugated waveguides

The most common means to excite the dominant HE U mode of a circular cross-
section corrugated waveguide is from a circular waveguide with a uniform wall
supporting the TE U mode. The problem was first analysed by Clarricoats and
Saha [29] with later studies by Cooper [49], Dragone [56], James [74, 75, 77]
and Kuhn [171], among others. The configuration is shown in the inset to Fig.
3.20.
O-3O r

0-25 -

0-20

0-15

0-10 -

0-05 -

2-2 5-2

Fig. 3.20 Reflection coefficient of a TElx mode incident on a smooth-wall to corrugated wave-
guide junction. Parameters rx fr0 and rx jr^, r^ = radius of smooth wall waveguide

To determine the reflection and transmission coefficients at the junction, we


expand the electric and magnetic fields in the plane of the aperture in terms of
normal modes of each waveguide. An outline of the modal expansion method now
follows. If e t | n , h t m and e t M h t M are, respectively, the transverse electric and
magnetic fields of the mth andMh mode in the two waveguides,

(l+p)e ti + I ametm = £ •M (3.87)


m =2 M-l
Cylindrical corrugated waveguides 51

O-P)htl- I amhtm = £ AMhtM (3.88)


m-2 M~l

By applying the orthogonality condition

j etnxht*mdS = 0 m^n (3.89)


J
s
and the corresponding equation for modes in the other waveguide, a doubly
infinite matrix of equations is generated. By taking a finite number of terms, i.e. a

V32

1-28

124

L 1-20
in
>

1-16

1-12

1-08

- o o 0 <) O

1-04 o o
o
1 i
8 9 10 11
1-00 frequency. GHz

Fig. 3.21 Measured VSWR of a 12 degree semi-flare angle corrugated horn and theoretical
curve for rJrQ = 0.22 and r1/rh= 1.1

finite number of modes in each waveguide, the equations can be solved. In general
the coefficients am and aM are complex and if p modes are taken in each waveguide
a 4p x 4p matrix equation has to be solved.
52 Cylindrical corrugated waveguides

In a rigorous solution, fields in the corrugated waveguide must be obtained using


the space-harmonic formulation discussed in Section 3.2, but to a good approxi-
mation, in most cases they can be neglected and a surface impedance model used
instead.
Fig. 3.20 shows theoretical values of TE n mode reflection coefficient obtained

Or

-50
16 19 22 25 28 31 34 37
frequency, GHz

Fig. 3.22 Return loss of a corrugated horn with semi-flare angle of 4 degrees and aperture
radius 3-14 cm [56]

by Clarricoats and Saha [29] in a study in which the parameters r1/r0 and r1/rh
were varied rh being the radius of the uniform circular waveguide.
A comparison between experiment and theory is shown in Fig. 3.21 where the
agreement lies within experimental accuracy. It is to be noted that the reflection
coefficient falls to zero at those frequencies where in the surface impedance model
of the corrugated waveguide, the wall impedance is zero, i.e. the corrugation depth
is approximately A/2.
In an early study of hybrid mode waveguides made by Clarricoats, an approxi-
mate value for p was shown to be given by

\P\ = (3.90)
01 + 02
where /3i and 02 a r e the propagation coefficients of the modes of the two wave-
guides. This result was established independently by Dragone [56] more recently
and Fig. 3.22, due to him, satisfactorily compares eqn. 3.90 with experimental
results.
James [74, 75, 77] has approached the discontinuity problem from a different
standpoint. He treats the corrugated waveguide as a cascade of uniform waveguides
Cylindrical corrugated waveguides 53

of very short length with radii corresponding to the inside of the corrugated wave-
guide and the base of the corrugation. The approach has the virtue of allowing one
to treat transitions in the throat region of a corrugated waveguide where the
corrugation depth changes rapidly with distance along the waveguide. Fig. 3.23
shows James's prototype configuration and theoretical return loss for different
types of transition. We notice that the transition yielding lowest return loss is one
in which depth is varied from X/2 at the throat entrance towards X/4 at the exit.

nnm.
smooth-wall corrugated
waveguide waveguide

-10
CD
•o

uT-20
in
o
c
1-30

-40

2-4 2-8 3-2

Fig. 3.23 a Throat region of corrugated horn showing junction between smooth-walled wave-
guides of differing radius.
b Theoretical return loss of seven slot corrugated mode converter
(i) Return-loss without convertor
(ii) Convertor with slot depth increasing from zero
(Hi) Convertor with decreasing slot depth, beginning with X/2 at the input {f0 is the
centre frequency where kr{ = 2'9)
(iv) Convertor with constant slot depth but varying t/b

A convertor with constant slot depth of X/4 is seen to be the least favourable of
those studied. Kuhn [171] has recently solved the discontinuity problem using a
scattering matrix formulation. The design of transitions is discussed again in
Chapter 5.
54 Cylindrical corrugated waveguides

3.7 Multimode corrugated waveguides

The feeds considered above are designed to support one mode of propagation and a
higher-order mode is present only as an undesirable feature. This situation contrasts
with the multimode feed where the presence of the higher-order mode is an essential
and desirable aspect of the design. Multimode feeds can be designed using smooth-
wall waveguide but in general their features are similar to those of multimode

| (i) side view (»i) end view


z=0

schematic diagram of waveguide junction geometry

Channel Aperture Distribution Constituent modes

HE,,

azimuthal difference

elevation difference

HE 2 1

Fig. 3.24 Corrugated monopulse feed and monopulse mode patterns

corrugated feeds, with one exception. The exception is the so-called Potter horn
[188], where a hybrid field is synthesised by the excitation of a T E n and TM n
mode in correct amplitude and phase. Such a feed emulates the HE n mode of a
corrugated feed, but over a limited bandwidth. Returning to multimode corrugated
feeds per se, two principal applications exist: in one, higher modes are present to
improve efficiency; in another, they provide tracking information. An example of
Cylindrical corrugated waveguides 55

the first kind is the feed combining the HE n and HE12 modes which was developed
as a prime-focus feed for the Parkes radiotelescope by Vu and Vu [157]. Another
example in this class, is the feed combining HE n and HE21 modes proposed by
Rudge and Adatia [189, 190, 132] to synthesise the focal field of an offset re-
flector antenna. In the second class lies the multimode feed shown in Fig. 3.24
together with corresponding modal fields. The HE U mode provides the signal
channel while combinations of orthogonal HE21 modes with, respectively, HOi
and E02 modes provide azimuth and elevation outputs. A feed of this kind has
the advantage over a pure-mode equivalent in that it affords polarisation purity
and the concomitant opportunity to extract polarisation information in sum and
difference channels.
The feed is operated at a frequency such that the normalised propagation
coefficient of all three of the tracking modes are nearly degenerate and lie close to
the condition J3 = A.
General expressions for the fields are as follows:
HE n mode (sum channel)
an k - - _ _
2 K

E
y = ~i "-? I {(A-J3)/2(A>)sin20} (3.92)

The form of the equation Ey shows that the aperture field becomes perfectly
linearly polarised when A = j3. However, recalling eqn. 2.83, other than for very
large apertures, the condition for zero cross polarisation of the radiated field
occurs at a frequency slightly above that at which A = 1. This latter condition is
referred to as balanced hybrid, and when it occurs the corrugation depth of a
waveguide with krx > 1 is approximately A/4, as explained in section 3.2.
H ol mode (contributor to the azimuth-difference channel)
k
Ex ~ JJ^o^oi "7 A(^)sin0 (3.93)

A: (3.94)
Ey = -iy0bu
K
E02 mode (contributor to the elevation-difference channel)

(3.96)

k -
/ , ( * , ) sin* (3.96)

HE21 (contributor to the elevation-difference channel)


55 Cylindrical corrugated waveguides

or

0.975

0.905 , 0.940 0.975


a/c

Fig. 3.25 Mode excitation levels as a function of a/c for square waveguide with waveguide
centres located at r{/2
a Excitation levels of modes in sum channel operation
b Excitation levels of modes in azimuthal difference channel operation
c Excitation levels of modes in elevation difference channel operation
Cylindrical corrugated waveguides 57

HE2i* (contributor to the azimuth-difference channel — asterisk denotes


orthogonal polarisation)

E
x — ~~\ — ~~ {($ + A) J\{Kr) coscp + (A — $) Ji{Kr) cos3$} (3.97)
y* 2 A
a2\ k - - - -
x* 2 K

In contrast with the H E n mode, the equation above shows that for the HE2i mode,
even if the term in (A —J3) is negligible, the field has equal copolar and crosspolar
components. But, assuming mode degeneracy, a judicious choice of the ratio of
coefficients a02 and a2i ensures, provided the fields are combined in-phase, that the
crosspolar component of the total field vanishes. Then the linearly-polarised
aperture field under nearly balanced-hybrid conditions generates an almost linear-
polarised far field. Similar arguments apply for the H ol and HE2i* modes.
In summary, the contributions to Ey from the H Oi , E^ and two types of HE2i
modes cancel when the terms in (A — j3) are neglected, then the aperture fields of
the difference channels have the form.
= c
Ex i Ji(Kr) cos 0 (elevation) (3.99)
Ex - C2Jx{Kr) sin0 (azimuth) (3.100)
As noted previously, any departure from the correct amplitude or phase of the
combination will introduce crosspolarisation, as also will a significant departure
from the condition that (A — J3 ) is negligible for the HE2i mode.
Clarricoats and Elliot [25] have made a detailed analysis of a feed of the kind
shown in Fig. 3.24 and have examined how the excitation coefficients depend on
the size and location of the square input waveguides. Typical results are shown in
Fig. 3.25. From their parametric study they conclude that a cluster of square wave-
guides close to the centre of the feed is best to achieve appropriate excitation of all
the modes, although they also demonstrate the need for a mode-selective phase-
shifter in order to produce the correct phase characteristics in the feed aperture.
A similar multimode feed which just uses the HE l l 5 HE2i and E02 modes has
been developed by Watson, Dang and Ghosh [161]. In this feed the azimuth and
elevation channels are extracted from a sum and difference of the latter mode pair.
Their feed system was developed for use in a spacecraft where accurate alignment
of the antenna with respect to a point on earth, is an essential requirement for
efficient transmission.
Propagation and radiation
characteristics of conical corrugated
waveguides

4.1 Introduction

Many corrugated feeds used with reflector antennas take the form of conical
horns or have a conical region even though the aperture region is cylindrical. For
horns with small semi-flare angles, in the range 6° to 15°, the horn behaves as a
cylindrical waveguide but a correction must be made for the spherical phase variation
across the horn aperture. This procedure, which was explored by Parini and Olver
[127], enables one to characterise the aperture field very accurately since space
harmonics can be described in a cylindrical corrugated waveguide, as discussed in
Chapter 3. Once we have the aperture field, the radiation field can be obtained
using the Fourier Transform method described in Section 3.5.
An alternative method which works well over the range of flare angles from
about 10° to 80° uses spherical modes to describe the field in the horn and in
particular in the horn aperture. The radiation pattern can be obtained by integration
over the spherical phase surface or by means of an expansion. Two expansion
methods have been tried: the spherical wave expansion method, developed by
Clarricoats, Saha and Olver [30, 43] and extended by Mahmoud and Clarricoats
[95]; and the Laguerre—Gaussian method developed by Bitter and Aubry [17].
The latter method is useful when a corrugated feed forms part of a multiple
reflector system designed using Gaussian beam approximations.
In this chapter we shall explore these methods, concentrating especially on the
spherical wave expansion method because of its wide applicability. We shall also
have need to explore mode conversion along the horn for in many cases the cross-
polar radiation pattern will be dominated by an undesired higher-order mode
rather than the dominant HE H spherical hybrid mode.

4.2 Modal characteristics of corrugated conical horns

A conical corrugated horn is shown in Fig. 4.1 (a). To a good approximation the
field in the horn may be described in terms of spherical modes which are hybrid
unless the field has zero azimuthal dependence. The boundary condition at the
Conical corrugated waveguides 59

wall 6=di presents a problem since it is not possible to describe fields in the slot
rigorously. However, one approximate method due to Jansen, Jeuken and
Lambrechtse [80], which works quite adequately in the aperture region of the
horn, treats the slot as if it were part of a conical waveguide, as shown in Fig. 4.1(b).

Fig. 4.1 Conical corrugated horns

We can then describe the boundary through an average slot admittance Y = —


(H<p/ER) an expression for which is given below (eqn. 4.5). In spite of its simplicity
the average surface impedance model has been found to work very well and
especially for wide angle horns.
We shall now develop expressions for the fields in the horn using the spherical
mode method.
60 Conical corrugated waveguides

Modal fields
As previously remarked the modes are, in general, hybrid and both electric and
magnetic fields exist in the outward radial direction R of Fig. 4.1 (a). These can
be written for any particular mode, assuming a time dependence exp Qcot) as

ER = A ^~p Hf\kR)P™(cos 6) cos m<£ (4.1a)


J tCJK.

V
yQHR =AA -^~^H^\kR)PT (cos 6) sin mtf> (4.1 b)

6) is an associated Legendre function of order m and degree u, while


is a spherical Hankel function of order v. A is the hybrid factor possessing
a characteristic frequency dependent value for each mode. Knowing ER and HR ,
the transverse fields are obtained as
REe = -AH?\kR) [j3/>;m(cos 0) + wAP™(cos0)/sin 6]
(4.2a)
REQ = AH™(kR) [AP'vm (cos 8) + m ^ m ( c o s 0)/sin 61] sin m<j> (4.26)
y<>RHe = - ^ ^ 2 ) ( ^ ) [ A ^ m ( c o s 0 ) + mP™(cose)/sin6l] sinm0
(4.2c)
] cosm<p

(4.2d)
where /3 is defined by

= }Hp\kR)/H?\kR) = j~ {ln[^ 2) («?)]} (4.3)


u/Cxv

from which it follows that

H<?\kR)/H?XkR0) = exp - j * J* ?(/J)d/i (4.4)

/? is recognised as a normalised complex propagation parameter. Various approxi-


mate formulae for /J in various ranges of kR are summarised below.
Since HU>(kR)l/2(kR) ={kR)V2H$V2{kR\ eqn. 4 3 becomes

0 = j [UlkR + ^ > i / 2 QeR)/H<81/2 (kR)] (4.5)

When &# < |P| 2 , we may use the asymptotic expansion of the Hankel function in
eqn. 4.5, then an approximate expression for j3 is
0 21 l - j ; ( | / + l)/2(^i?) 2 (4.6)
In the transition region where kR and |p| are of the same order, the logarithmic
Conical corrugated waveguides 61

derivative of the Hankel function can accurately be approximated by the Airy


function.
1/2 (kR) ~ (2/M) w\ (f)/Wl (t)
2/3
t = {kRl2) {{p + \fl{kRf-\}
and wx (f) is the Airy function which equals Bt(t) — At(t) and is defined in Refer-
ence 178. Computation of the latter is straightforward using small argument and
large argument expansions, also given in the reference. Certain simple expressions
for jf are then obtainable.
(i) kR < Re(v +\) > 1 whence Re(-~0 > 1
jf - (i - z2y/2 _ Q/2 kR)z2/(l __ 2 2 } (47)

where z = (p + \)/kR
(ii) Re(p + | ) > kR > 1 whence Re(f) > 1
P ^ - j ( z 2 - 1) 1/2 (1 ~-z2/2kR(z2 - 1)3/2) + [(z2 ~ 1)1/2 (4.8)
exp(-4f 3 / 2 /3)]
Formulae 4.7 and 4.8 reflect clearly the gradual cut off phenomenon which is
characteristic of spherical modes.

Modal equation and the hybrid factor


The boundary conditions require that at 6 = du E^ = 0 and —H^ER = Y where
Y is the average slot admittance. A good approximation is given by
Y = - j(cot kd +l/2kR sin <90/r?o (4.9)
By using eqns. 4.1 and 4.2, the boundary conditions are
P?(Q\) = ~" mPI'A sin Bx (4.10)
pfipx) = Yv(v+ 1) = -mAp/sindt (4.11)
where
p™(dl) = (dP™(cosB)ld6)/P™(co$d)[d = et] (4.12)
and
f = -jYzolkR (4.13)
Multiplying eqns. 4.10 and 4.11 together yields
P™0i) [pJPidO-fpip+l)] = m202/sin2e1 (4.14)
Subtracting eqn. 4.11 from eqn. 4.10 yields
A - I / A = ?p(y+1) sin OJm]} (4.15)
Eqn. 4.14 is the modal equation for p, while eqn. 4.15 gives the hybrid factor A
62 Conical corrugated waveguides

in terms of the normalised wall admittance Y. It is noted that when F = 0, A = ± 1


corresponding to balanced HE or EH modes, respectively. In general, the dependence
of the modal eqn. 4.14 on R through j3 and possibly Y violates the 'separation of
variables' assumption. However, near the balanced-hybrid condition F ~ 0 ,
numerical solutions of eqn. 4.14 indicate that v is only very weakly dependent
on kR even for kR sin 0i as low as 3. On the other hand, 0 can vary widely with
kR, taking mainly real values for kR > v2 and mainly imaginary values for kR < P.
For intermediate values of kR, 0 is complex. This behaviour, of course, reflects
the well known phenomenon describing the gradual cut-off of spherical modes.
Of particular interest is the fact that unless the mode is a pure E or H mode (A =
0 or °°), the variation of j8 with kR results in a varying transverse field distribution
as can be seen from eqn. 4.2. As a further consequence, the orthogonality relation-
ship between modes at any given R is not exactly satisfied.
On returning to eqn. 4.14, this has been solved for the first few low order
modes under the conditions corresponding to kR<v2 or 0 ~ 1 and F = 0 . We
designate the solutions in this case by v^; n = 1, 2 . . . Approximate solutions
for the H E n , EH12 and HE12 modes are now discussed.

Approximate solution of the modal equation


For small values of 01 and large i>, the Legendre function Pv(cos 01) is well approxi-
mated by
O ^/o(O> + §)0i) (4.16)
#
where /o( ) is the Bessel function of zero order. It then follows that
^(cosflO - dPp(co$d1)/61 s - ( i ; + i)/ 1 ((i> + i ) 0 i ) (4.17)
and the modal eqn. 4.14, for m = 1,0 ~ 1, becomes

an0lplv(0l)*01plv(01)*x ~^. = +1 (4.18)


Jl[X)

where x = (p + \)d\.
Solutions of the above equation are simply given by JQ(X) = 0 or J2(x) = 0 for
HE l n or EH l n modes, respectively. Using the first zero of J2(x) and the second
zero of/ 0 (x), we obtain immediately eqns. 4.20 and 4.21.
Using the first zero of J0(x) we get for the HEn mode i^?^ 2-405/0!. This
approximation, however, is in relatively large error (>2-5% for $!>30°). The
formula given by eqn. 4.19 is found to have much less error and is thus considered
a better approximation for the HEn mode.
y? + £ ^ 2-405/(0! sin0O 1/2 (4.19)
po + i - 5.1356/0j (4.20)
po + i ^ 5-5201/0! (4.21)
They deviate from the exact solutions by no more than 1% for flare angles 0 t up
Conical corrugated waveguides 63

to 60 . The constants on the RHS of eqns. 4.19 to 4.20 are zeros of the Bessel
functions of zero and second orders. The dependence of v on Bx for lowest order
modes is to be found in Fig. 4.2(a) while Table 4.1 tabulates v as a function of 0 x.

HE 15

1-0 0-5r

0-5 £0*25

0^30

30 60
angle, deg
b

Fig. 4.2 a Solutions of characteristic equation for HExn and EHxn modes in corrugated
conical horns
modes in smooth wall horn
b Function fv+ (6) as a function of d
c Function fv_ (6) as a function of 6
64 Conical corrugated waveguides

Table 4.1 First root v of flv (Qx)=0 for HEn spherical mode

o? V 01 V V

16 8-1649 41 2-9982 66 3 •8126


17 7-6617 42 2-9216 67 ] •7851
18 7-2148 43 2-8487 68 1 •7585
19 6-8152 44 2-7793 69 1 •7327
20 6-4560 45 2-7132 70 1I -7078
21 6-1313 46 2-6501 71 1i-6837
22 5-8364 47 2-5898 72 ][ -6604
23 5-5674 48 2-5322 73 1[-6379
24 5-3212 49 2-4770 74 11-6161
25 5-0949 50 2-4243 75 1I -5949
26 4-8863 51 2-3737 76 1I -5745
27 4-6933 52 2-3252 77 ]1-5546
28 4-5144 53 2-2787 78 11-5354
29 4-3481 54 2-2341 79 11-5168
30 4-1931 55 21912 80 JI -4988
31 4-0483 56 21499 81 11-4813
32 3-9128 57 2-1103 82 ]I -4643
33 3-7857 58 2-0721 83 L-4479
34 3-6663 59 2-0354 84 11-4319
35 3-5539 60 2-0000 85 1L-4164
36 3-4479 61 1-9659 86 11-4014
37 3-3479 62 1-9330 87 1-3869
38 3-2533 63 1-9013 88 1-3728
39 3-1637 64 1-8707 89 11-3591
40 3-0788 65 1-8412 90 L-3458

Finally, it is useful, for the purpose of calculating radiation patterns, to separate


the transverse modal fields into their copolar and crosspolar components. We first
define appropriate orthogonal unit vectors:
ico = cos 0?0—sin 0/0 (4.22)
ic. = sin 0/0 4-cos0?0 (4.23)
Then, on solving for iQ and 1$ and substituting in eqn. 4.2 with m = 1, the trans-
verse modal fields are recast, apart from a constant, in the form:
REtI = UP + A)fU6)ho + i(F~ A)/J-(0)'(cos 2(fiieo
(4.24)
+ sin 2<j> iCI)
)/J()( 0cr
(4.25)
-sin20z' c o )
Conical corrugated waveguides 65

where f™±{d) are defined by


f?± («) = W? (cos d)/dd ± mP™ (cos 0)/sin fl (4.26)
For the HE l n modes, under the balanced hybrid condition A = 1, the transverse
electric field is mostly in the copolar direction @~ 1). The ratio of the crosspolar
to copolar component is seen to be maximum at <j> = n/4 and is given by

(4.27)

Graphs of fl+{6) a n d / ^ ( 0 ) as a function of 6 are shown in Fig. 4.2(b) and (c).

4.3 Radiation from corrugated conical horns by the spherical wave expansion method

The prediction of the radiation fields from conical horns can be accomplished by
the Kirchhoff—Huygen, the Laguerre—Gaussian expansion, or the spherical wave
expansion methods. We begin by describing the latter method which is as effective
in predicting the near fields as the far fields. The spherical wave expansion (SWEX)
method was first used with corrugated horns by Clarricoats and Saha [30] who
derived expressions for copolar radiation under balanced-hybrid conditions. The
method has been generalised by Mahmoud and Clarricoats [95] whose analysis
is reproduced below. Simple closed from expressions are given for both copolar
and crosspolar radiation fields from conical corrugated horns. The Kirchhoff-
Huygens and Laguerre-Gaussian methods are discussed in Sections 4.5 and 4.6.

4.3.1 The spherical wave expansion method


The basic spherical wave functions with satisfy both Maxwell's equations and the
radiation condition are given by eqns. 4.28 and 4.29.

H$\kR) (4.28)

0)/30)S* mfie + (mP™(cos 0)/sin


i2)
H'n (kR)/kR (4.29)
where m and n are integers.
By using these functions, one can expand E a and Ha which are the transverse
vector fields at the horn's aperture as

I (4-30)
i=i

oo

Ha = (1/FO) £ (*J^oU + &iW o i,+ 6,w eW ) (4.31)


66 Conical corrugated waveguides

where #j and bi are, as yet, unknown coefficients. In the above, the subscript 1 in
the m and n functions refers to the $ variation of the modes which is assumed
to be either cos (f> or sin <j>, since only the dominant modes are of interest in this
analysis. This subscript will be omitted in what follows for convenience.
To obtain the coefficients an and bn, one can use either Ea alone, Ha alone, or
both E a and H a . Expanding E a alone, eqn. 4.30 is cross-multiplied by non and men
successively and then integrated over the spherical aperture. Upon using well
known orthogonality relationships among the m and n functions one obtains
ir r2n( non )
J
° \m- en)
) (4.32)
where yn = 2Tin2(n + l) 2 /(2n + 1 ) . Similarly, for an Ha expansion, one obtains

Cn C2n(mon
\ J x

(4.33)
The aperture fields of the HEi or EHi modes are given by eqns. 4.24—4.25 in the
range 0 < 6 < Qx and they may be taken as zero outside this range. By using these
equations along with eqns. 4.22 and 4.23 in eqn. 4.32 we obtain for an E a
expansion

Tn-(0i)] (4.34)

where

Tn±(Bi) = £*' fl±(0)fln±(P)sin0d0 (4.35)

This integral can be expressed in closed form:

sin0d<9 = {sin01/O>(*' + l ) - w ( w + 1))}


(4.36)

where f™±(6) is defined in eqn. 4.26.


Conical corrugated waveguides 67

When w approaches v, the above integration is obtained by taking the limit w -> v
oftheRHS.

lfvm±(d)]2sindde = P^

Similarly, for an Ha expansion one obtains

Tn-(Oi)] (4.38)
In eqns. 4.34 and 4.38 the approximation j5 — 1 has been used, which is valid for
large apertures where kR0>v2.
At this point, one notices that these equations are not identical. This is because
Ea and Ha are not quite exact since they are assumed to be zero for 6 > #r, also
the horn air discontinuity is neglected. However, for large apertures (kR0 > v2) one
expects that the lower order terms in the expansion will be the most significant in
an expansion of the aperture field. For these termskR Q >n 2 and Hn®\kR0) = — j
H%\kR0), then under this approximate condition, the above equations are indeed
identical. Now using this condition and substituting from either eqn. 4.34 or eqn.
4.38 in eqn. 4.30, the electric field E p in the radiation zone is expressed in terms
of its copolar and crosspolar components:

(4.39)
S-(6,el)exp(-}kR)/(R/R0)

(4.40)
n=1

= Tn±(d1)'n/yn

^(cos (9:)
2n2{n+\f n(n + l)-v(y+l) l v

fvl±(ei)-v(v+ l)Pl(co$ 81)f*±(6l)] (4.41)


A slightly different formula from eqn. 4.39 is obtained by averaging out the
coefficients of eqns. 4.34 and 4.38. This results in a correction term sn inside the
summation of eqn. 4.40 where
sn = [1 +H}i2){kRo)l)Hni2){kRQ)}l2 (4.42)
Again it is noted that for the lower order terms for which kRQ ^ ^ 2 , sn tends to
68 Conical corrugated waveguides

unity. Numerical results confirm that for sufficiently large apertures, e.g. URQ
sin $i > 15, the correction term in eqn. 4.42 does not change the radiation pattern
by more than a fraction of 1 dB.

4.3.2 Frequency dependence ofcrosspolar radiation field


From eqn. 4.39 the crosspolar radiation field attains its maximum in the 0 = 45°
plane. Relative to the boresight copolar radiation level, the crosspolar field in
this plane is given by
i) = [(1 - A ) / ( l + A)] 5 . ( 9 , 9 0 / ^ ( 0 , 9 0 (4.43)
The key factor that determines the frequency dependence of C is that which
involves A. Looking^ back at eqn. 4.15 and noting that for all practical cases Yv
0 + 1) sin 6t < 1, A is approximated by 1 + \ Yv(v+l) sin 6X for HE l n modes.
Using eqns. 4.5 and 4.13 for F, eqn. 4.43 then takes the form

which shows that the frequency variation of C is mainly controlled by the resonant
behaviour of the slots. So let / 0 be the frequency at which kos = ir/2 or s = X0/4
and define 5 = (f/f0 — 1), then eqn. 4.44 becomes for the HEn mode:

= [(2-4O5)2sin01/01-sin2a1/4]f=f4^

tan TTS/2 1
<4 45)
'
where eqn. 4.19 has been used as an approximation for v. In this form the cross-
polar ratio is expressed as flare-angle dependent terms multiplied by a frequency-
dependent term. It is noted that Cdecreases steadily with increasing horn aperture.
The latter actually determines the bandwidth of the horn as is demonstrated in
Fig. 4.3 where C is shown as a function of 5 with korla as a parameter. The de-
pendence of C on the flare angle 6X is almost as sin dijdi as is seen from eqn.
4.45 and hence it is a slowly varying function of the flare angle in the range of
interest. The radiation patterns of copolar and crosspolar fields are given by S±
(6,61), respectively. These are plotted in Fig. 4.4 for 30° and 45° horns having
the same aperture: krla = 15. The levels in the figure correspond to a frequency
deviation of + 10% from / 0 (i.e. 5 =0.1) and for both horns normalization is
made relative to the on-axis copolar radiation level £ + (0, 6t). The crosspolar
patterns of the EHi2 mode in the same horns of Fig. 4.3 are shown in Fig. 4.4.
The gain G of the horn supporting a given mode can be deduced from eqn.
4.39. For simplicity let us consider the case A = ± 1 corresponding to balanced
HE or EH modes.
G - 4nR2\Ep\2/y0P (4.46)
Conical corrugated waveguides 69

P= 2nR20{ei [fl±(d)]2 sin 6d6ly0


Jo

Upon using eqn. 4.39, this is reduced to

G = 2\S±\2/C [fl±(d)]2sinddd (4.47)

A further reduction is obtained by using eqn. 4.37 and the modal equation fp±
(Q\) — 0 under the balanced mode condition.
-2(2P+
G = (4.48)

-40

m
Xf

r
M
|-50
-55

Uo
-0.16 -0.12 -0.08 -0.04 0 0.04 0.08 0.12 0.16
6=f/fo-1

Fig. 4.3 Peak crosspolar radiation of H£n mode in a 30° corrugated conical horn against
frequency with krx as parameter

Gains of horns supporting H E n and EHi2 modes are shown in Fig. 4.5 as
functions of 6X for £R 0 sin Qx = 15. For other values of kRo, it is useful to note
that the gain is proportional to kR0 under the large aperture condition.

4.4 Generation of higher order modes

Excitation of the dominant H E n mode in a corrugated conical horn inevitably


generates other higher order modes. Of these the EHi2 mode most seriously affects
the purity of the polarisation of the radiated fields. This is so since near the
balanced-hybrid frequency the HE l n modes for n > 1 make negligible contribution
to crosspolar radiation. Generation of the EH12 mode can occur at the throat,
also along the horn due to conversion from the H E n mode. Mode conversion
along the horn occurs either when the normalised slot admittance Y is varying or
if the horn has a non-linear profile or both. Dragone [56] has derived a simple
formula for mode conversion in a narrow flare-angle horn under the large aperture
70 Conical corrugated waveguides
o

-60
20 40 60
angle, degrees
Fig. 4.4 Copolar and crosspolar radiation patterns of HEXI mode at a frequency of f — 1-1 f0,
krx = 15, Oi =30,45
Copolar patterns for EH 12 mode

25

20- kR o sine 1= 15

15
ao EH
T3

10-

10 20 30 40 50 60
semiflare angle 8j degrees
Fig. 4.5 Gain of HEn and EHl2 modes against flare angle 6,
Conical corrugated waveguides 71
condition. In this section his restrictions are removed and the generation of the
EH12 mode is deduced in a horn of arbitrary flare angle both at the throat
discontinuity and due to HEn mode conversion along the horn. It is shown that
while Dragone's cylindrical mode analysis overestimates the crosspolar contribution
due to mode conversion along the horn, the throat contribution is significant,
especially for wide flare-angle horns.

4.4.1 Scattering at an abrupt discontinuity


Consider the canonical problem of scattering at a junction between two corrugated
conical horns characterised by semi-flare angles and normalised wall admittances
(\jji, F a ) and (0 l9 F b ). The geometry and definition of symbols are illustrated in
Fig. 4.6. The dominant mode is incident from the LHS and reflected and transmitted
modes are excited at the junction. Let (e£,h*) be the transverse vector field of the
forward nth mode on the LHS of the junction and (ejj,h^) the corresponding
fields on the RHS. A reflected mode on the LHS has the fileds (e**, — h**)as can
be deduced from eqn. 4.2^ byjioticing that for a reflected mode the following
changes apply: 0 -> |3* and A -» A*, v -> v* where the symbol * signifies the complex
conjugate operation.

Fig. 4.6 Junction between two spherical cones with flare angles \p, and 6 u radii Ra and

Now matching the transverse fields on both sides of the junction at R =Rh
andO<0<0!:

r,ef*exp.G0aA(0)) = I
1=1
(4.49)

(4.50)

where £a = J3afc and A(0) is the distance between the two spherical caps R = Ra
72 Conical corrugated waveguides

and R=Rh (see Fig. 4.6) and is given from the geometry by:
A(0) = /? a [l+2(sin Vi/sinfliXcos^!—sin^i cotl?!)
(cos^-cos^)]172-^ (4.51)
The special case \jji = 0, corresponding to a cylindrical guide, is of interest. In this
case A(6) reduces to
(4.52)
To obtain the reflection and transmission coefficients eqns. 4.49 and 4.50 are
vector cross-multiplied by h^ and e}j, respectively, and then integrated over the
spherical cap R—Rh,0<B < 0 x to give

<ea exp (-J0 a A(0)), h*> + £ T,<ea exp Q


i

= Tn+ I r,<e?,hS> (4.53)

(et h a exp (-jj5aA(0))> + ^ r , <e£, h, exp (j|8a A(0)» =

= Tn+ I ri<e^h?> (4.54)

for n = 1, 2 . . . where the normalisation <e^, h^> = 1 has been assumed.


<ea, hb> represents the inner product, given by

(ea,hb> = Jo # J o sin B AB (Re*xRhh)-iR

Now following Dragone [56] a simple solution of the above infinite set of
equations is obtained by assuming that the coefficients Tn,n > 2 andR n ,n > 1 are
all small quantities relative to T\. This is usually the case for a weak discontinuity
as in all cases considered here. Since the modes on either side of the discontinuity
are not exactly orthogonal in our case, the products <e£, h^>, / =£ n are non-zero, but
they are small relative to unity. Taking this into account and neglecting quantities
of second order of smallness, one obtains:
r , ^ i [<ef exp (-j]3 a A), h?> + <e?, h?,exp (-jjj.A)>] (4.55)
<et e x p (
r ~ ~ j f e A ) ' h ? ) ~ < e ? » h ! e x p ( ~ j f 3 a A)>
a
<et exp 0 « A ) , h?> + <ebu ht ^ b )

a
Tn = i [<e exp ( - j f e A), h^> + <e^, h? exp ( - j(3
- i 7", [<e^, h£> + <e^, h?>]; n > 1 (4.57)
Conical corrugated waveguides 73

It is constructive to note that these results reduce to those of Dragone's [56] when

4.4.2 Higher order mode generation at the throat


The throat junction is composed of a cylindrical waveguide with smooth walls on
one side and the conical corrugated horn on the other. This corresponds to the
case treated in the previous section with i//j = 0 and F a = °°. The cylindrical wave-
guide normally carries the dominant TE U mode which is incident on the junction.
The HE U is excited in the horn along with other higher order modes. A numerical
example is shown in Fig. 4.7 for a 30° horn with a slot depth of X0/4 at the centre
frequency for which korit = 3. The transmission coefficients Tx for the HEn
mode and T2 for the EH12 mode and the reflection coefficient Ft oftheTE n mode
are plotted in decibels as a function of the frequency parameter 5. It is interesting
to note that the peak of \T2\ occurs when |j3| of the EH12 mode is about a mini-
mum, i.e. when this mode is in its transition region between a purely propagating
state and purely cut-off state.

-25-

-0.16 -0.12

Fig. 4.7 Reflection and transmission coefficients as a function of frequency for junction
between smooth wall waveguide and conical horn, krx = 3-3, 6, = 30°

As the EH12 mode, which is generated at the throat, travels to the mouth of
the horn, it suffers a loss which depends on the length of the horn over which the
mode is effectively in a cut-off state. This loss is thus a function of both the
frequency and the flare angle and is plotted in Fig. 4.8 versus 5 for three corrugated
horns with Qx = 12°, 30° and 45° but with each having the same aperture at the
throat and mouth. The net crosspolar level of the EH12 mode is then obtained and
is shown plotted for dx = 30° in Fig. 4.9.

4.4.3 Mode conversion along a horn with constant flare angle


The second mechanism responsible for the generation of the EH12 mode is the
74 Conical corrugated waveguides

continuous conversion of the H E a mode as this mode Jravels along the horn.
As previously stated it is caused by either a change of Y or 0\ or both. While
these two causes can be treated analytically starting from the results of the previous

100-

-0.16 -0.12 -0.08 -0.04 0 0.04 0.08 0.12 0.16


<S = f / f o - 1

Fig. 4.8 Transmission loss of EHl2 mode from the throat to the aperture of a horn as a
function of frequency, kr^ — 3 and kria= 15
ad, =12°
bdx =30°
c 6. = 45°

-0.16 -0.12 -0.08 -0.04 0 0.04 0.08 0.16


<S=f/f o -1
Fig. 4.9 Levels of £H12 mode at the aperture due to the mode generated at the throat and
the mode converted along the horn
mode generated at throat
mode converted along horn. Parameters as Fig. 4.8

subsection, we shall concentrate here on mode conversion along a straight horn,


i.e. a horn with constant flare angle 0U so that the only cause of mode conversion
is the variation of Y with R.
Conical corrugated waveguides 75

Since the expected levels of converted modes are quite low, certain simplifying
assumptions can be made. All reflected modes are neglected and the reconversion
of the higher order modes is not considered. In the following it is convenient to
adopt the normalisation <en, h n >= 1. This implies that the modal fields in eqn.
4.4 be multiplied by l/Nn where Nn = <En, Hn>, and
Nl = <En, H n )/ J o = Or/2)[G3n + KWnK + \)L+n - 0n - A n )
<PnAn-l)Ln] (4-58)
where

\l [/,l±(0)] 2 sin0d0 (4.59)

and is obtainable from eqn. 4.37.


Now let the complex amplitudes of the various modes be given by An(R), n =
1 , 2 . . . At the throat R=ROy Ax = 1 and An = 0, n>\. Considering mode
scattering in an infinitesimal region between R=R and R = R + dR where Y
changes by dF = (dY/&R)dR, the transmission coefficients are obtained by using
the results of the previous subsection. A differential relation An(R), n > 1 is

dAn(R)/dR = -)Wn(R)An(R)+^Al(R);n>l (4.60)


OKi
where dTn is obtainable from eqn. 4.57. After some algebra, an explicit expression
for dTn is
n
WvxPvn
\ • n /-df F d?,

, + (A, + A n )d? 1 ] (4.61)

g in
NxNn Nx
WvaPvb
Gab = {Vh-v) [ft* sin BiWiYJb - rb/ft.) + AabJ (4-62)
where
A ab = 0 W A a A b ) t t t . A b ^ -f b A a «; a )(l + A a A b )] V,
+ fafbAab(^b-^a) (4.63)
an =
d fa,b 1 ~ AM>> where all quantities are to be calculated at R, while the quantity
76 Conical corrugated waveguides

dFis obtained fromdf =?(/? + d # ) - F ( i ? ) and likewise for all other differentials.
The amplitude of the dominant mode Ax{k) is generally given by

At(R) = exp [ - j f t f 0i(/?)dK (4.64)

where /3j should, in principle, be different from |3i in order to account for the
loading of the higher order modes. However, in our case, because of weak coupling,
it is reasonable to take ${ = plt Now integrating eqn. 4.60, one gets

An(R) =\* A,{R'){dTJdR')^V -j*f* Pn(R)dR} dR' (4.65)

Substituting from eqn. 4.64

An(R) = exp (-#„)("* (dTJdR')exp - j * ! " * (&-ft,)d/i dR'

(4.66)
where

<Pn = k [%,&) dR (4.67)

Now eqn. 4.66 along with eqn. 4.61 give the amplitude of a converted mode at
any point along the horn. The integration can be performed numerically. However,
a better way is to approximate the quantity { ( ^ R ) 2 ^ —@n)} as a linear function
in (1/kR) and (dTn/dkR) as a polynomial in the same variable whereby the
integration can be readily expressed as sum of complementary error functions. In
certain cases it is necessary to break the interval of integration into two or more
subintervals where the above approximations are applied in each subinterval. This
procedure ensures an accurate computation of the integral.
Some insight into the mode conversion process to the EH12 mode can be gained
by examining eqn. 4.61 under the conditions kR>Vi, v2 and Y < 1. Then the term
involving dFis dominant and dT2 becomes

Y (4.68)

N2n - TtiPlicosB^viv +l)sin0! (4.69)


Using eqn. 4.69 for N\ andi¥ 2 under balanced hybrid conditions, and eqn. 4.19
and 4.20 for vx and p2, dT2 further simplifies to the following withx 0 — 2-405 and
x2^ 5-1356:
ring,)-"'
x
" sin
Conical corrugated waveguides 77

Now dF is nearly proportional to (1/kR) which in turn is proportional to sin 6X


assuming the horn's plane aperture is kept constant. The net result is that dT2 is
very weakly dependent on the flare angle 6\. Now, turning attention to eqn. 4.66,
and identifying the term in brackets as 0 1 2 , the real part is the difference of phase
delays of the first and second modes along the horn. This is obviously minimum
for the shortest horn (#j =90°), provided that the throat and mouth apertures
are kept constant. If this phase difference is less than n for 6X = 90°, the magni-
tude of A2 (Rm) will tend to increase if Qx is reduced, it attains a maximum when
this phase is — TT and goes into a damped oscillation with a further decrease of
di. This behaviour is illustrated in Fig. 4.10, which shows the level of EHi2 mode
converted along a conical horn versus the semi-flare angle $i with/// 0 = 1-1. Also
shown in the figure, curve b, is the upper bound predicted by using the cylindrical
mode theory of Dragone from which it can be seen that this theory overestimates
the role of mode conversion in wide flare-angle horns.

-35

-60
10 20 30 40 50
semiflare angle 8, ,degrees
Fig. 4.10 Level of EHl2 mode converted along a horn as a function of semi-flare angle B le
frequency = 1-1 f0, krlt = 3, krlQ= 15, s = K/4

Numerical results for the converted EH12 mode at the mouth of a horn versus 5
are shown in Fig. 4.11 for dx =30°, curve c. Compared to the level of the EH12
mode generated at the throat, the converted mode along the horn, curve b, is
lower over a broad band of frequencies but the reverse would be true for narrow
flare angles. For the 30° horn the mode generated at the throat dominates at the
higher end of the frequency band while the converted mode dominates at the
lower end. This behaviour is mainly attributed to the higher transmission loss of
the throat generated mode in the narrower horn.
One effective way of reducing the level of the EH12 mode in wider flare-angle
horns is to start the horn at a smaller throat aperture. This increases the total
transmission loss for an EH12 mode which is generated at the throat. But, eventually
there would be an increase in the level of the EH12 mode generated at the throat.
A remedy for reducing the return loss in narrow flare horns is to use the X/2 deep
78 Conical corrugated waveguides

corrugations at the throat. The corrugation depth is gradually reduced to X/4 at


the mouth of the horn (see Section 5.5). This procedure can also result in some
reduction of the EH12 mode.
The overall crosspolar peak radiation level of a 30° horn with corrugations of
X/4 depth at mid-band is shown in Fig. 4.11. For this corrugated horn korit = 3
and korla = 15. It is seen that, in this case, mode conversion along the horn is
insignificant over all the band. It is dominated by the mode generated at the throat
at the higher frequencies and by the intrinsic crosspolarisation of the H E n mode
curve a, at the lower frequencies. Experimental data for the same horn are shown
for comparison and a satisfactory agreement with the overall theoretically pre-
dicted level is observed.

-40 r
CQ

i -50-

1-55
(A

8
-60
-0.16 -0.12 -0.08 -0.04 0 0.04 0.08 0.12 0.16
<S=f/fo-1
9 io n
f,GHz

Fig. 4.11 Level of peak crosspolar radiation for a 30° semi-flare angle horn due to H£n
mode, curve a; EHl2 mode generated at the throat, curve b; EHl2 mode converted
along the horn, curve c
krlt = 3, krla = 15, vertical bars show experimental data

4.4.4. Mode conversion along a horn with variableflareangle


In some applications there is a requirement to produce a compact corrugated feed
with a plane phase front in the aperture. This must be of sufficient size in terms
of wavelengths, so as to produce a narrow beamwidth pattern. One way to create
such a feed is to utilise the profiled horn of Fig. 4.12(a). Such horns have been
fabricated but have been found to contain higher-order modes, generated because
of the changing profile. Generally their level exceeds that which arises in a linear
profile horn as discussed in the previous section.
Based on the work of Mahmoud [177], we shall now determine the effect of
mode conversion in a profiled horn as shown in Fig. 4.12(a). It is assumed that the
dominant HE n mode exists in the throat at z = 0 , also that the walls can be
modelled by a constant surface admittance. In practice the admittance may be
varied as well as the flare angle but for purpose of discussion the effects are
separated here.
Conical corrugated waveguides 79

From eqn. 4.66 we have

(4.71)

where

to = J ft(z')ds, / = 1 oxn
j3,- = the modal complex phase constant
ds = dz'(l + (dr/dz') 2 ) 1/2
and dCn(z) is the differential conversion from the dominant mode to the nth
mode per unit amplitude of the dominant mode in the region bounded by z — dz and
z. Now, the crux of the problem is to derive dCn(z) in a closed form in terms of
the local profile parameters and the local modal parameters of the modes involved.

Fig. 4.12 a Profiled corrugated horn


b Flare angle discontinuity

Let us consider a junction between two spherical caps of the horn at z — dz and z
with corresponding flare angles B\ — dd\ and 0 l s respectively (Fig. 4.12(b)). The
dominant mode (e?, h?) with a unit amplitude is incident at z — dz and we seek the
transmitted dC n , n > 1 emerging from z due to mode conversion. Let the modal
vector fields at the spherical cap at z be (e^, h)j), n = 1, 2 . . . .
These modes are the normal modes of a uniform cone having the local geomerical
and electrical parameters of the profiled horn. From eqn. 4.57, dCn for n > 1 is
80 Conical corrugated waveguides

given by

dCn a H<e?exp(-j/3!A(0)),h^ + <e£,


-i[<ei,J>S> + <eS,h^>] (4.72)
where the product <f,g) is defined as the integration of f *g over the spherical cap at
z. A(d) is the radial distance between the two spherical caps at z — dz and z (Fig.
4.12(b)). The second term in eqn. 4.72 arises due to the fact that, in general, the
modes are not exactly orthogonal. From the geometry, one can easily prove the
following relations:
A(0) = a(z) cosec2 6i(co$ 6-cos d^ddt (4.73)
d0/d0i = sin dfsin 0x (4.74)
Next, the 0 and <p dependence of the (ef, h?) atz — dz are displayed explicitly as
e?(0-dfl,0) = eUd,(i))-dd1-(deUd,<f>)/dd)'($mdl$me1) (4.75)
and a similar expression holds for hf (6 — dd, 0). Furthermore, since §\ A(9) < 1,

exp(-j0}A(0) ^ l~jff
(cos 6 — cos 6i)ddi (4.76)

Substituting from eqns. 4.75 and 4.76 in eqn. 4.72, dCn is recast in the form
dCn(z) = Qi-Q2cosoc91dd1-Q3)P&1(z)cosec2eldel (4.77)

where
Gi = £ [<(e! - e?), h^> + <e*, (h? - h?»] (4.78)

Q2 = | [<sin (9ae?/3(9,hS> + <e^, sin 6 dh?/30>] (4.79)


and
g 3 = I [<(Cos 0 - cos 00ef, h^> + <e^, (cos 0 - cos 0 0 h?>] (4.80)
where all fields are calculated at the spherical cap at z with arguments 6 and (p.
The superscripts a and b signify that the modal eigenvalues correspond to the local
parameters at z — dz and z, respectively. To continue, resort is made to the explicit
field expressions for (e n , h n ) given in eqns. 4.24 and 4.25, which reduce eqns.
4.78-4.80 to
Gi = A{LM^t)~~LM,vhn)} + B{L^uvl)
-LM,Vn) (4-81)
Q2 = ASM, vt) + BS-(v*l9 p*) (4.82)
Q3 = A T+(v*l9 ^ ) + BT-(vl A) (4-83)
Conical corrugated waveguides 81
where

L+ (?,<*>) =f01 /,, + (0)/ w + (0)sin0d0 (4.84)


Jo

(4.85)

+ (0) cos0 sin 0d0 - cos fl^ + ^ w ) (4.86)

/„ + (<?) = bPl (cos 0)/d6 ± (sin ey'PUcose) (4.87)


ft»±(0) = -(cos0Tl)/ i ; ± (0)-i;(i'+l)P£(cos0)sin0 (4.88)
The coefficients A and J5 are given by

(4.89)
JV, = TT/2 [(ft + A,)(l + PiAdL+fo, pt) + (ft - A,)(l - ft A,)
L
-(Pu vi)] (4.90)
where ft =ft/A:is the normalised phase constant of mode / relative to the free
space wave number, A,- is the normalised modal hybrid factor and Nt is a modal
normalisation factor defmed previously in eqn. 4.58.
The integrations in eqns. 4.84—4.86 can be obtained in closed forms, then dCn
(z), given by eqn. 4.77, is now completely derived. In order to get some insight
into the behaviour of dCn(z) with the change of the flare angle 6U let us consider
the simple case of conversion to the HE12 mode under the condition kr(z)^> 1 so
that both $i and $n are close to unity. Let us also assume that the balanced hybrid
mode condition holds, i.e. the wall admittance Y — 0, hence Ai = An = 1. With
approximations

CB(z)/3fl, ^ S W f l J s i n f l . ) [i-j^jsinfl,)(30,^/2x1)] (4.91)


X\ X2 Sltl u i/u 1

where xt = 2-4048, x2 = 5-5201 are the first two zeros of the Bessel function Jo
(x). For typical values of kr = 10, Bx < 2 5 ° , it is seen that dCn(z)/d6i is almost
a constant. The net conversion level determined by eqn. 4.72 is then highly
dependent on the integral phase factor, which, in turn, depends on the horn's
profile. Illustrative numerical examples showing this dependence are given below.
The behaviour of dCn(z) for the EH l n modes, is quite different from that for
HE l n modes, namely if the wall admittance Y=0 everywhere, dC n (z) = 0 ir-
respective of the flare angle variation. This is seen from eqn. 4.89, since then
Ax = An = 1, hence A = B = 0. So, when Y varies around the zero value, dCn(z)
can be shown to be
82 Conical corrugated waveguides

dCn(z) = -)(KydY + KddYd0) (4.92)


where dY is the change of the wall admittance (normalised to free space admittance),

-x\
and

__ 2x 1 (sin0 1 /e 1 ) 1/2 f x\ . 2krB\


(4.94)
kr{x\ —x\ sin BiJBx) \x\ —x\ sin Bx

where x3 = 5-1362, the first root of J2(x). The above applies for kr > 1 and Bx <
25°.

Numerical examples
A class of cosine squared profiles has been numerically investigated. Referring
to Fig. 4.12(a), the cross-sectional radius rx (z) is given by

-rlt = 2(rla-rlt)cos2{Trl2(l-zl2L1)}l(l+y)...Q<z<Ll

1+7
...la <z <L (4.95)
where the parameter y = L2/Li and determines the location of the maximum
flare angle $i of the profile which occurs at z = Li =L/(1 + 7) (see Fig. 4.12(a)).
The value of B\ is, however, independent of 7 and is given by
tan£ = dr/dz\z=Li = 7i(rla-rlt)/2L (4.96)
r
\{z), 6i(z) and dBi/dz are all continuous %\z—Li as can be verified by differen-
tiating eqn. 4.95 once and twice. The case 7 = 1 corresponds to maximum flare
angle Bx half way between the throat and the mouth, while for 7 ^ 1, Bx occurs
closer to the thro at/mouth, respectively.

HEn-HEn mode conversion


A horn having the profile of eqn. 4.95 with korit = 3*0 and koria = 12*0 is con-
sidered. In order to concentrate our study on the effect of flare angle variation,
the slot depth is taken constant along the horn. This does not mean, though,
that the wall admittance Fis constant, since the later depends also on ka and sin 0\.
The length of the horn is varied between k0L = 15 and 40 in Fig. 4.13, with 7 =
L2/Li = 1-0 and 0-4, and the mode conversion is computed using the above analysis.
As expected, the level of the converted HE12 mode at the mouth decreases with
increased L. More important is the decreased level of conversion for 7 = 0-4 which
corresponds to a maximum flare angle $1 closer to the horn's mouth than to the
throat. Explanation for this decrease can be obtained by studying eqn. 4.71. Near
Conical corrugated waveguides 83

-18 -

Fig. 4.13 Level of HEl2 mode as a function of normalised length of horn for L2/Lx = 1-0 and
0-4. The maximum flare angle d x is also shown as a function of kQL:krxf = 3-0 and

-20 -

Fig. 4.14 Level of HEl2 mode as a function of position along the horn giving by krx for k0L
= 27.0. Geometry as Fig. 4.12. Parameter L2/Lx
84 Conical corrugated waveguides

the point 6X of the profile, d02 changes from positive to negative for z ^ L ,
respectively. This means that dCn changes sign, and, if the phase integral term in
eqn. 4.71 is slowly varying, a good deal of cancellation occurs, reducing the overall
level of conversion. Now, since (j8i — j3n) decreases as krx (z) increases, the condition
of cancellation is better satisfied as the location of 0x approaches the horn's mouth,
or as 7 is reduced. To support this argument, the conversion level of the HE i2
mode is plotted in Fig. 4.14 as a function of kri(z) for 7 = 1 , 0-8, 0*5 and 0-4.
It is seen that the level of the HE12 mode reaches a maximum then drops near
the 61 point. For 7 = 1 , this drop is followed by a sharp increase since after this
point the differential conversions almost add in phase, as (fii — j8n) becomes small.
For 7 = 0*4, the process of cancellation extends over a wider region and gets
closer to the mouth, hence followed by a relatively smaller increase in HE12 power.

-18

^-22-

£-26-

-30 -

-0-08 0-12

Fig. 4.15 Level of HEl2 mode as a function of f/f0 where f0 is the frequency where slot depth
= X/4. Parameters as Fig. 4.14

Obviously dx cannot be located too close to the mouth since then the flare angle
will have to be reduced from 6\ to a low value at the mouth in a very small
longitudinal distance, which creates a sharp discontinuity. An optimum value for
7 in our present example has been found to be y = 0-4. A general conclusion may
then be drawn. A profile which has a maximum flare angle in a region of sufficiently
high krx (say, krx > 9) is likely to have low mode conversion due to phase
cancellation occurring in this region. It is to be noted that this cancellation will
have a broadband width, since it mainly depends only on the profile shape. This
broadband behaviour can be seen in Fig. 4.15, in which the conversion level is
shown versus the deviation of the frequency/ from f0 at which the slot is a quarter
of a wavelength deep. Generally, the conversion level increases with frequency
since the total attenuation of the HE12 mode along the horn is reduced with
Conical corrugated waveguides 85

increasing frequency. It is also noted that mode conversion to the HE12 mode is
mostly attributed to flare angle variations, and to a much lesser extent to changes
in wall admittance Y. Actually, for a linear horn of the same length and aperture
as for the horn of Fig. 4.15, the mode conversion due to 8Y has been computed
and found to be only of the order of — 50 dB or less.
-38

-r-42 "

x
LU

-50

Fig. 4.16 Level of EHl2 mode as a function of normalised length of horn for L 2/L t = 1-0 and
04. krlt = 3-0andkria= 120

-48
-0-04

Fig. 4.17 Level of EHl2 mode as a function of f/f0. Parameters as in Fig. 4.14
86 Conical corrugated waveguides

HEn-EH 12 mode conversion


Results for conversion levels of the EH12 mode for the same class of horn profiles
are given in Figs. 4.16 and 4.17. The behaviour of the level of conversion with
k0L and y = L2/Li in Fig. 4.16 is generally the same as in the case of HE i2 mode,
but here, considerably lower conversion levels occur. This may be attributed to the
heavy dependence on SY for the EH12 mode. This dependence is manifested in
Fig. 4.17, where a deep minimum of the conversion level occurs near the centre
frequency where 5 Y is very small.

4.5 Radiation from corrugated horns by the Kirchhoff—Huygen method

The aperture fields of a conical corrugated horn are given in eqns. 4.2 a-d. The
radiation from the spherical constant-phase surface in the aperture can be obtained
directly by means of a Kirchhoff—Huygen integration. Following Clarricoats and
Saha [30] we express the radiation field at P, E p (R', 0', $') as

- zQ x (/„ x H t a n g ) exp QkR0' iR' ds) (4.97)


are
iR', in are unit vectors and the fields i^ang* #tang tangential to the surface of
the spherical cap. On replacing the factor outside the integral by c, we obtain,
after some algebraic manipulation:
[- Ee cos {<j> - 0 ) - £ 0 cos 9 sin (4>' - <}>)
0 JO

+ zQHQ COS 9' sin (0' — 0) — z0H(j) {sin 6 sin 6' + cos 6 cos 6' cos
O' - 0)}] exp GfcK0{cos 0 cos 6f + sin B sin 6' cos (0' - 0)}]
sin 0d0d0 (4.98)

°l [Eecos B' sin (0' - 0 ) - £ > { s i n 0 sin 0'

+ cos B cos B' cos (B' - 0)} + zQHQ cos (0' - 0)


+ ZQHQ COS B sin (0' — 0)] exp [j&/?o{cos 0 c o s $'
+ sin 0 sin B' cos (0' ™ 0)}] sin 0d0d0 (4.99)
Let a = kR0 COS B' and b = kR0 sin B\ and let
Ee(09<l>) = - g ( 0 ) C c o s m 0 (4.100)
=
M0,0) K0)C sin m(p (4.101)
On substituting for £?# and £0, we obtain, after performing the integration and
Conical corrugated waveguides 87
combining the constants,

Epd(d',(l>r) = Dcosmcp'^1 [2jg(d)Jm(b sin 6) sin 6 sin 6'

+ Q)m +l
Jm+i(b sin 0) {g(9)(l + cos 6 cos 0f)
+ h(d)(cos 0 + cos d')} + /m~lJm_x (b sin 0)
g(6)(l + cos 0 cos df) ± h(6)(cos 0 + cos d')}]
eQacosd)sin6d6 (4.102)
r
The integral for Ep(f)(6 , 0') is similar to that for EpQ except that g(6) and h(6)
are interchanged and cosm0' is replaced by — sin m0\ When A = + 1, we have
mP™ (cos 0) dP™
/ (fl) + (4.103)

When A = — 1, we have
dPf(cos^) mP™ (cos 6)
g(6) = h(B) = / - (fl) = - ^ - ' ^ - — (4.104)
U17 S i n (7

Form = 1, A = 4- 1
Epd = ± Fr(0') + }Ftf'ffi 0' (4.105)
0

and
G r (^ ? 0') = (1 4- cos 0)(1 + cos 6f) sin 0 ^ s (a cos 0)Jo(b sin 0)}

- {(1 - cos 0)(l - cos 6r) sin 6^ (a cos 0)J2(b sin 0)}
+ {2 sin 0' sin 2 a^ s (s cos 0)Jx(b sin 0)}
From the above equation, we observe that, under balanced hybrid conditions,
pattern symmetry exists about the axis 6' = 0. A maximum in the boresight
direction d' = 0 occurs for HE lri modes, while for EH l n modes there is a null
on the boresight, and similarly for all modes with m>\. There is no cross-polarised
component of radiated field if the aperture field is linearly polarised.

Comparison between Kirchhoff-Huygen and S WEX methods


In Fig. 4.18, we compare copolar radiation patterns predicted by the Kirchhoff—
Huygen and modal expansion methods for dx = 50°, 60° and 70°; the agreement is
excellent. We have studied the convergence of the series and have found that for
various observation angles, the pattern converges to within 1% when NakR0. A
good approximation to the final pattern is also obtained with only about six terms
88 Conical corrugated waveguides

70°16

-30
10 20 30 40 50 60 70 80 90
9,'deg
Fig. 4.18 E plane radiation patterns for wide angle corrugated horns
modal expansion method
Kirchhoff— Huygen method Parameters 61 and kR0
or

-10
CD
•o

|-20
a

t. . 3 0 -

-40

-50 _L J_
0 20 40 60 80
9,deg
Fig. 4.19 Near-field of conical corrugated horn. Parameter normalised distance kR from horn
apex. kR0=20
Conical corrugated waveguides 89

in the series, especially for wide flare-angle horns. It is also possible to use the
above method to determine the near-field of corrugated horns as shown in Fig.
4.19, computed using spherical wave expansion methods, (see also Chapter 5).

10 20 30 40 50 0 10 20 / 30 40 50
e'deg
a b

50

Fig. 4.20 Radiation patterns of 30° semi-flare angle corrugated horn


x H plane experimental o E plane experimental points -Theory using spherical
wave expansion method a 8-5 kHz b 9-0 kHz c 11 0 kHz
90 Conical corrugated waveguides

4.6 Comparison of predicted and measured radiation patterns

The theory described in the preceeding sections has now been widely validated
for corrugated waveguides and horns, the latter with di up to about 75°. Results
for a corrugated horn with 30° semi-flare angle are shown in Fig. 4.20, and further
discussion of experimental results are to be found in Chapter 5.

4.7 Radiation from corrugated horns by cylindrical waveguide approximation

As we have seen in previous sections the copolar and crosspoiar radiation patterns
of conical corrugated horns can be predicted using spherical wave expansion and
Kirchhoff— Huygen methods. However, this is necessarily approximate because the
analysis must assume that the surface admittance of the corrugations is independent
of position along the horn. Only under this condition is the spherical wave
equation in a conical co-ordinate system separable into independent parts. A
second approximation exists for the wave admittance —H^jE^ at the boundary
0 = $i precluding a study of the slot geometry.
None of these assumptions is serious as far as the copolar radiation pattern
prediction is concerned, and many workers have obtained good agreement with
experimental measurements. The prediction of the crosspoiar radiation pattern
for narrow flare-angle horns requires a more precise theoretical analysis because
the geometry of the slots is a crucial parameter in determining the peak cross-
polar power level. The analysis of an open-ended corrugated waveguide, including
space harmonics, enables the influence of slot geometry to be studied. The horn
aperture is treated as an open-ended corrugated waveguide and the waveguide
aperture fields computed. A simple spherical phase factor, dependent on the
horn flare angle, is then added to the aperture fields and the radiation pattern
computed. The method is valid certainly up to flare angles of 20° and even for
larger angles the method gives a good approximation to the measured patterns.
We show that this approach is justified by comparing the characteristic equation
and aperture fields of narrow flare-angle horns and waveguides. We start by showing
that the aperture fields, expressed in cylindrical co-ordinates, for a narrow flare-
angle conical corrugated horn are identical to the aperture fields for a corrugated
circular waveguide with the addition of a spherical phase factor.
Consider the case when the slot depth is such that the wave admittance, H^/ER
at the mouth of the slots is zero; this is the balanced-hybrid condition. The charac-
teristic equation for the spherical-hybrid mode in a narrow flare-angle conical
corrugated horn with a small semi flare angle 61, and a large normalised slant
length, kR0, is given by

adl ,
(4.106)
Conical corrugated waveguides 91

where RBX = ru a = V M ^ + 1)}, ^ is the order of the Legendre function, and the
Bessel function approximation to the Legendre function has been used. For
comparison the characteristic equation for the corrugated circular waveguide,
radius r l5 is given by

Krx
= -0 (4.107)

Hence eqns. 4.106 and 4.107 are identical when a$i is equated to Kru implying
identical wavenumbers.
We now consider the horn aperture fields at the balanced-hybrid condition.
The aperture electric fields defined over a spherical cap for the corrugated conical
horn for small 61 and large kR are given by

5L. I I" »V""U „ i -U - ^4.1UOj

1 - 1 ^ 1 | ^ L ^ _ - |/0e-jre" (4.109)

The transform from spherical co-ordinates (R, d, 0) to cylindrical co-ordinates


(r, (p, z) is given by iR = zr sin 6 + z% cos fl, ie = ir cos ^ — iz sin 0 and z0 = z^.
Mien 0 is in the range 0 < \6\ <6X and 6X is small, eqns. 4.108 and 4.109 become

Expressing the phase factor in terms of the radial r component in the aperture,
Fig. 4.21 yields:
e-jfe* = e-ik(z1 + A) (4.111)

By moving the phase reference to the plane aperture z = z 1 ? and expressing A in


terms of r, the phase factor becomes

exp -jfcft| 1 -Vl - I^-J J U e^« (4.112)


Hence, the aperture electric fields for the narrow flare angle conical corrugated
horn, expressed in cylindrical co-ordinates, are given by
92 Conical corrugated waveguides

Comparison of these equations with the aperture electric fields for the corrugated
circular waveguide shows them to be identical for OLQI =Krx, if they are multiplied
by the spherical phase factor 0(r).

Fig. 4.21 Geometry for spherical phase cap

-72 -36 0 36 72
angle, degrees
Fig. 4.22 Measured and computed copolar radiation patterns for 12 degree semi-flare angle
corrugated horn
measured
computed using waveguide approximation
x x x x computed with spherical phase cap
rx = 63-5 mm, s = 9 mm, b = 7 mm, t/b = 2-5 mm

This analysis is also valid for the more exact space harmonic analysis where
the fields in the waveguide are represented as a series of space harmonics, and the
fields in the corrugations as a series of evanescent radial modes. The method is more
Conical corrugated waveguides 93

-60 -30 0 30
angle, degrees

Fig. 4.23 Measured and theoretical (K = L= 1) (see Section 3.3) crosspoiar patterns at
10 GHz for horn of Fig. 4.22

-30

m
•o

o
a

2
Q.
HE 12 cut-off

-50
8 9 10 11
GHz

Fig. 4.24 Measured and theoretical (K-L = 1) (see Section 3.3) peak crosspolarisation for
horn of Fig. 4.22
94 Conical corrugated waveguides

complicated than the surface admittance method but it does allow the exact profile
of the corrugations to be specified.
Thus the radiation pattern of a narrow flare-angle conical corrugated horn can
be represented by expressing the aperture fields as those of the HE n mode
(including space harmonics) of the equivalent waveguide, with the addition of a
spherical phase factor. The far-field radiation pattern can then be calculated using
an aperture field method.
Fig. 4.22 compares the measured copolar radiation patterns for a 12° semi
flare-angle corrugated horn with theoretical points predicted using the waveguide
approximation described here. Excellent agreement is obtained using the waveguide
approximation, with the first sidelobe of the 12° horn being accurately predicted.
Fig. 4.23 compares the measured crosspolar radiation pattern for the 12° semi
flare angle corrugated horn with the theoretical prediction using the method
described here. The peaks of the measured cross polar patterns plotted as a function
of frequency are shown in Fig. 4.24 and good agreement between experiment
and theory is obtained up to a frequency of 10*4 GHz. Above this frequency the
measured pattern becomes higher than that theoretically predicted and this is
attributed to coupling into the HE12 mode.

4.8 Radiation from conical corrugated horns by means of the Laguerre-Gaussian


expansion method

Although the spherical-wave expansion method provides a means to predict ac-


curately the near-field and far-field of a conical corrugated horn, when the semi
flare angle is small and the horn length large, in terms of wavelengths, many terms
are required in the expansion. With present generation mainframe computers this
presents little problem but there is attraction in a method which potentially brings
the prediction method for copolar patterns within the range of smaller computers.
Such a method is the Laguerre—Gaussian expansion investigated by Bitter and
Aubry [17]. The method offers another advantage for millimetre wave antenna
designers using Gaussian optics since the first term in the series expansion
corresponds to a Gaussian beam mode.
Under balanced hybrid conditions we have for the Cartesian component of the
aperture field Ex in a horn of length L

E*(r9O) = Jm(Kr)exp ( ~ j f"T ) cos *!0 (4.115)

We note that m = 0, corresponds to the HE n mode, m = 1, to the TMOi mode.


Next expand Ex as a Laguerre-Gaussian series

r, 6) = exp / -j - ^ \ J / M C ' , (4-116)


Conical corrugated waveguides 95

where

amplitude

-10

amplitude

-30

-40
4 6
degrees
Fig. 4.25 Radiation pattern computed using Laguerre-Gaussian theory [17]
theoretical {N = 9), d1 = 5.8 Ro = 2487 mm, Distance in front of aperture
= 1300 mm o o o o H plane measured x x x x E plane measured a 11-1 GHz b 14-25
GHz

L™ is the generalised Laguerre polynomial of order n and m and H^ is a parameter


defining the so-called beam waist of the Gaussian beam. Note that the first term
of the series with m = 0 (HEn mode) is the Gaussian function, while with m = 1
(TEOi mode) the first term is the derivative of the Gaussian function.
96 Conical corrugated waveguides

To proceed it is first necessary to select a value for W which maximises the


quantity 17 where
N Nmi2|/mi2

V= I ^ H (4-117)
and J ^ is the function Jm(Kr),r<rx.
To obtain the radiation field in cylindrical co-ordinates referred to the horn
aperture, the Kirchhoff-Huygens integration over the horn aperture yields

N
/?p(p,0,z) = fAnBexV(-)kz

exp [- jfc(l - B)p2/2z] (4.118)


where

7
27z ~ PV
L = length of horn.
A useful simplification can be made by taking only one term in the series, thus
for the HEJI mode

j — I (4.119)

Experience also suggests that H; = 0-64r1 optimises eqn. 4.117. Physically the
radiation behaviour of the corrugated horn is approximated by a Gaussian beam
with beam waist Wjust less than 2/3 of the aperture radius.
Bitter and Aubry have applied the Laguerre—Gaussian method and have compared
predicted copolar patterns with those obtained experimentally. Fig. 4.25 compares
far-field amplitude and phase measured for a 5-8° semi flare-angle horn with
predictions with N=9. Clearly the agreement is excellent. Obviously the patterns
could be obtained by the other methods described in Sections 4.3, 4.5 and 4.6.
However, we should note that for accuracy with the spherical wave expansion,
a very large number of terms would be required. Were that number, taken to be
equal to kL in the example of Fig. 4.25, 581 terms would be required, which is
obviously numerically unacceptable.
Chapter 5

Design of cylindrical and conical


corrugated horns

5.1 Introduction

This chapter is concerned with the factors which influence the design of conical
corrugated horns. We start by surveying these factors, then study the electrical
specifications and the procedures needed to design the horn to satisfy the specifi-
cations. A large number of design curves are included. The information needed to
design various types of corrugated horns is summarised. Finally computer pre-
diction is discussed.
There are a wide variety of antenna applications where a corrugated horn is used.
Although it is occasionally used by itself, for instance as a gain standard, it is
usually used as a feed for a reflector antenna. The desirable characteristics, which
have been discussed in the previous chapters, all make the corrugated horn an ideal
candidate for use as a feed. These uses can be divided into two types, prime focus
reflectors and multireflector antennas, the Cassegrain and the Gregorian reflectors
being the main uses, though sometimes more than two reflectors are used as in the
beam waveguide feed system for large earth stations of the INTELSAT-A class.
Millimeter wave radio astronomy antennas also often use more than two reflectors.
The increasing sophistication of spacecraft antennas is one area where new designs
have evolved, such as the compact profiled horns and multifrequency horns. The
types of corrugated horn together with typical uses are listed in Table 5.1. They
will be dealt with in more detail in Section 5.8. The reason for identifying the
different types is that each type has common characteristics and problems as far as
the designer is concerned. The larger aperture narrow flare angle horns, which form
the largest single class of horns built so far, are the easiest to design because the
large diameter makes the horn much less sensitive to crosspolarisation and mode
conversion problems. The small aperture horns have a performance which is strictly
limited by the aperture diameter. The designer must take into account the effects
of the flange which is inevitably present at the aperture of the horn. The bandwidth
of this type of horn is often a problem. The wide flare angle horns (sometimes
referred to as scalar feeds) are potentially attractive because the copolar radiation
pattern depends on angle rather than aperture size, they can have a very large
98 Design of cylindrical and conical corrugated horns

Table 5.1 Types o f corruga ted h orns and their uses


Types of horn Typical uses
Large aperture, narrow flare angle Large Cassegrain antennas
Earth stations
( ~ > 4 Q < 15° 1 Radio astronomy antennas
X /
Small aperture, narrow flare angle Spacecraft antennas
Prime focus
~< 4,6, < 15°
A

Wide flare angle Prime focus reflectors


(0i > 15°) Gain standard
Low frequency
Multimode horns Radars
Prime focus reflectors
Offset reflectors
Very broad band horns Electronic warfare
Multifrequency Multifrequency
communication
channels
Compact horns Spacecraft antennas

bandwidth with nearly constant copolar radiation characteristics. We have found


that they can also have very low crosspolarisation characteristics. However the
beamwidth of the patterns produced by the wide flare angle horns limits the range
of applications. The remaining types of horn listed in Table 5.1 are specialist types
which are derivatives from the first three types. The characteristic of the multimode
tracking horns, as the name implies, is the ability to propagate and radiate higher
order modes in addition to the fundamental mode. Very broad band horns are used
for situations where a number of frequency channels are required to be received, or
radiated by the one horn. The crosspolar characteristics and the match at the throat
of the horn are basically narrow band, so broad band horns can only be obtained
by some compromise of the characteristics. A related type of horn is the multi-
frequency horn where the aim is to pass two or more separate signal channels with
good characteristics at each channel frequency. Compact horns are horns where the
taper from throat to aperture is nonlinear, producing a shorter horn than with a
linear taper but retaining the desirable radiation characteristics of a narrow flare
angle horn.
The electrical parameters which determine the performance of the corrugated
Design of cylindrical and conical corrugated horns 99

Table 5.2 Design parameters for corrugated horns


Radiation patterns
Principal plane patterns
Copolar patterns
Pattern symmetry (E and H plane equality)
Beamwidths
Crosspolar characteristics
Frequency characteristics
Frequency of operation
Bandwidth
Multiband operation
Efficiency as a feed for a reflector
Total aperture illumination
Spillover
Phase
Crosspolar
G/T for a reflector
Gain
Input impedance and VSWR
Phase centre and movement
Differential phase shift along horn
Multimode operation
Mechanical
External diameter
Length
Weight
Ease of manufacture

horn are listed in Table 5.2 not all the parameters are of interest in one application.
Often one or two parameters are of overriding concern and will determine the way
the horn is designed and manufactured. These parameters will be referred to
throughout the remainder of this chapter. Most are self-explanatory or will be
defined when they appear in the text.
In any type of corrugated horn we can identify four parts of the horn for design
purposes, since to some extent each part can be considered separately. The four
parts are shown in Fig. 5.1. They are (i) the aperture diameter and flare angle which
principally determine the copolar beamwidth;(ii)the corrugations, which determine
the pattern symmetry and the crosspolar characteristics; (iii) the flare section
between the throat and the aperture which determines the position of the phase
centre and the generation of any higher order modes along the horn; (iv) the throat
region which determines both the impedance match into the section of waveguide
behind the horn and the mode conversion level at the throat. Each of these parts
will be investigated in detail in the next four sections. Where theoretical results are
100 Design of cylindrical and conical corrugated horns

presented they have been computed using as exact a model as is necessary for the
problem. Copolar patterns can be adequately predicted with a simple representation
of the horn aperture fields, but crosspolar patterns need a more exact model which
includes the corrugation geometry if accurate design data is desired. The theoretical
methods have been studied in detail in the preceeding chapters. Therefore, in this
chapter the background theory will be assumed and only the results and
explanation presented.

3 flare section
4 throat 1 aperture

2 corrugations

Fig. 5.1 Corrugated horn showing regions relevant for design

5.2 Copolar Radiation Characteristics

The copolar radiation characteristics of a conical corrugated feed are determined


mainly by the aperture diameter and the horn flare angle. We shall describe here

H plane
(yz plane)

E plane
(xz plane)

F ig. 5.2 Planes of polarisation

how they, and other factors, influence the shape of the radiation patterns. The term
copolar will be taken to include the radiation pattern of the horn in any plane, angle
$ with respect to the reference axis, Fig. 5.2, where the field is parallel to the field
of the source. The crosspolar field is the orthogonal component. The designer is
usually most interested in the principal planes, i.e. the E{<p = 0°) and H(<p = 90°)
planes and the <f> = 45° plane where the crosspolarisation is maximum. Because of
Design of cylindrical and conical corrugated horns 101

the importance of being clear about the definition of polarisation it is worth


restating the definition universally accepted for horn and feed studies. This is the
so-called Ludwig Third definition [185] which states that the reference polarisation
is that of a Huygens source (electric and magnetic dipoles with orthogonal axes
lying in the aperture plane and radiating equal fields in phase along the z-axis).
The crosspolarisation is then the polarisation of a similar source rotated 90° in the
aperture plane. This definition corresponds most closely to what is normally
measured on the antenna range and interchanging the copolar and crosspolar fields
as measured in any direction, corresponds to a 90° rotation of the reference source.
Expressed mathematically, the copolar £ p ( 0 , 0 ) and crosspolar Eq(d, 0) fields
are given in terms of the spherical co-ordinate components Ee (0, 0) and i^>(0, 0)
by
- sin0j ^(,0)1
sin0 cos0J U ) ]
where it is assumed that the principal electric field vector is aligned with the x-axis,
Fig. 5.2. If the electric field vector is aligned with the j-axis, eqn. 5.1 still applies
but the copolar compoent is given by Eq and the crosspolar component by Ep.
The transformation from the aperture fields of a conical horn to the far-field
gives £ 0 ( 0 , 0 ) = Q)(0) cosm0 and £ 0 (0, 0 ) = C 0 (0) sinm0, where CQ (0) and
C<p(d) are the pattern functions. Combining these relations with eqn. 5.1 for a unity
azimuthally dependent mode, gives
F p (0, 0) = Ce (0) cos 2 0 + C 0 (0) sin 2 0
H sin20
The crosspolar field is a direct measure of the difference between the E plane
(0 = 0°) and H plane (0 = 90°) copolar patterns and is a maximum in the 0 = 45°
and 135° plane.
Radiation patterns are usually plotted as Ep and Eq (dB amplitude) against the
0 angular co-ordinate for the 0 = 0, 45 and 90° planes.
The availability of fast computers with graphical output has increased the use of
contour plots, see Fig. 5.3, which displays the complete copolar or crosspolar
pattern. However, for ease of explanation the rectangular decibel plot is preferred
and will be largely used in the following sections.
We shall assume initially that the HE n mode alone is propagating in the horn,
the patterns and influence of higher order mode distortion will be dealt with later
in the section. As is well known, the copolar radiation pattern of a corrugated horn
or waveguide operating under balanced hybrid conditions is basically Gaussian in
shape, at least down to about the — 15 dB power level. This fact is shown in Fig.
5.4, which compares the pattern of an actual corrugated waveguide to an ideal
Gaussian pattern. This and subsequent theoretical computed patterns for the
corrugated horn have been obtained using the exact space-harmonic model unless
stated otherwise. Also, if only one copolar pattern is plotted this implies that the
102 Design of cylindrical and conical corrugated horns

horn radiates a low level of crosspolarisation and there is negligible difference


between the principal plane patterns.
In considering copolar patterns it should be remembered that we are principally
interested in the use of corrugated horns as feeds for reflectors. This means that the
copolar pattern is illuminating the reflector out to the edge angle of the reflector.

^—r—i—i—r
Fig. 5.3 Contour plot of copolar radiation pattern

This angle normally corresponds to a pattern level of about — 10 dB to give the


optimum efficiency (see Section 5.7) but in some low spillover designs an edge
taper of - 20 dB may be required. The patterns at angles outside this level are
rarely of interest, except in so far as the energy at these angles reduces the overall
efficiency. As a consequence, we can give portions of the pattern outside the main
beam only minor consideration, and note the justification for treating the corru-
gated horn as a Gaussian pattern radiator. This fact is particularly useful as it opens
up the possibility of treating the radiation from the horn as a Gaussian beam.
Multiple reflector feed systems have been designed in this way, for instance the
beam waveguide feed used in many INTELSAT Type A earth stations and the
Nasmyth and Coude foci used in radio-astronomy.
Design of cylindrical and conical corrugated horns 103

The Gaussian beam method assumes that the feed radiates a pattern which has
no sidelobes so it is not adequate for a precise design, hence from now on we shall
look at the radiation pattern in more detail. The general copolar pattern of a corru-
gated waveguide is shown in Fig. 5.5, indicating the E and H plane patterns. The
difference between the two planes provides, of course, a measure of the amount of
crosspolarisation present and this will be studied in the next section. Suffice it to
say at this point that the crosspolar characteristics are determined mainly by the
corrugation geometry, but this geometry has only a minor influence on the shape of
the copolar patterns and so can be ignored for basic design purposes. The horizontal
axes of Fig. 5.5 is plotted in normalised units (Dsind)/\, and so applies in general
to any aperture diameter producing that level of crosspolarisation.

D/A sin

Fig. 5.4 Comparison of Gaussian pattern and pattern computed from corrugated waveguide
theory

The radiation characteristics of an open-ended corrugated waveguide are similar


in shape to those of a circular aperture with a tapered aperture distribution.
The copolar pattern has nulls which theoretically go down to zero power level. The
addition of a spherical phase cap to the circular aperture fills in the nulls. This
occurs when the open-ended corrugated waveguide is transformed to a conical horn.
Typical results are shown in Fig. 5.6. For even modest flare angles the pattern
104 Design of cylindrical and conical corrugated horns

-20-

~40

r
-60

Fig. 5.5 Radiation patterns of corrugated waveguide

-10

-30

-40
12 18 24
degrees
Fig. 5.6 Influence of flare angle on copolar radiation pattern. rx =90mm, = 7'5mm,
11 GHz, Parameter horn semi-flare angle
Design of cylindrical and conical corrugated horns 105

differs substantially from the open-ended corrugated waveguide pattern. However


the differences are again mainly outside the angular region used to illuminate a
reflector. There is, therefore, considerable justification in modelling horns with
semi-flare angles up to about 8° as open-ended waveguides, thus simplifying the
analysis. It was shown by Thomas [149] that the patterns for narrow flare angle
horns can be plotted in a normalised form using a spherical wave error A given by

A
tan
(5.3)

The normalised patterns are shown in Fig. 5.7, taken from Thomas [149]. The
aperture diameter is assumed to be greater than about two wavelengths so that the
flange has little effect and for aperture diameters greater than about 4X, the abscissa
must be truncated at the value corresponding to 0 = 90°.

-10

<g-20

-30

-40
12
krj sin 0

Fig. 5.7 Normalised radiation patterns of small flare horns with A as parameter [149]

Figure 5.7 applies for semi-flare angles up to about 20°. Above this angle the
copolar radiation pattern becomes a function mainly of the flare angle and not of
the aperture diameter. This is illustrated in Fig. 5.8, which shows the patterns for
a 30° semi-flare angle horn of varying normalised slant lengths (assume frequency
constant so balanced hybrid conditions maintained). For slant lengths of R/X > 4
(Z)/X > 4) the patterns have approximately similar beamwidths with two particular
features. At the - 15 dB level the patterns cross at d = 30°, the semi-flare angle of
the horn. This feature can be used as a rough design guide for wide flare angle
horns. Secondly, the pattern around boresight loses its Gaussian-like shape. This is
due to the spherical phase cap which causes the energy radiated from the centre of
the aperture to be out of phase with the energy radiated from near the edge of the
aperture.
The radiation patterns discussed above assume that the frequency is held
constant. When the frequency changes, the normalised lengths change and the
106 Design of cylindrical and conical corrugated horns
o

-10

2-4

-30

-40
10 20 30 40 50
angle, degrees

Fig. 5.8 Copolar radiation patterns of 30° semi-flare angle corrugated horn. Parameter,
normalised slant length Ro/X

-40
0

Fig. 5.9 Influence of frequency on copolar radiation patterns of 12° semi-flare angle horn.
ry — 90 mm, s = 7'5 mm, Parameter: frequency (GHz)
Design of cylindrical and conical corrugated horns 107

normalised slot depth changes. This moves the propagation characteristics off the
balanced hybrid condition so that the pure hybrid mode conditions inside the horn
no longer hold. However, experience shows that the corrugated surface is very
tolerant of frequency changes, as far as the copolar patterns are concerned. Thus
Figs. 5.5, 5.7 and 5.8 can be used to indicate the effect of frequency variations of
about ± 20% for a practical horn. Over a wider frequency range, the copolar pattern
must be computed including the corrugation geometry. Fig. 5.9 shows the patterns
of a 12° semi-flare angle horn with a slot depth resonant at f0 = 11 GHz. At
15 GHz (36% above / 0 ) the spherical phase error has introduced a small dip in the
copolar pattern, while at 19 GHz (72% above/ 0 ) the pattern is again smooth but
the beam width is the same as at 15 GHz. The symmetry of the pattern has however
been lost and the crosspolar level will be high.

Fig. 5.10 — 3 dB half beamwidth against normalised aperture diameter. Parameter: semi-flare
angle

The designer of a feed is often given a copolar beamwidth specification and


needs a horn which will meet the specification. Curves of beamwidth against
normalised aperture size for fixed power levels, are useful to start the design process,
and these are shown in Figs. 5.10, 5.11 and 5.12. These curves assume that the
108 Design of cylindrical and conical corrugated horns

corrugation geometry has been adjusted to the balanced hybrid value at each point
on the curve. Fig. 5.10, showing the - 3 dB half beamwidth illustrates many of the
points discussed earlier in this section. The beamwidth of narrow flare angle horns
and open-ended waveguides decreases monotonically with increasing aperture size,
in contrast to the wider flare angle horns which oscillate about a fixed beamwidth.

l 1 I
60

70°

50

U0-

!30-

CD

§20

10 -

3 4 5 6 7 10
normalised aperture diameter,

Fig. 5.11 — 10 dB half beamwidth against normalised aperture diameter

The oscillations are caused by the spherical phase factor. The same general
behaviour is evident from Figs. 5.11 and 5.12, for the — 10 dB and - 20 dB half
beam widths, except that the oscillating behaviour is smoother. The wider angle
horns have a peak gain when the beamwidth is a minimum. This occurs when the
Design of cylindrical and conical corrugated horns 109

3 4 5 6 7 8
normalised aperature diameter, D/A

Fig. 5.12 — 20 dB half beamwidth against normalised aperture diameter

aperture is relatively small and the penalty for nearly constant beamwidth at larger
aperture diameters is a poor aperture efficiency.
The copolar radiation patterns considered above have been the far-field patterns,
computed using a space-harmonic model plus Fourier Transform pattern prediction.
Sometimes it is necessary to know the radiation pattern at a distance nearer than
the far-field, for instance if a subreflector is in the near-field of the corrugated feed.
The near-field patterns can be computed using the spherical wave expansion
technique (Chapter 4). However, with this technique the electric field model for the
corrugations assumes a balanced hybrid mode and therefore cannot take account of
the actual slot geometry. Near-field patterns for three horns with semi-flare angles
8°, 15°, 30° and 5*4 A aperture diameter are shown in Figs. 5.13, 5.14 and 5.15,
respectively. In the case of the 8° and 15° horns, the patterns are referred to the
aperture, while with the 30° horn they are referred to the apex. The pattern shape
is similar at all distances but the angle at which a given power level occurs changes
substantially, particularly for the - 20 dB power level. In terms of normalised
quantities the classic far-field distance (R = 2D2 /A) becomes kR = 4n(D/X)2. For
the horn computed in Figs. 5.13-5.15, this gives kR = 366. It can be seen from the
figures that even at this distance the radiated field is noticeably different from
the kR = °° pattern. Note that the practical phase centre of the horn is not at the
710 Design of cylindrical and conical corrugated horns

Fig. 5.13 Near-field copolar radiation patterns for 8° semi-flare angle horn with normalised
aperture diameter D/X -54. Parameter: kR from aperture

-10 -

-20-

-30-

Fig. 5.14 Near-field coplar radiation patterns for 15° semi-flare angle horn with D/X = 5'4.
Parameter: kR from aperture
Design of cylindrical and conical corrugated horns 111

aperture for the 8° and 15° semi-flare angle horns. This will modify the near-field
patterns as a function of angle 6.
The presence of higher order modes may be wanted or unwanted. In the former
case, where they are used for tracking or as part of a 'matched feed' [132], the
modes under consideration are the H 01 , E^ and HE 2 i. All these mode have
copolar patterns with a null on boresight and change phase through the boresight.

32

Fig. 5.15 Near-field copolar radiation patterns for 30° semi-flare angle horn with D/X= 54,
Parameter: kR from apex

Since the tracking capability is usually an 'extra', the precise pattern shape will
not be a fundamental specification for the designer. Typical patterns are shown in
Fig. 5.16. The three patterns have been normalised to the same peak level, and it
can be seen that they have the same shape in the angular region of interest.
When the higher order modes are present but unwanted, the copolar pattern can
be distorted. The two modes of interest are the HE12 and EH12 modes. Both these
can be excited in the throat and flare regions of the horn due to changes in the horn
cross-section with axial distance (see Chapter 4). Typical copolar patterns of the
two modes are shown in Fig. 5.17. The HE12 modes has a maximum on boresight,
good pattern symmetry and a first sidelobe nearly as strong as the main lobe. It
occurs at an angle corresponding to the HE n pattern level typically used to
illuminate the edge of a reflector (see Fig. 5.5). The EH12 mode pattern has a null
on boresight, a peak at about the same angle as the HE12 mode and a difference
between the E and H plane pattern which is sufficient to give rise to a significant
level of cross-polar radiation. The effect of adding an HE12 mode pattern to the
/12 Design of cylindrical and conical corrugated horns

HE U mode pattern is shown in Fig. 5.18. These are E plane patterns, the//plane
patterns are influenced in the same way. The amount of distortion depends not
only on the level of HE12 power but also on the phase angle between the two
modes. The sidelobe level may be raised or lowered. The presence of HE12 mode

-10 -

-20 -

Fig. 5.16 Copolar radiation patterns of - - HE2X and E02 modes, H0l mode

-30

Fig. 5.17 Copolar radiation patterns for, a HEl2 mode and b EHl2 mode
E plane H plane
Design of cylindrical and conical corrugated horns 113

distortion on a measured pattern can be confirmed by measuring over a frequency


band in order to detect the phase changes. The crosspolar patterns are virtually
unaffected by the presence of the HE12 mode power. The EH12 mode, at the same
relative level as the HE12 mode, influences the copolar patterns in a similar way,
except that there is a 180° phase difference between the EH12 E and H plane

-50
degrees
Fig. 5.18 Addition of — 25'4 dB of HE , 2 mode to HElx mode of an open-ended corrugated
waveguide with rx — 81 mm, s = 8mm, at 10 GHz
H E n mode only L0° HE12 mode Z_90° HE12 mode
£180° HE12 mode

patterns. This means that the two sum patterns are not affected in the same sense,
i.e. for zero degree phase angle, the HE n + EH12 E plane corresponds to the
HE U + HE12 (zero degree phase) E plane pattern, but the HE U + EH12 H plane
pattern corresponds to the HE n + HE n (180 degree phase) E plane pattern. In
addition the crosspolar level will be significantly raised. Thus examination of
714 Design of cylindrical and conical corrugated horns

distorted measured patterns can lead to knowledge of whether either HE12 or


EH2 modes are present.

5.3 Crosspolar radiation characteristics

We shall now discuss the design of corrugated horns to produce low crosspolarisation
as defined in the last section. We shall assume that the horn is propagating an H E n
mode and that the peak crosspolar power occurs in the 45° plane. The crosspolar
power can be attributed to four sources: the intrinsic crosspolarisation of the
corrugated structure; the higher order modes generated by the throat region; the
higher order mode conversion along the length of the horn and direct radiation
from the flange. The second and third factors will be considered in later sections,
the fourth factor is only of concern for small horns (aperture diameter less than
two wavelengths), while the first factor is the subject of this section. In general it can
be stated that, for most corrugated horns, it is the intrinsic crosspolarisation which
will dominate the total radiated crosspolarisation. Only when the intrinsic level has
been reduced to a sufficiently low level, will the other factors become significant.
The level of crosspolar power radiated at any frequency is determined by the
size of the aperture and by the geometry of the corrugations, particularly the slot
depth. Since the aperture size is the factor governing the copolar characteristics, the
crosspolar designer normally has no control over the aperture size. He can, however,
adjust the corrugation geometry to optimise the crosspolar radiation over the design
band of frequencies. In order to study the influence of the corrugations on the
radiation characteristics we need a theoretical model which includes the exact
geometry of the corrugations. For this purpose we will use the space harmonic
model (see Appendix) with one pair of space harmonics in addition to the
fundamental HE n mode in the inner region and one pair of slot modes. Experience
has shown that one pair of space harmonics is adequate for design and study
purposes. Including further space harmonics makes only a few percent difference to
the power level, under the worst circumstances.
The space harmonic model allows for an accurate description of the electric and
magnetic field throughout the waveguide, in contrast with the equivalent surface
impedance model. However accuracy is traded for simplicity. Before considering
the exact model let us review the main features of the crosspolar radiation from a
corrugated horn or waveguide, radiating the HE U mode. The crosspolarisation is zero
on boresight and also zero along the principal planes. It rises to a peak along the
45° planes, Fig. 5.19 (note that the crosspolarisation level is always referenced to
the copolar boresight). The first crosspolar lobe peaks at an angle corresponding to
a copolar power level of about — 9 dB. A result which is valid for most narrow flare
angle corrugated horns. For wide flare angle horns, the peak level moves inwards
towards the boresight direction. The significance as far as the user is concerned is
that the peak crosspolar level lies within that portion of the main copolar beam
which illuminates either the main reflector or the sub-reflector. Hence the cross-
Design of cylindrical and conical corrugated horns 115

polarisation is transformed through the reflector system to appear in the far-field


of the reflector antenna. For this reason the peak crosspolar level is a significant
design parameter. The variation of the peak level with normalised frequency (/// 0 )
is shown in Fig. 5.20. An equation for the peak crosspolar level is given in Section
3.5 (eqn. 3.82), but because of the model, it does not take account of the geometry
of the corrugations. However, it does indicate that the bandwidth over which the
peak crosspolar power lies below a given value is proportional to the aperture
diameter of the horn.

i i i r i 1 r i i i i

Fig. 5.19 Contour plot of crosspolar radiation pattern

The value of f0 in Fig. 5.20 is the frequency where the corrugations are
'resonant', and will change as the geometry of the corrugations is changed. It is
the balanced hybrid frequency.
The major limitation of the methods of analysis based on considering the
corrugations as an admittance surface is in the precise determination of/ 0 . The
simple model assumes that the normalised susceptance of the corrugations can be
represented by
Y = cot (ks) (54)
/16 Design of cylindrical and conical corrugated horns

and that the hybrid factor is given by (Section 3.2, Table 3.2),

Y _L_ Y\
A - 1 - (5.5)
2 krl

Further the crosspolar radiated field is proportional to the factor:


A-l
(5.6)
A+ 1
From the above equations it would appear the crosspolarisation will be zero when
the slot depth, s, is A/4 and the hybrid factor is unity. This is not borne out in
practice, except for large diameter horns. A better model is obtained by using the
full surface impedance approach of eqn. 3.26 but even here the longitudinal
geometry of the corrugations are not treated in the model.

Fig. 5.20 Sketch of peak crosspolar radiation against normalised frequency

The above discussion has highlighted the problem of using a simple theoretical
model to predict the crosspolarisation characteristics. It means that the hybrid
factor, A, which is so useful in explaining qualitatively the propagation and
radiation behaviour of corrugated waveguides cannot be used for precise crosspolar
design calculations. Computations with the space-harmonic model show that the
value of A to produce minimum crosspolarisation is always greater than unity, but
there is as yet, no simple analytic relationship to enable the value to be predicted
from the corrugation geometry.
It is the presence of space harmonics in the corrugated horn or waveguide which
modifies the balanced hybrid frequency,/ 0 . This is shown in Fig. 5.21, where the
power in the HE n mode of a relatively small diameter waveguide has been divided
into the components from the individual — 1,0, + 1 space-harmonics. The frequency
for minimum crosspolarisation depends not only on the fundamental harmonic,
N = o (which is the only one considered in the surface impedance model) but on
the higher order harmonics. The power in the space harmonics becomes less
significant as the diameter of the horn is increased. However, they retain their
Design of cylindrical and conical corrugated horns 117

influence on the balanced hybrid frequency long after they have ceased to be
significant for copolar predictions. This is because the fields of the space-harmonics
are confined near to the corrugated surface and this surface controls the filter-like
properties of the crosspolar frequency characteristic.

Fig. 5.21 Influence of space harmonics on peak crosspolar radiation of a corrugated horn
with r,/rn = 08

We shall now study how the corrugation geometry effect the balanced hybrid
frequency. Consideration of the space-harmonic formulation (see Appendix) shows
that the following parameters are significant:
D s b t
\'X' \'~b
that is, the normalised diameter (D/X), normalised slot depth (s/X), normalised slot
width (b/X) and the ratio of ridge width to slot width (t/b). A version of our space-
harmonic program was written to undertake a parametric study of the normalised
quantities for radiation from an open-ended corrugated waveguide. The pertinent
results are plotted in Figs. 5.22, 5.23, 5.24 and 5.25. They show the normalised slot
depth which gives minimum crosspolar power against the normalised aperture
diameter for either fixed b/\ or fixed tjb. The slot widths are varied between (HA
and 0-3A and the ridge width to slot width ratios between 0-1 (narrow ridge) to 1 -0
(equal slot/ridge width). This spans the range between 2 slots per wavelength and 9
slots per wavelength.
/18 Design of cylindrical and conical corrugated horns

It should be emphasised that throughout this range the form of the peak cross-
polarisation is as shown in Fig. 5.20. One of the most common questions asked
about corrugated feed design is 'how many slots per wavelength?'. The answer from
the point of view of the radiation characteristics is that, within a wide range it does
not matter as long as the slot depth is adjusted according to Figs. 5.22—5.25. A
number of features are evident from the figures. The slot depth is always deeper for

Fig. 5.22 Slot depth for minimum crosspolarisation against normalised aperture diameter.
Slot width = 0-1K. Parameter: ridge width to slot width ratio
0*4

Fig. 5.23 Slot depth for minimum crosspolarisation. Slot width = 0m2X. Parameter: ridge
width to slot width ratio. surface impedance result

small aperture diameters than for large aperture diameters. Only for large diameters
do the slot fields follow trigonometric function relations, then the slots can be
considered to belong to a plane surface. For aperture diameters under 2X the slot
Design of cylindrical and conical corrugated horns 119

depth to produce balanced-hybrid conditions is much deeper than the nominal


quarter wavelength deep slot.
The second observation which can be deduced from the figures is the influence of
the longitudinal geometry of the corrugations in determining the balanced-hybrid
frequency, even for large aperture sizes. Also it is clear that the ratio of the ridge
width to the slot width is more important than the normalised slot width. Small
values of tjb require deeper slots than when the slot and ridge have equal width.

Fig. 5.24 Slot depth for minimum crosspolarisation. Slot width = 0'3k. Parameter ridge
width to slot width ratio.

Fig. 5.25 Slot depth for minimum crosspolarisation. Ridge width to slot width ratio = 0'5.
Parameter: slot width, K

Slots which are relatively wide, e.g. 0-3A, Fig. 5.24, show more difference
between the minimum and maximum ridge/slot width ratios. The maximum value
which could be computed was t\b = 0-67. For larger values numerical instability in
120 Design of cylindrical and conical corrugated horns

the computations occurred. Note that for this value the normalised slot depth is less
than A/4. The fields around the corrugations are now so different from the values
assumed in a surface impedance model that concepts such as inductive and
capacitive slots have little meaning. In passing we note that wide slots are important
at millimetric and sub-millimetric frequencies where the practical problems of
making corrugated horns demand that there be few corrugations per wavelength.
Fig. 5.23 also shows the result produced by the surface impedance model of
eqns. 3.16-3.26. This is similar to the tjb = O5 curve. Its significance lies in the
fact that most corrugated feeds which have been designed so far would lie on this
curve (i.e. feeds with D/X > 3, b/X < 0-2 and t/b ^ 0-5). Thus a surface impedance
model, in which the slot fields are represented by Bessel functions, does in fact give
a good estimate of the frequency at which minimum crosspolarisation occurs, as
long as these restrictions are noted. Indeed it is possible to use the even simpler
model of eqn. 5.4 by computing the crosspolarisation assuming X/4 slots, but then
using Figs. 5.22-5.25 to obtain the real slot depth.
Experimental verfication of Figs. 5.22-5.25 has been obtained by comparing
the predicted crosspolar minimum with values measured on a number of feeds with
a wide range of geometries. The comparison is shown in Table 5.3. The generally
good agreement between theory and experiment is adequate confirmation of the
parametric study.
The discussion above specifically referred to narrow flare-angle horns or open-
ended waveguides. However, as Table 5.3 shows, the values apply to wider angle
horns as long as the electric depth of the slots at the aperture is used. This value is
not always easy to predict. For horns with semi-flare angles up to about 15°, the

Table 5.3 Comparison of theoretical and experimental results for the frequency
of minimum crosspolar radiation: Measured and computed results are
for the frequency where minimum peak cross-polar power occurs
Horn D slot ridge w Theory Measurement
X width slot w (K = L=l) (slot depth)/X
X (slot depth)/X
Open-ended
waveguide 1-64 0-33 0-16 0-36 0-35
Profiled horn
(9.7 GHz) 2-88 0-1 0-33 0-29 0-28
12° linear horn
(8.7 GHz) 3-68 0-23 049 0-29 0-28
20° linear horn
(11° GHz) 6-88 0-18 0-3 0-27 0-26
30° linear horn
(9.8 GHz) 5-0 0-26 0-31 0-28 0-26
5-3° linear horn
(14.1 GHz) 1-36 0-12 0-21 0-39 0-39
Design of cylindrical and conical corrugated horns 121

electrical depth can be taken as the depth in the centre of the slot, see Fig. 5.26; for
larger flare angles a mean depth is best chosen and must partly be done by trial and
error. If the slots are cut perpendicular to the axes, the central slot depth can be

Fig. 5.26 Orientations of slots

-20 r

-30-

1-40-

-50

0-6 0-8 1-0 1-2 1-4


/
Fig. 5.27 Peak crosspolar radiated power as a function of f/f0 where f0 is the frequency for
minimum crosspolarisation. Parameter: normalised aperture diameter
122 Design of cylindrical and conical corrugated horns

used, if they are cut perpendicular to the horn flare angle a value about 20% less
than the slot depth gives a reasonable estimate. The value is easier to estimate if the
slot width is narrow. However, experience shows that corrugated horns work just as
well whichever style of slot is chosen, so the choice can be made on mechanical
grounds rather than electrical grounds.

80r

60 below -30dB

c
2
below -40dB

20

0 2 4 6
diameter, wavelength
Fig. 5.28 Theoretical percentage bandwidth for peak crosspolar level below — 30 dB and
-40dB

When the balanced hybrid frequency, f0, has been determined or chosen for a
particular feed, the crosspolar bandwidth can be computed. Theoretical curves
showing the peak crosspolar power for three normalised aperture diameters are
shown in Fig. 5.27. From these curves, Fig. 5.28 has been obtained, showing the
theoretical percentage bandwidth below which the peak crosspolar level is less than
— 30 dB or less than — 40 dB. We should stress that both curves are the theoretical
levels assuming all the crosspolarisation is generated by the intrinsic geometry of
the waveguide. If crosspolarisation is present due to higher order modes or due to
the flange, it may well dominate the intrinsic crosspolarisation. The crosspolar
bandwidth will sometimes be the factor which determines the usable bandwidth of
the feed because it can be much less than the copolar bandwidth. Attempts have
been made to increase the crosspolar bandwidth of conventional corrugated horns
but without success. However the dual-depth horn exhibits two bands of frequency
in which the crosspolarisation is low, see Section 5.8.7.
Before finishing this section, a brief reference will be made to the near-field
crosspolarisation patterns. It was shown in Section 5.2 that the copolar patterns do
Design of cylindrical and conical corrugated horns 123

not change their shape until the observation point is very near to the horn. The
crosspolar results show a similar behaviour in that the peak crosspolar level remains
more or less constant from the aperture to the far-field. Sample results for a 12°

w\
A
\ V
- 1 0 --
\
\
\
\
\ copolar
\
1
\

•o
I
\
\
\
\ \ \
\
-20-
r'\ \

si
N. crosspolar

\\
l *
-30 « \
16
degrees

Fig. 5.29 Near-field radiation patterns for 12 degree semi-flare angle horn with kR0 =81.
kR = oo _ . _ kR = 280 from apex. kR = 81 from apex

semi-flare angle horn are shown in Fig. 5.29. The angle at which the peak cross-
polar level occurs moves inwards as the distance from the horn to the far-field is
reduced. Thus the horns can be operated well into their near-field region without
significantly altering the crosspolar radiation patterns.

5.4 Flare section of horn

We have discussed in the previous sections the design of the aperture of the horn to
achieve a specified copolar radiation pattern and the design of the corrugation
geometry at the aperture to achieve a specified crosspolar characteristic. This
section will deal with the design of the flared section between the throat and the
aperture. In some cases the copolar pattern specifications will determine the flare
angle, see Section 5.2, but there are three other considerations:
(i) the mechanical constraints on the length;
(ii) the position and movement of the phase-centre;
(iii) the mode conversion which takes place along the flared section.
124 Design of cylindrical and conical corrugated horns

For a given aperture diameter and a given throat design the safe solution is to have a
long horn with a narrow flare angle. This approach has been adopted in many large
dual-reflector earth-station antennas. There are however many cases where the long
length and consequent weight are not practical, and in these circumstances it is
important to know the electrical consquences of changes from the ideal.

Phase centre
The phase centre of any horn is the point on the axis of the horn which is the
centre of curvature of the phase front. It is usually found by measurement or by
deduction from the computed radiation pattern. In either case the assumption is
made that the fields can be considered to be spherical waves and hence, given the
amplitude and phase patterns, the centre of the sphere (the phase centre) can be
calculated. It is possible to find the practical phase centre by placing the test horn
on a turntable and recording the phase pattern for various rotation points along the
axis of the horn. When the phase patterns are constant, with changing azimuth
angle, the test horn is rotating about its apparent phase centre.

Fig. 5.30 Geometry for calculation of phase centre

The apparent, theoretical, phase centre can be computed from the amplitude
and phase of the radiated field at any distance from the horn by assuming that the
radiation at two adjacent points, in the same plane, emanates from a common point
on the axis of the horn. Let the two points, Px and P 2 , be at angles dx and 62 and
distance R from the apex of the horn, Fig. 5.30, and let the apparent phase centre
be at 0, distance L from the apex. If the difference between the radiation phase
angles at ?t and P2 is 81// then, by geometry, the distance L is given by

/ 2
2 \ \
X
1)± cos0 2 COS0J--1 + — + (cos0 2 -

(5.7)

where X =
Design of cylindrical and conical corrugated horns 125
The positive sign is for 5\[/ positive and the negative sign for 5i// negative. In the far-
field, R -> °° and the above equation simplifies to:
r £,/,
(5.8)
\ COS02 ~COS0!
These equations enable the apparent phase centre to be computed at each point in
the radiated field.
This assumes that the phase centre is well defined, i.e. that the fields at any
point in space are spherical waves. Unfortunately this is not true. There is no single
phase centre in most horns. The phase centre varies with the plane of polarisation,
the angle of boresight, the frequency and the distance from the horn. There is one
important exception; if the horn is an open-ended waveguide the phase centre
remains essentially fixed at the aperture of the waveguide. This fact has partly given
rise to the profiled corrugated horn as a means of achieving the ideal situation.
To return to the phase centre in linear flare angle horns, it is difficult to lay
down quantitative guidelines for the behaviour of the phase centre, so what follows
is a qualitative discussion of the affects. Each horn needs to be treated individually
and the phase centre computed for the desired radiation conditions. The phase
centre in the E and H planes, will, in general, be different. However, for low cross-
polar designs the E and H plane patterns are nearly coincident and hence the phase
centres can be assumed to be coincident.

Fig. 5.31 Normalised distance of horn phase centre from apex against A/X [149]

The horn flare angle and the frequency of operation also determine the position
of the phase centre. For wide flare angle horns the phase centre is near the apex, so
for horns with intermediate angles between a waveguide and a wide flare angle horn
the phase centre is in between the apex and the aperture. The far-field position can
be conveniently expressed in terms of the sperical phase factor A, eqn. 5.3. Fig.
5.31 shows the position for narrow flare angle horns. This gives the position for an
observer in the far-field and along the boresight of the horn. The phase centre varies
with angle of boresight because the radiated field is a summation of spherical
126 Design of cylindrical and conical corrugated horns

modes. The variation with azimuth angle is plotted in Fig. 5.32 for horns with semi-
flare angles of 8° and 15° and 5-4 A aperture diameter.
The distance of the observation point from the horn also determines the phase
centre. The position in the near-field is not the same as the far-field position. At the
aperture the horn fields are nearly spherical so the phase centre at the aperture is
always at the apex of the horn. As the distance from the horn is increased the phase

Fig. 5.32 Relative phase centre movement with angle for horn with normalised aperture
diameter — 5'4X. Parameter: semi-flare angle

100 200 300 400


distance from aperture, kR
Fig. 5.33 Position of near-field phase centre relative to far-field phase centre. Parameters as
Fig. 5.32

centre will move outwards from the apex until it reaches the far-field value given by
Fig. 5.31. The position is shown in Fig. 533 for the 8° and 15° semi-flare angle
horns of Fig. 5.32. Again the 8° semi-flare angle horn reaches the far-field value
sooner than the 15° semi-fire angle horn. This fact is particularly relevant when the
horn is in the near-field of a reflector and means that the horn and reflector need to
be closer than is apparent by considering the far-field patterns.
Design of cylindrical and conical corrugated horns 127

Mode conversion along the horn


The mode conversion along the horn needs to be kept to a very low level. The
theory of mode conversion, as applied to wide flare horns, has been studied in
Chapter 4 using spherical modes. Using cylindrical modes, Dragone [56] showed
that the mode conversion in small flare angle, large diameter horns with constant
surface reactance, can be expressed in the form:

where M is the mode conversion coefficient. This implies that it is necessary to keep
the flare angle small or to maintain a high wall reactance such that tan2 ks -> °°. The
spherical mode model shows that the mode conversion is a function of dd/dz, i.e.
the rate is zero in a constant flare angle horn and so indicates that the lowest
mode conversion is obtained with narrow, linear flare angle, constant depth slots.
The latter criteria conflicts with the need to transfer power from a smooth-wall
waveguide, which means that slots near the throat must have varying slot depths
(see next section). The former criteria tends to suggest that only linear flare angle
horns should be used. However, there are some cases where there are overriding
reasons for not using a linear flare angle horn.

Profiled horn
As has already been discussed, the phase centre of the horn can only be fixed with
changing frequency by radiating from an open-ended waveguide. A profiled horn
enables this to happen. Another very useful reason for choosing profiled horns is
to reduce the physical length of the horn and produce a compact horn. This is
particularly suitable for spacecraft applications. The form of the profile is sketched
in Fig. 5.34, and is a sine-squared type of profile. The concept is similar to the
profiled waveguide tapers developed for the TEOi overmoded waveguide. The
total length of the horn can be reduced by up to one third of the length required
for a linear taper, the limit of reduction being determined by the acceptable mode
conversion level. A short section of constant diameter corrugated waveguide is
added to the aperture to ensure that the phase centre conditions are met. A suitable
form for the profile is given by:

r{z) = > n + ( ' - 1 2 - > n ) | | 0 - ^ ) + ^ s i n 2 ( ^ | j (5.10)

The parameter A determines the amount of profiling. When A = 0 a linear taper


results, when A = 1 a sine squared profile results. A value of A = 0.7 gives a
reasonably gradual change in diameter. These types of horn have been built and
tested and found to work well, see Section 5.8.7.
The mode conversion along the horn due to changing flare angle is analysed
in Chapter 4. It is shown that low mode conversion to the HE12 mode occurs
when the profile has a maximum flare angle nearer to the aperture than half way
128 Design of cylindrical and conical corrugated horns

along the horn. This means that the profile is non-symmetrical and for the first half
of the horn L\, Fig. 5.34, the inner radius increases slowly but in the second half of
the horn, L2, the inner radius increases rapidly. The optimum profile was computed
to be whenZ^/Z/2 = 0-4. The reduced mode conversion is due to phase concellation
between the H E n and HE12 modes in the large diameter horn section. The effect is
relatively broadband because it depends mainly on profile shape.

Fig. 5.34 Geometry of profiled corrugated horn

5.5 Throat region of horn

Corrugated feeds are usually connected to a section of smooth-wall circular wave-


guide as a means of transferring power between the radiating horn and the trans-
mission line where signal processing may be conveniently undertaken. The junction
is one of the most difficult parts of the feed to design correctly for optimum
performance, particularly for wide band operation. Any mismatch between the
modes in the circular waveguide and the modes in the throat section of the horn
will give rise to mode conversion to higher order modes. Once excited, an unwanted
mode may radiate and cause increased cross-polar power and/or reduced efficiency.
The effects can become complicated if operation over an appreciable band is
required.
We have already studied theoretically (Section 3.6 and 4.4.2) the techniques for
analysing the match between a smooth wall waveguide and a corrugated waveguide.
The results from field matching studies support the well-used practice that the first
slot should be approximately half wavelength deep so that a short-circuit occurs at
the edge of the corrugations, that is, at the inner radius of the corrugated wave-
guide. This argument, however, already introduces a frequency dependent factor,
namely the half wavelength deep slot and thus does not answer the question what is
the best design away from the central frequency, nor does it indicate how to alter
the depths of the slots from the half wavelength value to the value required for
radiation purposes. Detailed computer studies, using field matching, have been per-
formed for the throat section, but there are a large number of parameters and
probably no one optimum solution so generalisations are not easy.
In this section we shall present a simple graphical method for designing the
junction between a conical corrugated horn and a smooth-wall circular waveguide.
Design of cylindrical and conical corrugated horns 129
This is based on the relationship between the reflection coefficient and the
propagation coefficients in a waveguide. It is not claimed to be a rigorous technique
but its value lies both in its simplicity and in the insight it gives into the frequency
dependent behaviour of the junction. The technique is based on the relationship
between the propagation coefficients 0i and j32 in two waveguides on either side of
a junction and the reflection coefficient, p, of the junction

\p\ = (5.11)

This relationship has been derived in Section 3.6. It is in practice more convenient
to deal with the corresponding guide wavelengths so

(5.12)
Agl

This means that a perfect match will be obtained when the guide wavelengths in the
two waveguides have the same value and that a knowledge of the guide wavelengths
will give an immediate estimate of the junction reflection coefficient.
1-8

1-6-
• \ \

\ \
smooth
y wall

A
1-4- corrugated
\
\ \
\ \
\ \

1-2 -

0
1-0
0-3 0-4 0-5 0-6

Fig. 5.35 Normalised guide wavelength of smooth-wall and corrugated waveguides

We shall concentrate on the problem of matching a T E n mode in a circular


waveguide to an HE11L mode in a corrugated circular waveguide. The nature of these
two modes is such that their guide wavelengths have different dependences on
frequency. This is shown in Fig. 5.35 where the guide wavelength is plotted against
normalised inner radius (fi/X) for a circular waveguide and a corrugated waveguide
130 Design of cylindrical and conical corrugated horns

with slot depth chosen to match at rxj\ = 0-5. The horizontal scale should be inter-
preted as changing frequency, not changing radius. The variation with frequency is
clearly different for the two waveguides, so a good match can be achieved at only
one frequency. The design aim is to ensure that the matched frequency is chosen to
give acceptable impedance values over the operating band. If low VSWR is an
important specification then Fig. 5.35, in conjunction with eqn. 5.12, tells us that
the frequency for perfect match should be nearer the lower end of the frequency
band because the rate of change of guide wavelength is greater at lower frequencies.
This will mean that the slots will be greater than half a wavelength deep at the
upper frequency, an undesirable situation but one which cannot be avoided in a
wide band, low VSWR corrugated horn.

0-3 0-4
slot depth/A

Fig. 5.36 Normalised guide wavelength against normalised slot depth for corrugated wave-
guide. Parameter rx /K

The above discussion referred to the junction between two waveguides, whereas
in reality we have the junction between one waveguide and a conical corrugated
section. This situation can be simulated by approximating the horn as a series of
constant diameter corrugated waveguides. The guide wavelength in each corrugated
waveguide section can be computed and plotted on a composite graph to give an
Design of cylindrical and conical corrugated horns 131

indication of the way in which the field changes as it propagates along the horn.
The guide wavelength will be high at the junction and decrease toward the free-
space wavelength at the horn aperture. Sharp changes in guide wavelength along the
throat section of the horn are indicative of a change in impedance that can cause a
higher order mode to be excited. Thus not only should the guide wavelength on
either side of the circular to corrugated waveguide junction be equal, but the
change in guide wavelength with distance along the horn, i.e. dAg/dz should be as
low as possible.
The information needed to design or analyse the throat section of the corrugated
horn can be conveniently displayed on a single graph. Curves showing the
normalised guide wavelength against normalised slot depth for various normalised
inner radii are shown in Fig. 5.36. These have been computed using the space
harmonic representation for the fields with a slot width of 0*1 X and a ridge width
of 0-Q5A. However, in practice the longitudinal slot geometry has only a small
influence on the propagation characteristics so the curves can be used for most
longitudinal slot geometries.

1-5 -

1-4

0-3 0-4 0-5


slot depth/X
F ig. 5.37 Trajectories o f typical junction
HE12 cut-off

The use of the curves is indicated in Fig. 5.37, curve A. This shows a trajectory
for a corrugated horn obtained by plotting the inner radius and slot depth for each
corrugation along the horn. The guide wavelength for the smooth-wall waveguide
occurs on the slot depth = 0-5X line, point A. The first slot (from the junction) in
the corrugated section gives point B. The second slot point C, and so on. When
the tapering of the slots has stopped and the slot depth is constant, point D is
reached and from then on the guide wavelength will decrease along the line E.
A good match at the design frequency would have points A and B coincident
132 Design of cylindrical and conical corrugated horns

and a smooth change along the horn. In the example shown in Fig. 5.37, there
would be a mismatch at the junction giving a return loss of about — 28 dB. The
mode conversion along the horn depends on whether a higher order mode is
propagating at the relevant radius. The most troublesome mode is the HE12 mode
because this is excited by changes in waveguide cross-section. The cut-off line for
the HE12 mode is shown in Fig. 5.37, and in this case the uneven curve occurs at a
point in the horn where the HE12 mode cannot propagate.
To use these curves to design a throat section, an ideal smooth trajectory would
be drawn on Fig. 5.36, between the chosen smooth-wall waveguide value and the
chosen final slot-depth value. Points would then be chosen along the curve at which
the slots could be placed. Care must be taken to ensure that the inner radius
changes smoothly with distance along the horn. A typical design is shown in Fig.
5.37, curve B, with values in Table 5.4. This design has ten slots between the throat
(s = 0-5X) and the final, constant, slot depth (s = 0-27X) for a horn with a constant
flare angle. Ten is a safe number of slots although horns have worked well with as
few as four slots. The figure of ten slots is supported by exact computations per-
formed by James [77]. Note that the slot depths shown in Table 5.4 do not decrease
linearly with distance. The curve for a linear decrease would be more like curve A in
Fig. 5.37, which gives a worse VSWR at the junction. The best match is obtained
with a non-linear profile to the horn. The profile can be partly designed with a low
VSWR as a design goal using the above procedure. A number of corrugated horns
have been designed using the curves and have performed satisfactorily in practice.

Table 5.4 Slot depths in throat region


Slot Inner radius Slot depth
number (wavelengths) (wavelengths)
1 0-45 0-500
2 0-51 0-410
3 0-57 0-360
4 0-63 0-330
5 0-69 0-310
6 0-76 0-290
7 0-82 0-283
8 0-88 0-278
9 0-94 0-273
10 1-0 0-270

A few further points can be deduced from the curves. A better match will be
obtained if the smooth-wall waveguide has a normalised radius greater than 0-5X.
This is not always possible, but too often a good match is sacrificed because the
corrugated horn designer has specified a circular waveguide which is too small in
diameter. The better match is obtained because the guide wavelength changes more
Design of cylindrical and conical corrugated horns 133

rapidly for small corrugated or smooth-wall waveguide diameter. A linear taper on


the horn does not necessarily produce a good throat section. Particularly in the
region of the junction some non-linear taper may give a better match.
The design problem becomes more complicated when a band of frequencies is
desired. Then the trajectories at the upper and lower operating frequencies must be
placed on the guide wavelength curves. Some compromise will be needed, as a
perfect match is possible at only one frequency. The situation is illustrated in
Fig. 5.38. Clearly a bad mismatch will occur at the lower frequency, and this can
only be rectified by making all the slots slightly deeper so that the design
frequency, / 0 , is shifted towards the lower end of the operating band. This will
improve the lower frequency mismatch, but make the higher frequency mismatch
slightly worse. However, since the guide wavelength changes more rapidly at the
lower frequencies, the overall effect will be an improved VSWR.

0-3 0-4 0-5


slot depth A
Fig. 5.38 Trajectories of three frequencies at a junction

As already mentioned, the longitudinal geometry of the slots has only a marginal
influence on mismatch behaviour. It has more influence on the mode conversion
along the horn, and in practice we have found it to be better to have a large number
of slots per wavelength in the vicinity of the throat. The exact number of slots per
wavelength will depend on the application and operating frequency. A value which
is ideal, say ten, will be impractical at millimetre wavelengths. James [74, 75, 77]
has studied the junction between a T E n mode waveguide and an HE U mode
corrugated waveguide (see Section 3.6). His conclusions confirm our discussion.
He has designed in detail a throat region and has extended the work to study
ring-loaded slots, Fig. 5.39, as a means of increasing the bandwidth of the throat
region. He found, theoretically and experimentally, that the return loss could be
made to exceed — 30 dB over a 145 bandwidth ratio with conventional slots. By
134 Design of cylindrical and conical corrugated horns

Fig. 5.39 Ring-loaded slots

2-6

0-1 0-2 0-3


slot depth/A

Fig. 5.40 Normalised guide wavelength against normalised slot depth for HE21 mode in
corrugated waveguide. Parameter rx A

0-1 0-2 0-3


slot depth/A
Fig. 5.41 Normalised guide wavelength against normalised slot depth for Bo2 mode in corru-
gated waveguide. Parameter rl A
Design of cylindrical and conical corrugated horns 135

using ring loaded slots [75] a bandwidth ratio of 1-55 was possible, but this could
be increased to at least 2-0 if some HE12 higher order mode was accepted. These
results are impressive, unfortunately the increased manufacturing complexity and
the limitations of the waveguide components probably make it impractical except
for a few cases.
There is one further complication to deal with before we complete this section.
Sometimes it is required to pass tracking modes (HE 2 i, E02) through the junction as
well as the signal mode (HE U ). This necessarily implies that the smooth-wall
circular waveguide diameter will be larger enough to accommodate the tracking
modes. This will help reduce the mismatch on the HE U mode but a perfect match
is not possible for all modes. Guide wavelength design curves similar to Fig. 5.36 are
shown in Figs. 5.40 and 5.41, for the HE21 and E^ modes, respectively. The design
procedure can be repeated for these modes and a compromise arrived at between
the matching of the signal mode and the matching of the tracking modes.

5.6 Gain and directivity

The power gain of any antenna in a specified direction (#, 0) is


r,,n x 4TT power radiated per unit angle in direction 6,<p .
G(p,(p) = • (5.13)
total power accepted from source
This quantity includes the ohmic or dissipative losses arising from the conductivity
of the metal. It is therefore difficult to evalute theoretically and the quantity
normally calculated is the directivity. This has a definition similar to gain but the
denominator is replaced by 'total power radiated by the antenna'. The peak
directivity is normally evaluated, which, for a horn, implies the directivity along the
boresight. The peak directivity can be expressed in terms of the tangential aperture
electric fields,/^, of the horn as:

where the area A and the fields are expressed in either cylindrical or spherical co-
ordinates for open-ended waveguides or conical horns, respectively.
The tangential field expressions inserted into the above equation can be one of
the simplified forms discussed in Chapter 3. It is unnecessary to use a precise model
unless the horn is a long way from the balanced hybrid condition since the cross-
polar components of the horn aperture field do not contribute to the directivity.
The peak directivity has been computed for conical corrugated horns using a
spherical wave expansion method, Chapter 4. The results for a wide range of
aperture diameters and horn semi-flare angles are shown in Fig. 5.42. Also shown
is the directivity for a uniformly illuminated circular aperture, which indicates
that the corrugated horn is a relatively inefficient radiator, especially for wide
136 Design of cylindrical and conical corrugated horns

angle horns. As the aperture diameter of the horn is increased, the directivity
reaches a maximum value and then oscillates about a lower value. The maximum
directivity occurs when
157
(5.15)
do
and has a value

A>(0,0) *. 201og10 — (5.16)

In both these equations 0O is in degrees and they are accurate to within 2% for
5° < 0o < 70°. The formulae are mainly of use for calculating the gain of wide flare
angle horns, since the peak directivity of narrow flare angle horns occur at very
large aperture diameters.
30 " I I I i - 1 \^>

circular
aperture^N^

25

20°

CD
-

T3

S 15
//y-—--. 30°

40°_
a

70°
10 ---—«_-——

0 ! 1 | I I 1 i i
" 1 2 3 4 5 6 7 8 9 10

Fig. 5.42 Directivity against normalised aperture diameter

The corrugated horn can be used as a standard gain horn [22]. In some respects
it is better than the more usual pyramidal horn because the pyramidal horn has high
sidelobes and its E plane edge diffraction causes uncertainity in the computed gain.
Design of cylindrical and conical corrugated horns 137

Very good agreement between the calculated and measured gains of corrugated
horns have been recorded. For instance, Chu and Legg [24] obtained agreement of
O2 dB for a 100 GHz corrugated horn, a difference between theory and experiment,
which they attributed to the ohmic losses along the horn.

5.7 Efficiency when used as a feed for a reflector

Most corrugated horns are used as feed for reflectors. The user and designer is
interested in the efficiency of the feed as an illuminator of the reflector, that is, a
measure of how closely the feed approaches the ideal situation of uniform
illumination across the reflector and zero illumination outside the solid angular
region subtended by the reflector at the feed focus. In this section we shall consider
those components of the efficiency which directly attributable to the feed horn.
For most reflector antennas, the feed efficiency components are the main factors
deteriorating the overall efficiency, but other components (reflector surface
tolerance, aperture blocking, sub-reflector scattering and alignment errors) must not
be forgotten.
The gain of a reflector antenna is given by

where D is the diameter of the reflector and r\ is the total efficiency. This is the
product of a number of separate components. The feed efficiency components are
identified in Table 2.1, for a paraboloidal reflector. In that table the feed is
assumed to possess a radiation function of the form

E= (F E sin0 + F H cos0) (5.18)


P
The illumination efficiency, r?l5 is a measure of the effectiveness of the feed to
accept (or match in reception) the uniformly distributed incident energy in the
reflector aperture. It is a function of the amplitude of the feed pattern, and
accounts for the different path lengths from the feed to the reflector surface at
different angles B, the space loss. It is a function of the angle $ 0 , subtended by the
edge of the reflector. This dependence on $0 implies that the efficiency of a narrow
subtended angle reflector (high f/D) will be greater than a wide subtended angle
reflector. Hence, a large aperture horn will give a higher illumination efficiency than
a small aperture horn. An ideal feed with rji = 100% would need to have a feed
pattern such that
(5.19)

to yield uniform amplitude in the aperture after optical reflection from the
138 Design of cylindrical and conical corrugated horns

paraboloid. This pattern increases in amplitude as $ is increased, the opposite


behaviour of a single hybrid mode feed pattern. Hence, the illumination efficiency
of a single mode horn can never approach 100%. There are two ways of improving
the illumination efficiency, either by using a feed which propagates one or more
higher order modes, or by shaping the reflectors in dual reflectors antennas. The
former approach which is discussed in Section 5.8.4 can give a pattern which more
nearly approaches the ideal. It has been used in a number of radio-astronomy
reflector antennas. The latter approach means that the main reflector will no longer
be paraboloidal, and in these circumstances the formulae of Table 2.1 no longer ap-
ply. In general, the multimode horn approach is preferred for small reflectors and
the latter approach for large, Cassegrain type reflectors.
The spillover efficiency, r}s, is a measure of how much energy spills past the edge
of the reflector (or sub-reflector). The wasted energy not only lowers the spillover
efficiency but also contributes to raising the antenna noise temperature because the
angular regions outside 60 'see' background noise. The spillover is the energy
outside angle 0o > s o spillover = (1 — r?s). The feed designer must choose a value of
the feed edge taper which gives acceptable spillover. The actual value will depend
on the application. In very low noise systems it may be desirable to have a very low
spillover with the disadvantage of a reduced illumination efficiency since the outer
portions of an unshaped reflector will be underilluminated. In many small earth
station antennas, the receiver noise will dominate the total system noise and then
the spillover efficiency will be a less important design factor.
The crosspolarisation efficiency, r]x, is a measure of the proportion of the energy
which is contained in the the crosspolar radiation pattern. The crosspolar lobes
receive unwanted energy in the same way as the sidelobes contribute to the spillover
efficiency. In practice the crosspolarisation efficiency can often be taken as 100%
because corrugated horns with peak crosspolar levels below — 30 dB contribute to a
negligible reduction in efficiency.
The last feed efficiency component, the phase efficiency, T?P, is a measure of the
phase error across the reflector caused by non-uniform phase of the feed radiation
pattern in the same way as the illumination efficiency is a measure of the non-
uniform amplitude pattern. An open-ended waveguide has a theoretical uniform
phase pattern across the main beam and so have 100% phase efficiency. A horn
with a finite flare angle has a non-uniform phase pattern across the main beam so
the phase efficiency is less than 100%.
The efficiency formulae given in Table 2.1 have been used to compute the
efficiencies of some corrugated horns when used as feeds for a parabolic reflector
with edge angle 60. The results are shown in Fig. 5.43 for open-ended corrugated
waveguides with aperture diameters of 1, 2 and 4 wavelengths, and in Fig. 5.44 for
a 4X aperture diameter corrugated horn with semi-flare angles of 15 and 30 . In all
cases a frequency near balanced hybrid has been chosen so that the crosspolarisation
efficiency is effectively 100%. The curves show the general behaviour to be expected
when a corrugated feed illuminates a reflector. Except for the IX diameter feed, the
precise numerical values are not applicable to a practical reflector because the
Design of cylindrical and conical corrugated horns 139
larger feeds would be used in Cassegrain reflectors and the scattering loss from
the sub-reflector has not been included in the calculations. The total antenna
efficiency would also be degraded by the blockage loss and the surface error loss.

100

Fig. 5.43 Efficiency for an open-ended corrugated waveguide. s/X = 027


overall efficiency — illumination efficiency spillover efficiency
Arrows indicate copolar power levels, a D/k — 1 0 b DA = 2 0 c D/X = 4*0
140 Design of cylindrical and conical corrugated horns

Figs. 5.43 and 5.44 show the compromise between illumination efficiency and
spillover efficiency. The form of Fig. 5.43(a) to (c) is similar apart from the in-
creased efficiency as the aperture size is increased. When phase loss is also present,
Fig. 5.44 shows that the additional loss in efficiency can be quite significant, which
means that a wide flare angle horn is, in general, not an efficient illuminator of a
parabolic reflector. The copolar power level is shown in the figure, the peak
efficiency in all cases occurs for an edge taper of— 10 dB. This should therefore be
the value chosen for most reflector antennas unless the spillover leads to unac-
ceptable noise.

100
uu
\ •

/ \
80 \ ^V
\ \
' / ^\ \
/ / \ \ \P
// \ \
60 \ \ \
I \ \ \
V
\

40

20
-20

/ 11i •
n/ 1 1
10 20 30
degrees

Fig. 5.44 Efficiencies for corrugated horn. Notation as Fig. 5.43


a d x = 15° 6 0 ! = 30°

G/Tparameter
The ratio of gain to noise temperature may sometimes be a more important design
parameter than the efficiency alone. This can be computed for a circular reflector
by multiplying the efficiency (as calculated above) by (TLD/X)2 and then by dividing
by the noise temperature. If the G/T value is dominated by the 'T5 value from the
spillover, it may be better to accept a lower edge taper for the feed illumination.
The total noise contribution comes from a number of sources:
(i) the noise received by the main beam of the antenna which will have a mini-
mum level set by the isotropic background radiation level of 2*7 to 3-0 K;
(ii) the spillover noise due to the spillover past the feed;
(iii) the receiver noise.
Only the second component, the spillover noise, is under the control of the feed
Design of cylindrical and conical corrugated horns 141

designer. The spillover noise temperature is given by:

r. = ra(i-j?,) (5.18)
where Ta is the noise temperature 'seen' by the sidelobes of the radiation pattern of
the reflector. This will vary according to the location and orientation of antenna.
The values of rjs given in Figs. 5.43 and 5.44 show that it usually lies between 0-85
and 0-98. This gives typical values for Ts in the range 1 K to 24K. Uncooled re-
ceivers currently have equivalent noise temperature of 100 K and above. Thus
antennas with uncooled receivers will have G/T values dominated by the receiver
noise and the spillover need not be the most important design parameter. On the
other hand, a cooled receiver may have a noise temperature which is similar to, or
less than, the spillover noise. In these circumstances the spillover will determine the
G/T value and design of the feed becomes the dominant noise controlling factor.

5.8 Horn design

In the previous sections we have considered each part of the horn and the electrical
characteristics separately. Now we shall consider the horn as a whole and discuss
the procedure for designing the different types of horn listed in Table 5.1. In most
cases this consists of collating the information in previous sections and chapters,
but for some special types, e.g. the small aperture diameter horns, additional
information is presented. The basic design procedure is similar for all corrugated
horns and is enumerated in the following section. It applies to all the types with
minor modifications. We shall assume that, unless stated, the corrugated horn is
required for use as a feed for a single or multiple reflector antenna. This
automatically means that the copolar radiation pattern is the characteristic of
principal interest, and it is here that the design of a corrugated horn starts.

Fig. 5.45 Corrugated horn with semi-flare angle of 8 degrees

5.8.1 Large aperture, narrow flare angle horns


This type of corrugated horn has seen the widest use as a feed for Cassegrain
reflector antennas. The large aperture diameter produces a narrow copolar radiation
pattern to illuminate the sub-reflector. The narrow flare angle gives a physically
long horn, which is not usually a problem with Cassegrain reflector, and also means
142 Design of cylindrical and conical corrugated horns

that the phase centre of the horn varies less with frequency (see Section 5.4). A
sketch of a typical horn is shown in Fig. 5.45. While it is not possible to be precise
about what is meant by large' and 'narrow', since these are partly tied up with
analysis and use, in general the aperture diameter should be greater than 4X and the
semi-flare angle less than 15 degrees. These two criteria mean that approximate
methods of analysis for copolar radiation pattern prediction can be used with
confidence. The narrow flare angle enables the horn to be treated as an open-ended
waveguide which then enables the fields in the waveguide to be expressed as Bessel
functions, and the radiation pattern evaluated'in closed form. The large diameter
also simplifies the analysis and enables further approximations to be made. These
have been discussed in Chapter 3 and the equations expressing the radiation
characteristics are given by eqns. 3.74—3.81.
The design procedure is as follows:

(i) Choose the aperture diameter. This is done in order to satisfy a specified edge
taper for the reflector, or to optimise the efficiency, or to give the highest G/T
ratio. The curves of Figs. 5.10-5.12, 5.42 and the discussion of Sections 5.2 and
5.7 can be used for this purpose. The copolar radiation patterns have now been
determined and this also fixes the bandwidth for which the crosspolarisation is
below a specified level (Section 5.3).
(ii) Choose the geometry of the corrugations. As discussed in Section 5.3 the
corrugation geometry determines the frequency at which the minimum level of
crosspolarisation occurs. For many applications the crosspolar frequency band-
width will be much wider than the communication channel bandwidth so the
crosspolarisation generated by the horn will be much lower than the specified level.
In these circumstances the choice of the corrugation geometry is not critical. In
other cases the curves of Figs. 5.22-5.25 can be used to choose the corrugation
geometry, and Fig. 5.28 then indicates the bandwidth of the crosspolar character-
istics. As a general guideline, for horns with Dj\ > 6, the slot depth can be 0-25A
and the ridge width to slot width ratio should then be between 0-5 and 1-0. The
slot dept is measured in the centre of the slot. The number of slots per wavelength
will be partly determined by the frequency of operation and the consequent
mechanical maufacturing restrictions. At frequencies of a few gigahertz it is possible
to contemplate ten slots per wavelength, while at frequencies of hundreds of
gigahertz even two slots per wavelength may be difficult to fabricate. We have
designed and built horns with a wide variety of a number of slots per wavelength,
and we have always found that the horns perform satisfactorily. Similarly, there is
no tight constraint in the choice of a ridge width to slot width ratio of > 0-5. If it
is more convenient to make the horn with a smaller ratio, then the only consequence
is the need to deepen the slot depth in order to maintain the resonant condition,
see Figs. 5.22-5.25. The slot depth also needs to be deeper if the aperture dia-
meter is less than six wavelengths, again see the above figures,
(iii) Choose the flare angle of the horn. Assuming that a linear taper is being used,
then the size of the aperture will already place a constraint, thus effectively deter-
Design of cylindrical and conical corrugated horns 143

mining the flare angle. Otherwise the main consideration is the movement of the
phase centre as discussed in Section 5.4. If this is not important, perhaps because
a relatively narrow bandwidth is required, then choose a semi-flare angle which will
reduce the length and hence cost of manufacture, say 10°~~15°. The patterns of a
15° semi-flare angle horn differ from the open-ended waveguide patterns so some
adjustment to the aperture size may be needed. Where the phase centre movement
is important a semi-flare angle as small as feasible should be chosen, say 5°.
(iv) Design the junction and throat section of the horn. This was covered in Section
5.5 and the procedure described in that section can be followed. If the ability
exists to choose the diameter of the smooth wall waveguide then the matching
problem is made easier by choosing a diameter which is well away from the cut-off
of the TE U mode, but not large enough to excite high order unity azimuthal
dependent modes.

Attentuation: A long corrugated horn will have some loss of power due to
attenuation along the horn. This is of particular concern at millimetre and sub-
millimetre wavelengths, partly because the loss in the metal increases with
frequency, and partly because it is practical to make a horn which is long in wave-
lengths. The attenuation characteristics have been dealt with in Section 3.4 and are
considered in detail in references 39 and 40. The corrugated waveguide propagating
a balanced hybrid HE n mode has inherently low attenuation characteristics. This
means that there are very few cases where the attenuation is likely to be significant
for the corrugated horn user. This is confirmed by measurements carried out at
100 GHz by Chu and Legg [24]. They measured an attenuation of only 0-2 dB due
to a 76X long corrugated horn.

5.8.2 Small aperture horns


Small aperture horns, with a narrow flare angle, are effectively open-ended
corrugated waveguides. They can be used as prime feeds for reflector antennas; an
example is the antennas used on the European OTS satellite. The design procedure
is identical with that for large aperture horns, set out in the previous section.
However, in this case, the characteristics are very dependent on the horn
dimensions, in particular the crosspolar characteristics will have a narrow band-
width, as is evident from Fig. 5.27. The slot depth required for resonance will be
quite deep as Figs. 5.22-5.25 indicate. This does, however, mean that the matching
of the corrugated horn to a smooth-wall waveguide is made easier because the dif-
ference between the half wavelength deep slot required for a match and the
resonant depth is less than for a large aperture horn.
The characteristics of the small aperture horn are very dependent on the
influence of the flange, since the flange is a significant proportion of the total area
of the aperture. This has consequences both for analysis and performance. The
method used for predicting the radiation patterns must include the effect of the
corrugations and the effect of the flange. The Fourier Transform technique should
be used since this assumes an aperture is an infinite ground plane, which is, in most
144 Design of cylindrical and conical corrugated horns

cases, a good approximation to the finite flange. The currents on the finite flange
have decayed to a low level by the time they reach the outside edge of the flanges.
In cases where this is not true, say for feeds with diameters of under 1.4A, the
effect of the finite flange can be taken into account by adding a diffracted term to
the radiation from the aperture. The aperture fields cannot, in general, be approxi-
mated by simple formulae and it is necessary to take account of the space harmonic
to obtain a reliable estimate of the crosspolar levels.

Fig. 5.46 Small diameter corrugated horn (see text for dimensions)

-72° -36° 0° 36° 72°


Fig. 5.47 Measured copolar radiation patterns of horn of Fig. 5.46. Parameter: frequency,
GHz
The cross-section of a horn with an aperture diameter of 1-36X at the design
frequency of 14-25 GHz is shown in Fig. 5.46.
The horn is 3-25X in length and the semi-flare angle is 5-4°. The figure is drawn
to scale. The slot depth is 0-39X at the aperture and tapers linearly to the throat.
Design of cylindrical and conical corrugated horns 145
Fig. 5.47 shows the measured E plane patterns over a range of frequencies and Fig.
5.48 shows the measured and computed (using full space harmonic model) 45°
plane crosspolar patterns.

-72° -36
Fig. 5.48 Measured and computed crosspolar radiation pattern at 140 GHz for horn of Fig. 5.46
computed

The minimum peak crosspolar level was measured as — 37 dB, a level set by the
radiated contribution from the edge of the flange.
A way to reduce this level is to add a small quater-wavelength deep slot, or
choke, in the face of the flange [68, 69, 71,125] so that the current flow is inter-
ruted. Because of the need to limit the flange thickness the choke is re-entrant as
shown in Fig. 5.49. This figure also shows the effect of the choke on the crosspolar
pattern of a 1-7X diameter open-ended corrugated waveguide. The crosspolar level
has been reduced by 8 dB. The theoretical results shown in Fig. 5.49 are due to
Hockham [71] and have been obtained with a matrix solution relating the fields in
the aperture to the field in the choke slot. An alternative technique is given by
Thomas [144, 149] who extended the corrugations around the corner onto the
flange.

5.8.3 Wideflareangle horns


Wide flare angle horns have copolar patterns which are determined mainly by flare
angle rather than aperture size. Apart from this fact, the design procedure is similar
to the two previous types. The beam width curves of Figs. 5.10—5.12 can be used to
determine the horn semi-flare angle and the horn length. Although it is possible to
design a wide band horn, the boresight directivity is a maximum at the minimum
half beamwidth which occurs at a relatively small aperture size where the constant
beam width does not exist. The peak boresight directivity enables an optimum size
to be determined for any semi-flare angle, eqn. 5.15. This is shown in Fig. 5.50 and
146 Design of cylindrical and conical corrugated horns

the corresponding - 3 dB, ~ 10 dB and - 2 0 dB half beamwidths in Fig. 5.51.


Note that the — 15 dB half beamwidth level is approximately equal to the semi-
flare angle of the horn.

-36
(
36° 72°
Fig. 5.49 Effect of choke in flange on the crosspolar radiation pattern of a 1.7X diameter
corrugated horn
measured theory

The corrugation geometry is designed in the same way as for large aperture
corrugated horns, namely by choosing the slot depth according to the radial slot
geometry. The crosspolar characteristics are frequency dependent so the wide band
properties will only apply to the peak crosspolar level if the aperture size is large in
wavelengths. It is normal with wide flare angle horns to cut the slots perpendicular
to the wall of the horn rather than perpendicular to the z axis, but this does not
modify the design procedure.
The phase centre in a wide flare angle horn will remain essentially fixed at the
apex. This can be an advantage since, unlike the narrow flare angle horns, the phase
centre will remain fixed as frequency changes. A comparision of efficiencies shows
that the wide flare angle horn has a lower efficiency than the narrow angle horn but
if the moving phase centre is taken into account, the wide flare angle horn can
perform better as frequency is increased above the design frequency.
The throat section of the horn is crucial because the change from a smooth wall
waveguide to the conical section can be very abrupt and thus may give rise to higher
order modes which will considerably distort the radiation patterns (this is discussed
Design of cylindrical and conical corrugated horns 147

20 U0 60 80
semi-flare angle, degree
Fig. 5.50 Optimum normalised aperture diameter against semi-flare angle for wide angle
corrugated horn

20 40 60 80
semi-flare angle, degrees
Fig. 5.51 Half beamwidths corresponding to Fig. 5.50
148 Design of cylindrical and conical corrugated horns

in detail in Section 4.4). We have found that the design of the throat section can be
performed using the procedure of Section 5.5. This will normally mean that a small
section of smooth-wall conical horn will be needed between the smooth-wall wave-
guide and the conical corrugated section. Fig. 5.52 shows the cross-section of a 30°
semi-flare angle horn with very low mode conversion properties. This is confirmed by
the measured peak crosspolar characteristics, Fig. 5.53, which follow the
theoretical prediction (using a spherical wave expansion method, Chapter 4) very
well. Note the low level of crosspolarisation over a appreciable band of frequencies.

Fig. 5.52 Scale diagram of 30 degree semi-flare angle corrugated horn

The slots are of constant depth along the horn. This may appear to be in con-
tradiction to the previous discussions, but if the inner radius of the first slot (0-55X)
and the slot depth (0-27X) are plotted on Fig. 5.36, we can see that it has a guide
wavelength nearly identical to the smooth-wall waveguide value (Xg/X = 1 -4). The
measured return loss at the design frequency is —30 dB.
Wide flare angle horns can be made to work when they are very small in wave-
lengths. This has been demonstrated theoretically and experimentally by Estin et al
[61] who used a 45° semi-flare angle horn with an aperture diameter of between
1-2X and 2-8X. The peak crosspolar level at midband was —30 dB, a very good level
Design o f cylindrical and conical corrugated horns 149

-60

Fig. 5.53 Measured peak crosspo/ar power against frequency of 30° horn of Fig. 5.51
theory

HE,,

H O1 HE 21 E O2

n
TM O i
TE,,
TE 01

Fig. 5.54 Multimode corrugated horns


a Horn with tracking modes b Monopulse corrugated waveguide with four feeder
waveguides
150 Design of cylindrical and conical corrugated horns

for such a small horn. The horn was investigated for use as a gain standard at low
microwave frequencies, when the physical size of an antenna becomes the limiting
design factor.

5.8.4 Multimode horns

Tracking horns: Tracking horns use a number of modes to provide tracking or


monopulse information, which may be in addition to a signal mode or it may be
the only function. Thus a primary requirement is that the horn should support
all the modes throughout the length of the horn. This is clearly no problem at the
aperture but means that the minimum size of the throat region is determined by the
mode with the highest cut-off frequency. A major design limitation with these
types of horn is the need to separate the power from the various modes into the
tracking channels. The separation can be done in two ways. Either the horn has a
cross-section similar to a standard horn with the modes being transformed from the
corrugated section to the corresponding modes in smooth-wall rectangular wave-
guide (Fig. 5.54(a)), and the signal extraction is then done in smooth-wall
waveguide. Or the end of the corrugated horn is terminated abruptly and a number
of smooth-wall waveguides butt directly on to the corrugated section (Fig. 5.5 4 (b)).
The smooth-wall waveguides then carry the tracking modes. An example of the
former approach is Watson et al [161], where a sophisticated mode extraction
network is described, and an example of the latter approach is Clarricoats and
Elliot [25]. Watson et al needed a corrugated horn which had a primary function
of transmitting or receiving a communication channel so the design of the horn
followed standard procedure except for the need for an oversized throat region. In
contrast, the horn described by Clarricoats and Elliot consisted of a length of
corrugated waveguide with no flare. The end was terminated directly in a four horn
cluster. Because there was no significant bandwidth requirement this was the easiest
approach. The length of the waveguide and the inner and outer diameters are
chosen to propagate the modes in the correct phase relationship. The principles of
this type of horn have been covered in Chapter 3.
A modification to the four horn cluster feed is to add a fifth horn [50, 154].
This improves the tracking performance but involves some compromise in the
corrugated horn design. A thorough investigation has been carried out by Viggh
[154].
The design procedure for a tracking horn follows the same principles as enumer-
ated for large aperture horns, except that the design of the copolar and crosspolar
patterns are replaced by the design of the modes required for tracking. This means
an aperture field which radiates a null on boresight. Since the essential feature is the
null, the level and position of the peak on either side of boresight is not normally of
concern, although the slope of the difference pattern with angle is important. The
form of the radiation patterns of the HE 2 i, E^ and HOi modes are shown in Fig.
5.16. These have normalised to 0 dB, but the shape of all the modes is seen to be
nearly identical.
Design of cylindrical and conical corrugated horns 151

The throat section can be used as a filter of unwanted modes by choosing the
inner diameter to just propagate the tracking modes. It is desirable for the tracking
modes to have the same differential phase shift as they propagate down the horn.
This implies that the dispersion characteristics should be nearly coincident. An
example is shown in Fig. 5.55 where the E02 and HE2i modes have similar propa-
gation coefficiencies over a range of frequencies. Design of the tracking horns is
best accomplished for the particular specification by using a computer program. The
simple waveguide model is adequate for this purpose. The corrugation geometry is
normally chosen to be resonant for the HE n mode at the design frequency, even
though the higher order modes may have a different resonant depth.

Fig. 5.55 Dispersion characteristics of HE2l and Eo2 modes in corrugated waveguide with
rx/rQ =0 6

Shaped pattern horns: A class of multimode horns have been used for prime focus
paraboloidal reflectors. These produce a small dip on boresight and peak power a
few degrees off boresight [132, 157]. The purpose of this pattern is to overcome
the space attenuation loss in a paraboloidal reflector and produce nearly uniform
illumination over the reflector surface, as illustrated in Fig. 5.56. The pattern can
be obtained by exciting a proportion of HE12 mode which is added to the main
HE H mode. This is normally done in a section of corrugated waveguide abutting
152 Design of cylindrical and conical corrugated horns

angle
Fig. 5.56 Feed pattern required to increase aperture efficiency in a prime focus reflector
aperture field field from feed
0

40
6,deg
theoretical •
Fig. 5.57 Patterns of two hybrid mode horn
Relative phase of modes: a - 0° b - 36° c - 72° d - 108° e - 144° f — 180°
Design of cylindrical and conical corrugated horns 153

directly onto a smaller diameter smooth-wall circular waveguide. The step excites
the HE12 mode in the correct proportion. Fig. 5.57 shows the patterns which can
be obtained by adding together HEn and HE12 mode power in various phases.
The pattern marked 'a' is the desired pattern. The junction between the corrugated
waveguide and the smooth-wall waveguide should be designed with a modal
matching procedure [49].
A dual hybrid mode (HE n + HE12) horn is described by Thomas and Bathker
[153]. This used a step junction between a smooth-wall circular waveguide and a
conical corrugated horn with a semi-flare angle of 6-25°, to generate the HE12
mode. The horn exhibited 1% improved spillover efficiency and 5% improvement
in illumination efficiency by comparison with the equivalent single mode horn
feeding a Cassegrain antenna. Overall the gain of the antenna improved by 0-36 dB.

Matched feeds: Another type of shaped pattern feed is the 'matched feeds'
invented by Rudge and Adatia [189]. These are designed to produce a copolar and
crosspolar pattern which matches the focal region fields of an offset reflector, so
eliminating the inherent crosspolarisation present in an offset paraboloid. Although
the matched feeds have been mainly designed using smooth-wall horns, corrugated
matched feeds can be produced. A proportion of HE2i mode is added to the
fundamental HE n mode to give the desired pattern. The theory and design is fully
covered in Reference 132.

5.8.5 Broadband horns


Horns which have nearly constant copolar radiation patterns over a very wide band
of frequencies can be produced. We should emphasise that the broadband charac-
teristic applies only to the copolar characteristics. The crosspolar characteristics,

GHz
Fig. 5.58 Half beamwidth and peak crosspolarisation against frequency of a broadband
corrugated horn with 15 degree semi-flare angle and aperture diameter — 262 mm
being determined by the resonant nature of the corrugations, are necessarily
relatively narrow band. Reference to the — 3 dB beamwidth curves of Fig. 5.10 will
show that a horn with a semi-flare angle of between 10° and 20° has a constant
beamwidth over a wide range of D/X. The optimum semi-flare angle is 15° when
the - 3 dB half beamwidth is between 5° and 6° from D/X = 7 to D/X = 14. The E
154 Design of cylindrical and conical corrugated horns

plane, H plane and 45° crosspolar patterns across the band are as shown in Fig.
5.58. Because the aperture is large, the theoretical crosspolarisation is, in fact,
low across the whole of the band.
The throat region is a limiting factor for broadband horns. If the first slot is
chosen to be half a wavelength deep at the top end of the frequency band then it is
found that the HE U mode is cut off at the lowest frequency. Thus the slots must
be greater than half a wavelength deep at the upper frequency, a consequence
which could give rise to mode conversion and thus a useful range of frequencies less
than one octave. We are not aware that corrugated horns have been measured over a
very wide range of frequencies, so we have no practical evidence of the limits of
operation of corrugated horns.

5.8.6 Multi-frequency horns


This type of horn is distinguished from the previous type in that the horn is required
to operate at a number of discrete frequencies rather than over a wide band of
frequencies. The corrugated horn has the ability to operate whenever the slots are
(2/2 — l)X/4 deep (n = 1, 2. . .). An example of a large dual frequency band corru-
gated horn is provided by the feed for some of the NASA Deep Space Network
reflector antennas. This is described by Williams and Withington [163] and
operates, simultaneously, from 2 0 GHz to 2-4GHz (where the slot depth isO-34X
to 0-4X) and 7-1 GHz to 8 6 GHz (where the slot depth is 1-21X to 1-44X). The
semi-flare angle of the horn is 17*1° at the 1*07 m aperture. Thus the copolar
patterns are functions of angle, resulting in similar beamwidths at the two operating
bands.

-20-

* -30 \ P\
a
\\
\\ \ x / /
-40- s\///

-50
8 10
\ 12 16 18

Fig. 5.59 Dual depth corrugated horn


a Cross-section b Peak crosspolar power rx = 41*5 mm, b = 3 mm, t = VBmm,
Sj = 8-5 mm, sjs2 =1*5, sjs2 = VI, sjs2 - 1-9

For many applications the specified channel spacing does not satisfy the above
slot depth criteria. An example is the transmit and receive bands for direct broad-
cast satellites. An alternative approach is to use a dual depth corrugated horn, Fig.
Design of cylindrical and conical corrugated horns 155

5.59(a) [5, 63, 64, 124]. One of the slot depths is resonant at one frequency, while
the adjacent slot is resonant on a second frequency. Thus reasonable performance
can be expected at two discrete frequencies. An experimental model has been
measured by Ghosh et ah [63] who showed that low crosspolarisation could be
obtained in two frequency bands, spaced 1:1-5 apart. The theoretical prediction of
the crosspolar performance needs to be performed using the space-harmonic model
because the influence of the dual depth slots cannot be accurately represented by a
surface impedance. We have extended the space harmonic analysis given in the
Appendix to the dual-depth case. The characteristic matrix equation which must be
solved to determine the propagation relationships is extremely complicated because
each term in the matrix is complex. The space harmonics modify the frequency at
which minimum crosspolarisation occurs in a manner similar to the results presented
in Section 5.3 for the standard corrugated waveguide. Fig. 5.59(b) shows typical
results for fixed slot width and ridge width and varying depth of the shallower slot.
The ratio of the upper resonant frequency to the lower resonant frequency is seen
to correspond to the ratio of the slot depths. The results are for an open-ended
corrugated waveguide but will apply also for narrow flare angle corrugated horns.

5.8.7 Compact horns


The compact, or profiled, horn, Fig. 5.60, has already been discussed in Sections

Fig. 5.60 Profiled corrugated horn

4.4.4 and 5.4. The design of the horn is straightforward, following the standard
procedure laid out for large aperture horns. That is, design the aperture diameter
for the copolar pattern using open-ended corrugated waveguide results; then design
the aperture corrugation geometry for the crosspolar characteristics. The flare
section is designed according to eqn. 5.10 and the throat section is designed ac-
cording to Section 5.5. It is advisable to allow for a small length of corrugated
156 Design of cylindrical and conical corrugated horns

/-10

loo
I
1 -20

/
rf\ / -30
;
\ v \
\
/

\
\
\
1 \1 \
/ » I 1
/ ^
1
1 ^ I i
\ J+20dB
\
1 t

-72° -36° 36° 72°

Fig. 5.61 Measured radiation patterns of profiled horn (see text for dimensions)
copolar crosspolar

9-5 10 10-5
GHz

Fig. 5.62 Measured peak crosspolarisation of profiled horn of Fig. 5.61


Design of cylindrical and conical corrugated horns 157

waveguide at the aperture, after the profiled section, to ensure that the phase centre
remains fixed at the aperture. Measured copolar and crosspolar patterns of a 2-9X
diameter horn are shown in Figs. 5.61 and 5.62. The horn has an aperture diameter
of 89 mm, a slot depth of 8-7 mm, a slot width of 3 mm and a ridge width of
1 mm. At the throat the slot depth is 13-4 mm and the diameter of the smooth-wall
waveguide is 33 mm. The total length of the corrugated horn is 129 mm which
includes a 17mm waveguide section at the aperture. The measured peak cross-
polar level is seen to be less than —40 dB over a 0-75 GHz band. The excitation of
some EH12 mode limits the high frequency performance. As discussed earlier this
is a consequence of using a non-linear taper. It can also be seen from Fig. 5.61
that the copolar patterns are distorted by the presence of some HE12 modes. This
causes the sidelobe at 40 degrees. Fortunately it is outside the angle of the beam
needed for illuminating a reflector so is not serious.

5.9 Numerical prediction of performance

The design of corrugated horns usually involves a first stage of analysis, then a
second stage of numerical prediction of the theoretical performance and finally a
third stage of manufacture and testing. In this section we shall collect together and
summarise the available methods for numerically predicting the characteristics of
corrugated horns. In order to predict the radiation patterns, the propagation
coefficient in the circular corrugated waveguide or the conical corrugated wave-
guide must first be computed, and then aperture fields predicted. From these fields
the near-field or far-field radiation patterns can be computed. The design of the
throat section also requires a knowledge of the propagation coefficient.

5.9.1 Propagation characteristics


(a) Corrugated circular waveguides: When the normalised diameter of the corru-
gated waveguide is large, and the waveguide is operating near to the balanced hybrid
condition the propagation characteristic can be estimated from eqn. 3.37, i.e.

= 0
where q = 0 for HE l m modes and q = 2 for KH lm modes, and the wavenumber
x\ — {kri)2 — (fri)2. This equation has the advantage that the relationship between
krx and $rx is in closed form, a situation not normally encountered in inhomo-
geneous waveguides. For the modes of interest the values of xx are given by:
Mode Xi
HE,, 2-405
EH 12 5-135
HE, 2 5-520
EH 13 7-016
HE, 3 8-654
158 Design of cylindrical and conical corrugated horns

The equation has a surprisingly wide range of validity. Even for frequencies far
from the balanced hybrid condition, the values predicted are often good estimates
of $rx and krx. Dragone [56] has derived a series relationship for xx which gives an
improved estimate for the wavenumber. The first two terms are

x = xx ( l -
2krx

where Y is the admittance of the corrugations. However, the relationship is only


valid near balanced hybrid (where Y — 0) and for large krx so the second term
contributes only a small perturbation to x ^ Higher order terms converge slowly
and need using with care.
A more accurate computation of the propagation characteristic involves the
solution of a transcendental equation. The formulation based on the surface
impedance approach is given by eqn. 3.26. In order to solve this equation for fixed
krx trial values of $rx must be used and an iterative scheme used to converge on the
correct value of firx. The equation is valid for all m and both fast and slow wave
regions of the propagation characteristic. A sample computer program is given
below. The main computational difficulty lies in locating suitable trial values for
$ru and experience is the only reliable guide as to how to make the choice. Because
the propagation equation involves combinations of Bessel functions there can be
regions where poles occur and the value of the equation changes very rapidly as the
parameters are altered. The solution to this problem is to rewrite the equation in
such a way as to remove the pole.
The most accurate method of computing the propagation coefficient in corru-
gated waveguides is to use the space harmonic formulation given in Appendix. The
characteristic equation is given by eqn. A29. A determinant must be evaluated.
Each term of which is a combination of Bessel functions. The procedure for solving
the propagation coefficients is similar to the surface impedance case with the
difference that the computational time is longer due to the increased complexity of
each term. The space harmonic formulation is mainly of use for accurate prediction
of the aperture field and then of the crosspolar radiation patterns. It is also useful if
the corrugated waveguide has non-standard corrugation geometry so that the
influence of the slots cannot be represented as an impedance surface.

(b) Corrugated conical waveguides: A closed form solution for the spherical
wavenumber, p9 is given by eqn. 4.19, namely:
2-405
* = ,„ -:._n M/2-Q-5 (forHEn mode)

This gives values to within 1% of the true values for semi-flare angles up to 60°. The
formula is valid for small flare angles (< 10°) but for these horns it is better to
approximate the conical horn as an open-ended waveguide and use the simplier
circular waveguide function.
Design of cylindrical and conical corrugated horns 159

To derive more accurate values for p, the characteristic eqn. 4.14 must be solved.
This is a transcendental equation with Legendre functions, and is solved in the same
manner as the characteristic equation for corrugated circular waveguides. See
Table 4.1.

(c) Program to compute propagation and attenuation characteristics: The


program listed in Program 1 computes the propagation and attenuation charac-
teristics of corrugated circular waveguides by implementing the surface impedance
formulation, eqn. 3.26, and then using perturbation formulae equations, eqns. 3.66,
3.68-3.71, to calculate the attenuation in the waveguide. The program is written in
BASIC and is suitable for running on many types of personal computers. It
computes the value of prx at a fixed frequency. An initial value of f}rx is either
input as data or calculated using xx =2-405, then the values of characteristic
equation are calculated as firx and is incremented until a root is found. Conver-
gence on to the root is done with a Langrangian inverse interpolation procedure.
The input data is specified in Table 5.5 and sample data and results in Table 5.6.
The program can be used to compute the values of firx appropriate to the throat
region of the horn when the simple approximation of eqn. 3.37 does not apply or
when the modes with non-unity azimuthal order are required.

Table 5.5 Input data for Program 1


N = Azimuthal order
RA = Slot depth (metres)
Rl = Inner radius (metres)
D = Slot width (metres) (used for attenuation calculation)
TT = Ridge width (metres) (used for attenuation calculation)
RH = Conductivity of walls (mhos/m)
FR = Frequency (GHz). If FR = — 1 jumps to start of data.
C(l) = Initial trial for r^ If C(l) = — 1 calculate from xx =2*405.
After computation returns to line 1110. The accuracy of computation is set by line
1290

(d) Bessel functions: The computation of the propagation and radiation charac-
teristics of circular waveguides need Bessel functions. Since most of the computer
time is taken up with computing the Bessel function it is important to use an
efficient routine. For inhomogeneous waveguides, Bessel functions of the first and
second kind (/ and Y) are needed and also the corresponding modified functions
(/ and K). There are two main methods of computing the functions, either using a
polynominal approximation or using a recurrence relationship. The polynomial
approximation method (e.g. Abramowitz and Stegun [178], Chapter 9) requires
a reasonable amount of memory to store the coefficients. It is the method usually
implemented in the maths libraries of large computers but is not suited to small
computers. The recurrence relationship using backward recursion is an efficient
160 Design of cylindrical and conical corrugated horns

Table 5.6
RUN
? 1,.O1,.O3,.OO7,.OO3,1-557E7
? 10, - 1

Corrugated Waveguide
Azimuthal No. = 1
R1/R0 = 0-75
Rl = 0-03 metres
Slot depth - 0-01 metres
Slot Width = 7 E - 03 metres
Ridge Width = 3 E - 03 metres
Conductivity = 1-557E + 07
Frequency = 10 GHz
k*rl - 6-28318
Beta*rl = 5-84885
Hybrid factor = 1-22057
Attenuation = M8206E - 04 dB/m

method for the typical argument values needed in corrugated horn design. Also,
when more than one Bessel function is required, the same formulae can be used
to generate all functions. This method is implemented in Program 1, lines 2160 to
2520, which generate Jm (x), Ym (x), Im (x), Km (x) and their derivatives. The code
is compact and has the advantage that the accuracy can be changed to suit
individual circumstances. This is done by changing the argument in line 2230. The
accuracy, as given, is six decimal places.
A compact simplified version of the same routine to calculate J0(x) a n d / ^ x ) is
contained in Program 2, lines 1400 to 1500. This routine is particularly useful for
radiation pattern prediction programs.

5.9.2 Radiation characteristics

(a) Open ended waveguides and narrow flare angle horns: The copolar and
crosspolar radiation patterns can be predicted either by the Kirehhoff-Huygen
method or by the Fourier Transform method. The Fourier Transform method is
better because it requires only the transverse electric fields in the aperture of the
horn, and because the assumption of an aperture in an infinite ground plane is a
good approximation to the finite flange which must be present on corrugated
structures. The Fourier Transform method also leads to expressions which are easy
to compute for open-ended waveguides They are given in eqns. 3.74—3.81, for the
case when space harmonics are neglected. The values of J3 and co can be found either
by the balanced hybrid formula, eqn. 3.37, or the surface impedance characteristic
eqn. 3.26.
Design of cylindrical and conical corrugated horns 161

The Fourier Transform method can also be applied to the space-harmonic


formulation to give analytic expressions for the radiated fields, and these are given
in the Appendix. This formulation is needed for an accurate prediction of the cross-
polar characteristics when either the corrugation geometry is unusual (e.g. for few
slots per wavelength) or the aperture diameter is small so that the slots have to be
deep in order to resonate.
When the horn flare angle is significant a spherical phase factor can be added to
the aperture fields to improve the prediction of the copolar sidelobe region. This
has been studied in Section 4.7 and eqn. 4.112 gives the phase factor. The radiated
field components are given by eqn. 3.73 with integrand multiplied by exp (j 0(r)].
The radial integral is no longer given in an analytic form and numerical integration
is needed to obtain the far-field results. The aperture field of corrugated horns is
well behaved so most integration routines (e.g. Simpson's rule) converge rapidly.

(b) Wide flare angle horns: When the semi-flare angle of a horn is greater than
about 15° the spherical wave expansion method must be used. The expressions for
the radiated fields, are given in eqns. 4.39-4.41. These expressions are valid at any
distance R from the aperture and mean that near-field patterns or far-field patterns

Table 5.7
RUN
Radiation Pattern of Corrugated Waveguide
Inner radius in Wavelengths = ? 3*0
Slot depth in Wavelengths = ? 0-28
Slot admittance = - 0-190764
Hybrid factor = 1-03346
Increment angle, Final angle = ? 2-20

ANGLE COPOLAR X-POLAR


(dB) (dB)
0 0 -100
2 -0-290878 - 69-4263
4 -1-17578 -58-0885
6 - 2-69489 -52-2385
8 -4-92867 -48-961
10 -8-02951 -47-3925
12 -12-3091 -47-2265
14 -18-5556 -48-4368
16 -30-2667 -51-3027
18 -33-8935 -56-9102
20 -27-8657 -74-0154
162 Design of cylindrical and conical corrugated horns

can be computed using the spherical Hankel functions and associated Legendre
polynomials. Neither function is time consuming to compute but these routines are
not normally available in standard computer software libraries.

(c) Computer program for radiation characteristics of narrow flare angle horns:
A simple BASIC program to compute the copolar and crosspolar radiation patterns
of the HE n mode in narrow flare angle corrugated horns or open-ended corrugated
waveguides is listed in Program 2. This is based on eqns. 3.80 and 3.81 with the
wavenumber computed using the first two terms of eqn. 3.39. The input data is
self-explanatory and typical results are given in Table 5.7. The program is very
quick to run and is well suited to a small personal computer. It should be noted
that the crosspolar values will only be strictly valid for a practical horn with a
large number of corrugations per wavelength and approximately equal slot width/
ridge width as discussed in Section 5.3.
Chapter 6

Manufacture and testing of


corrugated horns

6.1 Manufacture of corrugated horns

Corrugated horns are not easy, nor cheap, to manufacture. Considerable manpower
needs to be devoted to the production of a horn. Changes in the design after
manufacture are hard to implement and there is an obvious desire to produce a
design, which, when made and tested, will agree with the theoretical predictions.
The corrugations are designed for operation over a specified frequency band. It
thus immediately follows that all dimensions and tolerances on dimensions need
to be considered in relation to wavelength rather than absolute distance. We will
deal with tolerances in more detail later in this section. To start with we shall
discuss some of the methods which have been used to manufacture corrugated
horns.

Machining
Making the horn by machining, Fig. 6.1, is the most common method of manu-
facture for horns designed to operate at microwave frequencies, say below 20 GHz.
Aluminium is the preferred material because of its relatively low density, but
sometimes brass is used for small or intricate horns. Long horns need to be made in a
number of sections and joined together. The part of the horn which presents the
greatest difficulty in machining is the throat region. The requirement of an approxi-
mately A/2 deep slot at the throat, when the inner diameter of the corrugated
section is only about one wavelength diameter, limits the size and shape of the
cutting bar which can be inserted into the horn. This is particularly a problem
when the slot is narrow (i.e. a few millimetres). In this case the cutting tool tends
to drag on the metal, either distorting the walls of the slot or giving an uneven
surface to the slot walls and base. Since one mistake can ruin the complete horn,
considerable care is needed when cutting the throat section. One solution is to
make the horn in a number of separate sections and join them together. We have
made horns successfully by this method, even when each section consisted of
one slot and one ridge, Fig. 6.2, but there are disadvantages. The outer diameter
needs to be increased in order to take the fixing screws. This increases the weight.
164 Manufacture and testing of corrugated horns

Each section needs to have ends which are very flat and perpendicular with the
horn axis, otherwise some slight bending of the horn can occur. Since this will
excite azimuthally dependent modes it must be avoided. Each section needs to be
bedded firmly to the next section so that there is no gap anywhere around the
periphery. A gap will lead to a discontinuity in the current flow which then distorts
the fields in the slot and can cause effects out of all proportion to the gap size.

Fig. 6.1 Corrugated horn machined in one piece

We have used a thin conducting paste between sections to electrically seal the
gap between sections.

Electroforming
Corrugated horns have been successfully electroformed, particularly for use at
millimetre wavelengths. A mandrel of aluminium is turned with the inverse groove
pattern and then copper, or other metal, is electroformed onto the mandrel. The
mandrel is finally removed with a solvent. The problem is to ensure that the
deposited metal is 'thrown' evenly down to the base of the slots in the mandrel.
For this purpose it is better to have a horn with equal slot and ridge widths. At
the throat region the slots cannot be very narrow, otherwise the metal will not
go down to the base of the half wavelength deep slots.
A hybrid electroforming/machining technique has been described by Dragone
[54] for making a corrugated horn, designed for 100 GHz. The mandrel is made
up of brass discs (to form the ridges) and aluminium spacers (to form the slots).
A copper wall is then electroformed onto the composite mandrel. It adheres to the
brass discs but not to the aluminium. The inner is then machined to the correct
Manu facture and testing o f corrugated horns 165

cone and finally the aluminium spacers are removed by solvent. This method has
the advantage that thin ridges can be made but the disadvantage that good contact
around the base of the slots is difficult to guarantee. As mentioned above,
interruption to the current flow can be highly damaging to the operation of the
horn.

Fig. 6.2 Corrugated horn manufactured in sections

As the frequency of operation is increased above 100 GHz, the construction


of mandrels require some care. Often slots are required which have widths of less
than 200 jum. These are difficult to machine on a lathe. A spark erosion technique
has been successfully used by Wilde*, Fig. 6.3. A computer controlled wire spark
erosion machine is used to cut the slots. A thin brass or molybdenum wire is
continuously moved over the surface of the slowly rotating mandrel. A high
frequency pulsed voltage is applied to the wire causing controlled erosion of both
the wire and the mandrel. The wire is continuously being replaced by fresh wire
so it acts as a bandsaw. The process can cut slots lOOjum wide and 500 jitm deep,
making it feasible to construct corrugated horns at 300 GHz.

Other techniques
Small corrugated horns for use at a few gigahertz have been made by casting.
Because the wavelength is relatively long at these frequencies it is usually unnecessary
to do anything with the surface. The finish which can be obtained by casting is
adequate for making a feed for a prime focus reflector intended for direct boradcast
satellite reception. Casting is thus a cheap and simple way of making horns in
large numbers.
* Wilde, R. (private communication)
166 Manufacture and testing of corrugated horns

Corrugated waveguides have also been made by a thermal contraction process.


The ridges are formed from rings of brass or copper which are held on a specially
constructed mount. The outer wall consists of a length of smooth wall circular
waveguide. The brass discs have a slightly greater outer diameter than the inner
diameter of the waveguide. The rings are cooled by immersion in liquid nitrogen
so that they contract enough to allow insertion into the circular waveguide. When

Fig. 6.3 Millimetre wave corrugated horn made by electroforming for use at 100 GHz
(Courtesy Thomas Keating Ltd.)

the rings re-expand they grip the wall of the waveguide. The method was used
to make lengths of waveguides for low attenuation experiments at 10 GHz. Brass
rings (coefficient of expansion 15-7 x 10"6/°C) of diameter 100-065 mm, were
inserted in a copper waveguide of inner diameter 100 mm, Fig. 6.4. The samples
worked well in the laboratory showing that the method does give good contact
between ridges and wall.
Manufacture and testing of corrugated horns 167

An alternative to the above method is to use r.f. heating as a way of soldering


rings (to form ridges) to an outer smooth wall.

Tolerances
The tolerances on a corrugated horn can be considered in three ways; the absolute
tolerances in millimetres, the relative tolerances in wavelengths and the consistency
of manufacture from one part of the horn to another part. It has already been
mentioned that the absolute tolerances are not as important as the relative tolerances
in wavelengths. Thus the absolute tolerances become progressively tighter as the
operating frequency is increased. However, the relative tolerances do not need to
be as tight as is often supposed. We will now examine briefly the effects of changes
in each of the horn diameters.

Fig. 6.4 Rings which will form ridges in corrugated waveguide made by thermal contraction

Changes in the horn aperture will modify the copolar radiation patterns but
hardly affect the crosspolar performance. Reference to the beamwidth curves,
Figs. 5.10-5.12, shows that changes in diameter produce a proportional change in
beamwidth. For corrugated horns, at the balanced hybrid frequency an approximation
to the main beam copolar pattern is given by the function cos7(0-48 (D/X)0) where
6 is in degrees. This allows an easy assessment to be made of changes in pattern
shape. For instance, a 5% increase in the diameter of a two wavelength horn will
decrease the — 10 dB beamwidth by 5%.
Changes in the slot depth influence the crosspolar radiation characteristics
as reference to Figs. 5.22-5.25 will show. The slot depth determines the resonant
frequency at which minimum crosspolarisation occurs so changing the slot depth
moves this minimum with frequency. The above figures can be used to assess the
affect, which is quite severe for large diameter horns. However, in compensation,
168 Manufacture and testing of corrugated horns

a large diameter horn has a wide crosspolar bandwidth, so the change in frequency
of the minimum is less important than appears.
Changes in the slot width and ridge width have less affect than the slot depth,
but again they move the crosspolar minima, as shown by Figs. 5.22-5.25. They
have negligible influence on the copolar patterns over the whole range of feasible
values.
The consistency of all the dimensions along the horn is far more important than
the relative or absolute tolerances. An error in the dimensions of only one slot
acts as a discontinuity in an otherwise smoothly changing horn. The discontinuity
can be expected to generate high-order modes if it is at a section of the horn
where the higher-order modes are capable of propagating. The performance of the
horn is then adversely affected due to one error. We have experienced many cases,
particularly at high frequencies, where the horn has been rendered useless by one
discontinuity. It seems that horns made by electro forming are more likely to have
isolated errors than machined horns, perhaps because the electro forming has
failed to form one part of one slot.

Shape of corrugations
There are a few features concerning the shape of the slots which need discussing.
It has already been mentioned that there is no theoretical reason why the
corrugations should be cut perpendicular to the horn axis, Fig. 6.5(a), or
perpendicular to the walls of the horn, Fig. 6.5(b). Since it is normally easier to
make narrow flare angle horns with the former slots, this has become common
practice.

Fig. 6.5 Shapes for corrugations

The electrical depth of the slots is not always easy to assess, particularly for
wide slots and wide flare angle horns. It seems best to use the mean depth of the
slot at its centre as the design depth.
Although it has been assumed throughout this text that the slots are rectangular
in shape, this does not have to be the case. The corrugations can be some other
shape, as long as they are consistently the same throughout the complete length
Manufacture and testing of corrugated horns 169

of the horn. Al-Hariri [4] investigated the effect of different slot shapes by
modelling the slot with a staircase representation of rectangular shaped slots. He
then used a radial transmission line approach to derive the equivalent admittance
looking into the slots. The results for various slots shapes are shown in Fig. 6.6.

\
>
> > > /
0-3
8 1 2 3 4 5
c / y y
3 / / / / /
E
§ o i
0-25 /
/
/
/
/
/

0
1
' 3
/
/
/ i
y/
/ y
S 0-4
D
0; / S/A /
to / /
a /
E / / /
5-0-3

Fig. 6.6 Normalised admittance against normalised slot depth for shaped slots
1 Square 2 Arcs of circles plus flats 3 Arcs of circles 4 Cosine 5 Triangular

The assumption is made of a large number of slots per wavelength. This indicates
that any of the shapes shown will give a zero normalised admittance at r = rt
provided the depth of the slot is increased. Triangular slots need to be nearly 0-4 X
deep and arcs-of-circle slots (curve 3) 0*32 X deep. This behaviour is expected from
an intuitive assessment of the electrical characteristics. The equivalent inner diameter
is also increased by the shaped slots. We have measured the radiation characteristics
of a length of open-ended corrugated waveguide (1*7 X diameter) with approximately
cosine shaped slots. Apart from the shift in operating frequency to a higher value
compared to the rectangular slot waveguide, the radiation characteristics were
undistorted.
The effect of bevels and corners, Fig. 6.5(c), on the edges of the corrugations
can also be assessed by the above technique. We have found that the effective slot
depth does not change significantly as long as the ratio sjs is greater then 0*8.
Again the bevels should be on all corrugations along the horn.

6.2 Testing of corrugated horns

6.2.1 Introduction
The radiation characteristics of corrugated horns need to be measured and compared
to the theoretical predictions or the design specifications. In principle the testing
of corrugated horns is no different from the testing of any other type of horn.
However, the design specifications are often more stringent, particularly with
regard to the crosspolarisation levels, than for other types of horn.
170 Manufacture and testing of corrugated horns

Most corrugated horns are used as feeds for reflectors, so that there are in
principle two sets of measurements to be performed. Measurements on the feed
horn and measurements on the reflector plus feed horn. However, it is not always
practical to measure the characteristics of the reflector plus feed horn due to the
large electrical size of the reflector and hence the large test range required for
measurement, thus measurement of the feed horn alone is then the only measure-
ment which can be undertaken. This is, in principle, adequate since modern test
ranges and techniques enable the feed to be accurately measured and computer
prediction can be used to give a good estimate of the total antenna radiation
pattern.
The general procedure for testing antennas can be found in IEEE Standard
Test Procedures for Antennas [184] and Appel-Hansen, Chapter 8 Rudge et al
[191]. The radiation characteristics of any test antenna are found by placing
the antenna in a region of pseudo plane electromagnetic waves. The antenna is
rotated and the power received by the antenna for specified orientations of the
test antenna with respect to the plane waves are recorded, Fig. 6.7. This assumes
that the test antenna is being used in a receive mode, the usual situation, but there
is no reason why it should not be used as a transmitter if that is more convenient.

Fig. 6.7 Antenna measurement system

The recorded patterns will differ from the ideal pattern because the wavefront
in the test region will not be truly 'plane', i.e. constant in amplitude and phase.
The 'planeness' of the wavefront, for a given test antenna, is a measure of the
performance of the test range. The nature of distortion away from a plane wave
depends partly on the type of test range being used and partly on the quality of
the test range. This latter factor is usually a function of the size and hence cost
of the range. Because a test range can be very expensive to construct there is a
natural desire to make a range which is only just adequate for the intended
Manufacture and testing of corrugated horns 171

measurements. This is fine until more stringent specifications are required, in which
case the performance of the range needs to be reassessed. Ideally an assessment of
range performance would be carried out for each test antenna. This is impractical
but the relationship of the test antenna to the test range should always be in the
mind of the antenna engineer. Fortunately most corrugated horns are relatively
small in size and do not require large test ranges.

6.2.2 Types of test range


There are, in principle, four types of test range which could be used for measuring
horns:
outdoor far-field range;
indoor far-field range (anechoic chamber);
compact antenna range;
near-field or intermediate-field range.
The far-field range (outdoors or indoors) creates a pseudo plane wave region by
utilising a small portion of a pherical wavefront, Fig. 6.8(a). If the test aperture
diameter is small by comparison with the length of the range, the amplitude and
phase of the electric field perpendicular to the axis of the range will taper only
gradually from the axis. Usually the minimum length of a far-field range is taken
as 2D2/X where D is the test aperture diameter, though for high quality horn
testing, a length of, say 4D2/\ is preferable.

Fig. 6.8 Types of test range


a Far-field range
b Compact antenna range
The outdoor far-field range is needed for testing large antennas. One, or both,
of the source and test antennas are mounted on towers spaced some distance
apart. It has many disadvantages. The performance of the range is difficult to
evaluate, operation is susceptible to weather and other environmental factors, the
land utilisation is high and it lacks security. However, for a large antenna there is
usually no other viable alternative. Since most corrugated horns are relatively
small in wavelengths, an outdoor range is neither necessary, nor recommended,
for horn testing. The indoor far-field range or anechoic chamber is, by contrast,
ideally suited for horn testing Measurement of the crosspolar radiation pattern
requires a low inherent crosspolar level in the range and this can be achieved in an
anechoic chamber. It also requires precise alignment of the test antenna with
respect to the source antenna and this is more easily achieved in an indoor
environment. The design of a suitable anechoic chamber is described below.
/ 72 Manufacture and testing of corrugated h orns

The Compact Antenna Range generates a pseudo plane wave region by collimating
the radiation from a point source with a parabolic reflector, Fig. 6.8(b). A portion
of the collimated beam will be a plane wave with superimposed small amplitude
ripples caused by scattering from the edges of the reflector. Both high quality,
high cost Compact ranges and medium quality, low cost Compact ranges, have
been constructed and widely used. Antennas up to about 2 m in diameter can be
measured in a compact, indoor environment. There is, however, a severe dis-
advantage as far as the measurement of corrugated horns is concerned. The offset
paraboloidal reflector generates a crosspolar component and the Compact range
reflector feeds have a relatively small aperture diameter. This means that their
inherent crosspolarisation is high. These two factors mean that the range cross-
polarisation is of the order of — 30 dB, far too high to enable low crosspolar
corrugated feeds to be measured. A dual offset reflector can offer a solution.
In near-field ranges the test antenna is usually the source antenna. The radiated
field of the test antenna is sampled in its near field region by scanning a probe
over a surface. The surface may be plane, cylindrical or spherical. The amplitude
and phase of the radiated field is recorded, digitised and processed with algorithms
to yield the far-field pattern. It is a very accurate method of antenna pattern
measurement, particularly suited to cases where the antenna is mounted on bigger
structures such as a satellite. It has not been widely used for feed horn testing,
to date, but there is no reason in principle why the patterns of corrugated horns
should not be measured by this technique.

6.2.3 Anechoic chamber design for low crosspolarisation horn measurement


An anechoic chamber should be designed to give a good plane wave region which
will be large enough to cover the horn in all orientations and at all operating
frequencies. It should have a level of background radiation, due to scattering off
the walls of the chamber, which is at least 10 dB below the minimum pattern level
to be measured. Thus, if the minimum copolar and crosspolar levels to be recorded
are — 40 dB then the chamber should have a background radiation level of — 50 dB
for its crosspolar scattering. The background radiation level in the chamber is
measured by the reflectivity level, of which more later.
The main design factor is the length of the chamber which will determine the
amplitude and phase taper across the test antenna. A range length which just
satisfies the far-field criteria of 2D2/X gives a phase difference of 22^° between the
centre and edge of the test aperture. The observed recorded effect on a radiation
pattern with theoretically deep nulls is to fill in the nulls. For a uniformly illumi-
nated circular test aperture the minimum value of the first null of the copolar
pattern will be — 27 dB instead of — °°dB and the first sidelobe level is increased by
0*2 dB. The amplitude taper is determined by the aperture size of the source horn.
For a 2D2/X range, the angle subtended at the source by the test antenna is
Manufacture and testing of corrugated horns 173

With D = 8 A this gives 6 = 3-6°. Reference to the pattern of a horn (for instance,
Fig. 5.4) shows that a 4X diameter source horn will give an amplitude taper of 0-6
dB at the edge of the test aperture. This level of amplitude taper will increase the
observed si delobe level by about 0-2 dB and slightly increase the observed copolar
beamwidth. A smaller source horn will generate a wider beam and hence a smaller
amplitude taper, however, it will also increase the amount of energy incident on
the walls of the chamber. Taken together the result of this discussion on the
amplitude and phase taper is to indicate that a range length of greater than 2D2/X
is preferable for high quality horn testing. Doubling the range length to R = 4D2/X
gives a test range which is capable of measuring deep nulls in radiation patterns
and distorts the pattern shape by a negligible amount if a source horn of 4 A
diameter is used.
The shape of the anechoic chamber can be of two main types, either a rectangular
or funnel shape. The latter tapers from one end to reach a point at the other end.
The amount of absorber needed on the walls is reduced by more than 50 per cent
over the same length rectangular box. But its design has disadvantages. Most
materials are relatively poor absorbers at high angles of incidence and reflect a
proportion of the incident energy. The walls of the funnel chamber cause it to act
similar to a pyramidal horn and consequently the wave propagating from the
source at the apex is not a plane wave but a waveguide mode. This need not be
detrimental as long as the phase centre of the horn source antenna is at the phase
centre of the tapered chamber. Care in setting up and locating the source antenna
is required.

Fig. 6.9 Sketch of Queen Mary College anechoic chamber

The rectangular chamber has none of these technical problems. It is simple to


operate and set up. In addition it can be used for scattering and near-field probing
measurements.
A compromise between the two designs is possible which gives the advantages
of the rectangular chamber and some of the cost saving of the tapered chamber.
One half of the chamber, that used for the test region, can be rectangular in shape
an the other half can be tapered to a smaller cross-section but not a point. We have
built and successfully operated a chamber based on this compromise design. A
174 Manufacture and testing of corrugated horns

sketch of the chamber is shown in Fig. 6.9. It is intended principally for use be-
tween 8 and 18 GHz. The chamber has a cross-section of 3 m by 3 m and tapers
down to cross-section of 1 -2 m by 1 -2 m.
The cross-section at the quiet zone needs to be such that the test antenna is
electrically 'isolated' from the walls. Energy scattered off the test horn and the
mechanical support system should have a path loss of about 40 dB before reaching
the walls. The cross-section of the chamber is made square in order to preserve
symmetry. This is important for crosspolarisation measurements to ensure that
most reflections have a similar path loss.
The walls of the anechoic chamber are completely covered with radar absorbing
material. This should have a reflectivity at normal incidence of at least — 40 dB
for good crosspolar measurements. The absorber on the end wall has to absorb
all the incident energy which is not recieved by the test horn. The absorber on this
end wall should therefore have a better reflectivity than the side walls. In our
chamber the end wall and the square box section are covered with — 50 dB
reflectivity absorber (at X band) while the tapered section is covered with — 40 dB
reflectivity absorber.
The horn used for the source can be any linearly polarised horn which satisfies
the beam width criteria discussed above. It does not need to have a low cross-
polarisation as long as the crosspolarisation is zero on boresight. Reference to any
of the crosspolarisation patterns in the previous chapters will show that the cross-
polarisation within the small angular region illuminating the test aperture is very
low and much less than the level generated by reflection off the walls. For this
reason a pyramidal horn can be used as the source antenna even though its peak
crosspolar level is relatively high. A pyramidal horn generates a nearly perfect linear
polarised field in the boresight direction. Pyramidal horns are readily available and
so tend to be the preferred source horn for anechoic chambers.
The final component in the anechoic chamber is the turntable and mechanical
support for the antennas. These need to be rigid for crosspolarisation measurements,
but also have the smallest possible cross-section when viewed from the source
antenna. Supports made from plastic or fibreglass can reduce the reflected signal.
The complete pattern measurement of a corrugated horn requires two turntables;
an azimuth turntable and a polarisation rotating turntable, Fig. 6.7. It is not necessary
to have an elevation turntable but some form of elevation adjustment should be
provided for alignment purposes. Similarly, a convenient means of adjustment in
the vertical, horizontal and longitudinal directions will also aid alignment.
The correct alignment of the test horn with respect to the source antenna is
one of the most time-consuming aspects of low crosspolarisation measurements.
Copolar measurements are less critical and can usually be aligned by eye, but to
be able to measure crosspolarisation patterns at levels below — 30 dB requires
precise alignment. The axis of the test horn must be exactly coincident with the
axis of the source horn. Various optical instruments can be used but we have found
that a visible light laser is a very convenient method of alignment. The laser is set
up to shine through and along the axis of the source antenna waveguide and along
Manufacture and testing of corrugated horns 175

the axis of the chamber. The source antenna is first aligned with the laser using
temporary cross-wires as guides. Then the test antenna is carefully aligned by the
same method.
The instrumentation required for the measurement of the copolar and crosspolar
patterns is the same as any antenna pattern measurement, namely, stable source,
stable reciever, pattern recorder and associated control instrumentation. The use
of digital control and the immediate digitising of data for subsequent processing
and display are increasing both the flexibility and the accuracy of the measurement
sequence.

6.2.4 Assessment of test range performance


The basic performance of ananechoic chamber is determined by the energy scattered
off the walls, floor and ceiling of the chamber. This can be understood by
considering the situation shown in Fig. 6.10. The signal picked up by the test
antenna will be the sum of two components, the direct signal and the signal reflected
off the walls, i.e.
V = Aie^ + A2e]a* (6.2)
A i and A2 are the amplitudes of the radiation pattern of the test antenna in the
direct and reflected signal directions plus any path length loss differences. A2 also
includes the attenuation which the reflected wave has undergone at the walls of
the chamber and normally A2 <A^ OLI and a2 are the path phase lengths of the
respective directions plus any test pattern phase differences in the two directions.

Fig. 6.10 Spurious reflections from wall of anechoic chamber

Now if the test antenna is moved sideways, the path length OL2 will change
considerably compared to ax. The signal received at the test antenna will oscillate
between a maximum value (\AX | 4- \A2 |) and a minimum value (\AX \ — \A21). The
peak-to-peak deviations of the received signal will be:

\Al\-\A2\

In practice there will be a number of unwanted reflected components. This will not
affect the principle of the above and if the peak-to-peak deviation in the recieved
signal is measured as the test antenna is moved around the test area, then the total
/ 76 Manufacture and testing of corrugated horns
peak-to-peak deviations recorded will be given by:

t = 20 log! (6.4)
\At\-\y\
The quantity y is the reflectivity level of the test range for direct signal level A.
This equation is plotted in Fig. 6.11. Note that the reflectivity level, as here defined,
is a function of a particular test antenna since the value is partly determined by the
radiation pattern. A test antenna with a near omnidirectional pattern will receive
unwanted components from more directions simultaneously than a highly directional
antenna. It is therefore important to assess the chamber performance with a test
antenna which has similar radiation characteristics to the horns to be measured.

-70

-10
0-1 0-2 0-5 1-0 2-0 5-0 10-0
peak-to-peak deviation, dB

Fig. 6.11 Reflectivity graph for use with pattern comparison assessment. Parameter: pattern
level, dB

The reflectivity level can be measured in two ways, by either the VSWR method
or the Pattern Comparison method. Both methods are described in detail in Appel-
Hansen (Chapter 8 of Reference 191). In the VSWR method, the field in the test
region is recorded by moving a test antenna continuously across the test region.
The test antenna is mounted on a movable carriage and the method is similar to
sampling the VSWR in a transmission line using a moving probe. It is capable of
giving a detailed picture of the test region field but requires specialised equipment
which must be set up in the chamber. Also great care is needed to ensure that there
are no extra reflections from the movable carriage which must necessarily be large
to be rigid.
The pattern comparison assessment is easier to implement and requires only
standard antenna test equipment. The test antenna is mounted on a turntable and
its radiation pattern recorded. The turntable is physically moved to another
Manufacture and testing of corrugated horns 177

position in the chamber and the pattern recorded, the positions are chosen to
maximise the chance of observing maximum amplitude variations. All the recorded
pattern are superimposed on top of one another, Fig. 6.12, and at various signal
levels below the peak on a reference pattern, the maximum difference between
patterns is noted. The signal level is \AX\ and the maximum difference is t, hence
eqn. 6.4 can be used to compute the reflectivity level;;. A large number of values
are measured and an average value deduced for the reflectivity level. For cross-
polarisation measurements the procedure must be carried out twice, once for the
copolar reflectivity level and then again for the crosspolar reflectivity level. The
test antenna need not have low peak crosspolarisation but it should go to a low
level at some azimuthal angles.

angle
Fig. 6.12 Superimposition of patterns

Experience shows that the pattern comparison method tends to given a lower
reflectivity level than the VSWR method, probably because the test region field
is being sampled rather than continuously probed. However, it is a more convenient
method to implement and the reflectivity level can be assessed using the actual
corrugated horn for which accurate measured patterns are desired.

6.2.5 Measuremen t of copolar and crosspolar patterns


The co-ordinate system used for radiation pattern measurements is shown in
Fig. 6.13. The radiation pattern is always recorded by rotating the test antenna
in the 0 = 0° plane (xz plane). The patterns in various 0 planes are obtained by
rotating the source and test antennas about the z-axis so as to align the electric
field vector at the desired angle 0 with respect to the xz plane. This scheme means
that the boresight copolar signal level is constant for all planes. The measurement
procedure for the amplitude patterns of a corrugated horn with a relatively low
crosspolarisation is as follows:
(i) The corrugated horn to be measured is set up on the turntable and carefully
178 Manufacture and testing of corrugated horns

aligned so that the axis of the test horn and the axis of the source horn are
coincident. If the horn pattern has a crosspolar null in the boresight direction (as
is normally the case), this fact can be used as an aid to alignment. The test horn
should be placed such that its phase centre is over the centre of the turntable, then
rotation of the test horn will keep the electrical path lengths from the source to
the test horn constant at different azimuth angles. In practice, the centre is not
uaually known precisely and so a guess has to be made. If only far-field patterns are
required then the precise location of the test horn with respect to the turntable is
not important. If phase patterns or near-field amplitude patterns are required then
the electrical phase centre must be found, see below. Note that, if the phase centre
of the horn is not at the aperture of the horn, then the effective test aperture, in
thexz plane will be larger than the horn aperture.

Fig. 6.13 Co-ordinate system for feed measurement

(ii) The H-plane copolar pattern is recorded by orientating the electric field
vector of both the source and test antennas so that they are parallel with the
j-axis. Then rotate the test antenna in the xz plane and record the pattern over
the desired angular range of 0. The gain controls of the receiver should be set
up to give an undistorted output with enough dynamic range for the measurements.
The variable gain controls should not be adjusted after this initial set-up. All
recorded signal levels will then be with respect to the peak copolar level. Any
change in gain should be done using calibrated switched attenuators,
(iii) The H-plane crosspolar pattern is next recorded by rotating the test horn
through 90° in 0 so that its electric field vector is now parallel with the x-axis.
The source antenna is unchanged. The test antenna is again rotated in the xz
plane to record the crosspolar pattern. Of course this pattern level should be
zero, so that only the inherent reflectivity level of the chamber will be recorded.
In practice, it may be difficult to find the H plane copolar orientation in (ii) pre-
cisely, so use can be made of the knowledge of zero crosspolarisation in the H
plane by reversing (ii) and (iii). The H plane crosspolar orientation is found by
adjusting the test horn 0 angle until no signal is received. This then defines the
direction of the electric field vector in the test horn. The test horn can now be
rotated through precisely 90° in 0 to record the copolar pattern. This scheme
Manufacture and testing of corrugated horns 179

implies that there is an accurate method of measuring the angle of rotation of


the test horn about the ^-axis, a necessity for good crosspolar measurements.
(iv) The E plane copolar pattern is recorded by orientating the electric field
vector of both source and test horns so that they are parallel with the x-axis, and
again rotating the test horn in the xz plane.
(v) The E plane crosspolar pattern (which should be zero for single mode horns
with unity azimuthal dependence) is found similarly to (iii) above,
(vi) The 45° plane copolar pattern is recorded by orientating the electric field
vector of both source and test horn so that they are parallel to the 0 = 45° or 135°
direction. The pattern is again recorded by rotating the test horn in the xz plane.
This case is shown in Fig. 6.14(a).

Fig. 6.14 Test and source orientation for 45 degree plane measurements
a Copolar
b Crosspolar

(vii) The 45° crosspolar pattern is recorded by turning the test horn through
precisely 90° in 0, leaving the source horn unchanged, and repeating the 0 pattern
rotation. The situation is shown in Fig. 6.14(b). This measurement is one of the
most important recordings for single mode horns with unity azimuthal dependence
because the 45° crosspolar pattern should give the peak crosspolar level. The
pattern levels recorded are with respect to the copolar peak level.
180 Manufacture and testing of corrugated horns

(viii) The patterns in any other plane can be similarly recorded by repeating the
above procedures. A complete contour plot requires that the patterns are measured
in enough planes to give a smooth record of the horn radiation characteristics.

6.2.6 Measurement of phase patterns and gain


The phase patterns need some form of phase-amplitude receiver. Usually the
purpose of the phase measurements is to locate the phase centre of the horn and
to see how nearly spherical is the copolar main beam. High sensitivity is not there-
fore needed, and a network analyser can be used. The shape of the recorded phase
patterns will depend on the location of the test horn phase centre with respect to
the axis of the azimuth turntable. Thus it is first necessary to locate the phase
centre, as discussed in Section 5.4. The phase centre changes with azimuth angle
0, so it is generally assumed that the horn phase centre means the phase centre
along the boresight direction. If the crosspolarisation is significant, the phase centre
may be different in the E and H planes.
The phase centre can be located by recording the phase pattern over a narrow
angular range about the boresight direction, then either move the horn along the
z-axis and repeat the phase pattern measurement until the phase is constant, or
use the recorded pattern and geometry to locate the phase centre.
After location of the phase centre in a particular plane the phase pattern in that
plane can be recorded, indicating the way in which the horn radiation pattern
deviates from a spherical wave.
The gain of a corrugated horn can be measured by any of the standard methods
described in IEEE Standard Test Procedures for Antennas [184], and Rudge etal
[191]. The gain transfer method where the gain of the test horn is compared to
the gain of a standard horn is easy to implement in an anechoic chamber. The gain
of the standard horn should be similar to the gain of the test horn in order that
the extraneous signals in the chamber affect both horns similarly. The gain needs
to be known at all specified frequencies, otherwise another method such as the
three antenna method must be used. In this case three horns are measured in all
combinations and the free-space path loss formulae used to compute the gain of
the test horn.
Chapter 7

Rectangular and elliptical


corrugated horns

7.1 Introduction

Corrugated horns and feeds with non-circular cross-sections have received relatively
little attention in the past. Out of nearly 180 papers cited in the bibliography
only a few have dealt with horns of rectangular or elliptical cross-section. There
are good reasons for this bias, for it turns out that unless a shaped beam is required,
circular corrugated horns are superior in almost all respects to non-circular corru-
gated horns. Furthermore, rectangular and elliptical corrugated horns are difficult
to analyse. The simple theoretical models are only an approximation to the true
electromagnetic behaviour of the horns and do not give good predictions of cross-
polar performance. The electromagnetic fields in the horns are more complicated
and the fields in the slots are dispersive. The throat region is difficult to design
so as to suppress unwanted mode excitation. Finally, the horns are difficult to
manufacture and thereby expensive to produce. But, as stated, their one significant
advantage compared to circular corrugated horns in certain applications, is their
ability to generate a beam with a non-circular symmetry. This is desired in some
spacecraft antennas and also some radar antennas. In this chapter we shall discuss
their properties but the approach will be limited to a study of the basic character-
istics, without the detail which has been devoted to circular corrugated horns.

7.2 Rectangular corrugated horns

7.2.1 Background
The first paper on rectangular corrugated horns appears to be that of Lawrie and
Peters [92] who introduced corrugations onto the E plane walls as a way of
tapering the aperture electric field to reduce the E plane sidelobes. The two-walled
rectangular corrugated horn can be analysed exactly, and a theory has been given
by Baldwin and Mclnnes [11]. A number of authors have subsequently reported
applications of the two-wall corrugated horn [50, 96, 113, 135, 142]. The first
analysis of the four-walled rectangular corrugated horn with corrugations on all
182 Rectangular and elliptical corrugated horns

four walls was presented by Bryant [20] who used it as a horn for producing
circularly polarised patterns with low sidelobes [19]. The four-walled corrugated
horn cannot be analysed exactly using model analysis because the boundary
conditions cannot be made coincident with the reference direction of a rectangular
co-ordinate system. Bryant [20] used a superposition technique and treated the
four-walled waveguide as the sum of two two-walled corrugated waveguides, an
approximate procedure which ignores the influence of the corners of the slots. The
theory presented by Bryant was not rigorous and started a controversy. Dybdal et
al. [57] claimed that it was incorrect and subsequently [58] published an
impedance compatibility relationship. This showed that the four-walled corrugated
horn did not satisfy the relationship and so they deduced that it could not carry
any power. However, they neglected the evanescent modes in the slots and the
theory was shown by Isaac [73], Narasimhan [112] and Bach Anderson [8] to be
incorrect. Since then no major advances in the theory have been reported and the
best method of modal analysis is still that based on Bryant's superposition method
with an additional orthogonal mode set.
A number of papers have reported applications of rectangular corrugated horns.
The ability to shape the radiation pattern with a four-walled corrugated horn having
the corners filled in was reported by Manwarren and Farrar [96]. Its use as a feed
horn was reported by Baldwin and Mclnnes [14] and Davis [50]. The attenuation
properties of two-walled corrugated waveguides were studied by Al-Hariri et al [3,
4] and Baldwin and Mclnnes [10]. A four-walled corrugated horn was built and
extensively measured in order to investigate its copolar and crosspolar radiation
characteristics by Adatia et al. [2]. They found that it was necessary to include an
EH set of modes, in addition to the HE modes to account for the crosspolar levels,
and that higher order modes deteriorated the performance. The throat region of the
horn was particularly difficult to design correctly, partly due to the overmoded
nature of the waveguide and partly due to the lack of a reliable theory.
In the following analysis we shall start by studying the propagation characteristics
of two-wall rectangular corrugated waveguides. The four-walled corrugated wave-
guide and horn can then be analysed by superimposing an orthogonal mode set. The
corrugated walls are treated as an anisotropic surface impedance. The impedance
boundary conditions lead to hybrid modes in the waveguide, as in the circular
corrugated waveguide. The hybrid modes have fields with only one of the transverse
electric or magnetic components zero. The two mode sets use the x direction as the
reference and have either Hx = 0 or Ex = 0. The former gives HE modes (or TE^.
modes) and the latter EH modes (or TMX modes). The TM^/TE^ designation has
been used by most workers but because it leads to confusion with the smooth-wall
rectangular waveguide where the mode designation is referred to the z direction of
propagation, we will refer to the modes as HE nm and EH n m . The subscript n refers
to the number of variations in the x direction and the subscript m refers to the
numerical order of modes, of type n, as frequency increases from zero. Modes of
order n can have transverse electric fields which are either symmetrical or asym-
metrical with respect to central axis. This leads to four sets of field equations (i.e.
Rectangular and elliptical corrugated horns 183

symmetrical x, symmetricaly\ asymmetrical x, asymmetrical^). Consequently for


each mode set (HE or EH) there are two characteristic equations and two sets of
cut-off frequencies. The mode structure is thus considerably more complicated than
circular waveguides, though because the fields are represented by trigonometric
functions rather than Bessel function the characteristic equations are quicker to
compute.

7.2.2 Two-walled rectangular corrugated waveguide


The analysis of the two-wall corrugated waveguide shows most of the features of
the more general four-wall waveguide so will be studied here. A detailed analysis
would require that the fields in the central region are represented by a fundamental
wave and a set of space harmonics and the fields in the slots by an infinite sum of
standing waves, most of which are evanescent. However, we shall assume that the
space harmonics are zero and that fields in the slots are due only to the lowest
order TE10 mode. It will be assumed that there are a large number of slots per wave-
length and that the ridges are infinitely thin so that the longitudinal geometry is not
a parameter in the analysis. The relevant geometry of the waveguide is shown in
Fig. 7.1. At the surface of the corrugations,^ = ±b/2,Ex=0 andE y ^ 0 . Thus the
boundary conditions will be satisfied by modes which are TE to x or HE modes.

i ky

ts
b/2
T

2 ^
-b/2

-a/2 a/2

Fig. 7.1 Two-walled rectangular corrugated waveguide

The waveguide may be conveniently analysed by starting with a Hertzian magnetic


vector potential (time dependence e ja; * assumed).
*a* (7.1)
E = (7.2)
2
H - k FH + 7
H) (7.3)
= 0 (7.4)
where k = 2TT/X0 and K% = k2 —j32. The solution to the wave eqn. 7.4 is
i//H = [A cos (Kxx) + B sin (Kxx)] [C cos (Kyy) + D sin (Kyy)] e'ipz
(7.5)
where K2C=K2X+Kl.
184 Rectangular and elliptical corrugated horns

Applying the boundary conditions that — 0 at x = ±#/2, leads to Kx =


nn/a and four possible solutions:

=
H, An cos I — x cos (Kyy) n, odd (7.6)

= Bn cos |—x) sin(Kyy) ft, odd (7.7)

=
3 Cw sin( — x| cos (Kyy) /t,even (7.8)

Dn sin (— x) sin (Kyy) n, even (7.9)

^ H , s ^ H a r e symmetric in y and \^H2> ^ H 4 a r e asymmetric in.y. ^ H 1 ? I / / H 2 a r e


symmetric in x and i//H , ^ H 4 are asymmetric in j . The field components are
obtained from eqns. 7.2 and 7.3. For 0 H l (symmetric, symmetric) these are:
Ex = 0

cos

= — ]<joixKyAn cos |—-xj %ix\{Kyy)

(7.10)
in cos j — x | cos (Kyy)

Hy = — ^ y ^ 4 n sin j — xj si

H2 = \ ~ &4n sin | — JC| COS(A V J)


a \a

where |8? = k2 -1 ~ • Hence Kl = fx - ]32

The impedance parallel to the surface of the corrugations at y = ± b/2 is

(7.11)

In the slots we assume a cut-off TE10 mode, so Ky = j3j and from eqn. 7.10 at
Rectangular and elliptical corrugated horns 185
= ±b/2:
Ex = — Hz — 0

It
Ez = — —x
a

Hx = -faE^osl-x] cos (fas) (7.12)

Hy = —Exsm\—x\ si

14
HE 2 0 / / /
HE3O / / /
1 2 --

\y/7
10-

6 - /

HEio/ /HE,.
2-

i / 1" 12 16

Fig. 7.2 Dispersion curves for HE modes in rectangular corrugated waveguide: b/a = 7, s/a =
0-2, BH — 'balanced hybrid' frequency

The surface impedance is

Zs = (7.13)

Note that Zs is a function of s, a and frequency. It is therefore dispersive, in


contrast to circular corrugated waveguides where the surface impedance is non-
dispersive. The purpose of the corrugations is to create a 'magnetic wall' a t j = ±bj
2, i.e. Hx = 0 or fas = m + n/2 or s/X = 0-25y/{\ + 4(s/a)2 }. The slot depth
required to create this 'balanced hybrid' condition is thus dependent on the size of
the main waveguide.
186 Rectangular and elliptical corrugated horns

16

Fig. 7.3 Cut-off characteristics for HE modes in rectangular corrugated waveguide


a s/a = 0-05
b HE12 mode, parameter s/a
Rectangular and elliptical corrugated horns 187
Equating eqns. 7.11 and 7.13 gives the characteristic equation

Ky tan I — I = - f t tan (/M) (7.14)

This applies for all n and \I/ symmetric in y. For \j/ asymmetric in >>, the charac-
teristic equation is

\= + ft tan CM) (7.15)

If /} is greater than px, Ky becomes imaginary and eqn. 7.14 becomes

^ H = + ft tan (jM) (7.16)

This equation gives slow waves bound to the corrugations. For computational
purposes the equations are conveniently normalised to give for eqn. 7.14

(Kya) tan hr-J = (ftaJtanOM) (7.17)

Where (Kyaf = (fixaf - (pa)2 and (Pxa)2 = (ka)2 -(mrf. Also Kyb = (Kya)
b/a and /M = (Pxa)s/a. This shows that the characteristics can be normalised to a
and that the controlling geometric parameters are b/a and s/a. In practice the slot
depth will be determined by the balanced-hybrid condition and consideration of
the usual aperture sizes for horns shows that as a general guideline 0*04 < s/a < 0*2.
Propagation characteristics for b/a - 1-0 and s/a = 0-2 are shown in Fig. 7.2. As
stated earlier, the mode nomenclature is such that HE l m modes are numbered
consecutively from the lowest frequency. This is convenient but does mean that the
field dependence is not immediately clear. Thus the HE10 mode is the symmetric
slow wave solution of eqn. 7.16. The HE n mode is the lowest order asymmetric
solution of eqn. 7.15. The HE12 mode is the lowest order solution of eqn. 7.14.
This mode is symmetric in both* and^ and has a single cosine distribution in both
planes. It is the mode with the first balanced-hybrid condition and is the mode
required for low crosspolar radiation characteristics. The designation of the other
modes is similar. There is no HE21 mode, corresponding to the HE n solution.
A number of features of the HE12 mode are evident from Fig. 7.2 It is not the
dominant mode in the waveguide and its solution at cut-off corresponds to a TE12/
TM12 mode in a smooth-wall rectangular waveguide. This implies that excitation of
the mode in the throat region of a horn will not be as easy as in a circular corrugated
horn. Also the lower frequency of operation should be chosen so that the HE10/
HE n modes have gone to their high frequency cut-off. The region of operation is
overmoded. This is for a small waveguide size. As the size becomes larger the over-
188 Rectangular and elliptical corrugated horns

moding increases considerably, and thus the possibility of higher order mode
excitation becomes more likely.
The special points on the propagation characteristics can be identified by
substitution into the characteristic equations. The cut-off frequencies (]3 = 0) are
given for modes symmetric in.y by:
P
{kaf = / \ 2 +(nn)2 (7.18)

\a 2a)
and for modes asymmetric in >> by:

(7.19)

where p = 0 , 1 , 2,. . ., andg = 1 , 2 , . . . .


These are plotted in Fig. 7.3(a) and (b) for typical parameters. The higher
frequency cut-off (/? -> °°) for all modes is given by:

(ita)3 = (>
"(,/a)'V +(nn)2 (7 20)
-
The balanced hybrid frequency occurs when fixs = TT/2 for the HE12 mode. This
leads to the same condition as j3->°° and hence eqn. 7.20 also gives the balanced-
hybrid frequency. It is a function only of s/a so as the waveguide size increases
(implying smaller s/a) the balanced-hybrid frequency becomes correspondingly
larger.
One potentially useful feature of the rectangular corrugated horn is that the
balanced-hybrid conditions are maintained over a wide frequency range. This is
shown in Fig. 7.4 where the wavenumber Kyb is plotted against normalised fre-
quency for the HE12 mode. The balanced-hybrid condition occurs when Kyb = n.
It is seen that Kyb is nearly constant over an octave of bandwidth for s/a = 0*1. For
a smaller s/a the bandwidth is even greater. However, the need to avoid the
propagating band of the HE U mode means that the useful frequency range is some-
what reduced.

1.23 Four-wall rectangular corrugated waveguide


The rectangular corrugated waveguide, shown in Fig. 7,5, with corrugations on
all four walls is not amenable to exact analysis. The method used to obtain an
approximate solution is to treat the waveguide as the superposition of two two-
wall corrugated waveguides. With the principal electric field vector polarised along
the y, or vertical axes, the electric fields of the HE type modes will see the vertical
corrugations (containing a large number of corrugations per wavelength) as plane
electric waveguide walls. Thus one of the two-wall corrugated waveguides is identical
with the waveguide studied in the last section and all the results given there apply
Rectangular and elliptical corrugated horns 189

to HE modes in four-wall corrugated waveguides. The other two-wall corrugated


waveguide has its corrugations on the side walls. Examination of the fields of the
EH mode set to be supported by this waveguide, indicates that the top and bottom
walls, at y = ± 6/2, should be treated as perfect magnetic conductors, while the
impedance of the side walls at x = ±a/2 will be due to the presence of a TE10
mode in the corrugations.

"0 10 20 30

Fig. 7.4 Wavenumber Kyb against ka. Parameter s/a

y i1
1s
L t
b
X

*L a

Fig. 7.5 Four-walled rectangular corrugated waveguide

The analysis is similar to the previous section, except that the boundary
condition for eqn. 7.5 is i//H = 0 at y = ± b/2. This leads to an impedance at x = ±
a/2 for modes symmetric in x of:

z = EA cot \KX — (7.21)


190 Rectangular and elliptical corrugated horns
where $\ = k2 — (nn/b)2. The surface impedance due to the slots is:

(7.22)

The characteristic equation follows as:

Kr cot tan (7.23)

Comparison with the characteristic equations of the HE modes shows that this
has the same form as HE modes which are asymmetric in y, eqn. 7.15. EH modes
with fields symmetric in x have a characteristic equation similar to eqn. 7.14. Thus
in a square corrugated waveguide with constant slot depth on all walls the HE and
EH modes will be degenerate. The propagation characteristics shown in Fig. 7.2
apply to EH modes with appropriate change of parameters.
The degeneracy of the HE and EH modes in a square corrugated waveguide is
a serious disadvantage. Adatia et al. [2] have found that the EH modes are easily
excited and contribute to the crosspolar radiation from a horn. Fortunately there
is little reason to choose to use a square corrugated waveguide since a circular
corrugated waveguide outperforms it in almost all respects. The main reason for
considering a rectangular corrugated waveguide is to obtain different beamwidths
with the electric vector polarised along the x andy axes. For a ¥= b, the degeneracy
disappears and the main practical effect of the EH modes is to increased the
number of modes which can potentially propagate in the waveguide. Since the
field patterns of the EH modes are similar to the HE modes there is a strong
possibility of the modes coupling together.

7.2.4 Radiation characteristics of rectangular corrugated horns


The radiation patterns of a rectangular corrugated horn with narrow flare angle
can be calculated using the Fourier Transform, or E-field, technique [132]. The
far-field is:

F(x,y) = j j (Eyzx —Ex2Ly) exp [}k(x sin 0 cos 0 -¥y sin 6 sin $)] dxdy

(7.24)
where Ex and Ey are the aperture electric fields.
For the HE modes Ex=0 and Ey is given by eqn. (7.10). Substituting into
eqn. 7.24 and evaluating the integrals leads to the following expression for HE
modes which are symmetric in y.

ntx nn\ Ika


— sin i — I cos I— sin 6 cos <
Fx = -
— i — (k sin 6 cos (j))2
Rectangular and elliptical corrugated horns 191
1

(kb \ yb\
Ky sin cos sin 9 sin 01 K sin 9 sin <b
j cos COS sin t1 sin (
\2 2
/

[K l-(k sin 9 sin 0 ) 2 ]

(7.25)
The copolar and crosspolar components of the far-field are given by

Eco = (cos2 ( - - s i n 2 - I cos 20| Fx

Ecx = sin2 l y i sin20F x (7.26)

The far-field expressions for the asymmetric HE modes and EH modes can be
similarly derived.
Since we are mainly interested in the behaviour around the balanced hybrid
frequency of the HE12 mode, where Kyb = n, eqn. 7.25 can be simplified to:

.ka kb
cos I — sin 6 cos 0 cos I — sin 9 sin (j>)
(kb)(fid) (7.27)
_ 7i2 — (ka sin 9 cos 0) 2 7T2 - (kb sin 9 sin 0) 2

Examination of this equation shows that the first square bracket controls the H
plane radiation pattern and the second square bracket controls the E plane radiation
pattern. The two planes have identical forms so that a square horn will have a
symmetric radiation pattern (ignoring the cos 6 factor in eqn. 7.26). This, of
course, was the aim of the corrugations on the top and bottom walls. In a plain
wall waveguide the E plane pattern characteristic is different and gives rise to an
asymmetric pattern with very high sidelobes, as sketched on Fig. 7.6. This figure
shows a universal pattern for the HE12 mode in a corrugated waveguide of side
a operating near the balanced hybrid frequency. The pattern is relatively broad
with a first sidelobe level of — 24 dB. As expected, the efficiency of radiation from
the physical aperture is low. Rectangular corrugated horns are bulky and possess
a wide flange to encompass the corrugations which are in general much deeper
than in their circular counterpart. Fig. 7.6 can be used to assess the patterns of a
non-square horn by appropriate scaling of the factor a. From eqn. 7.27, the — 3
dB half beam width in either the E or H planes is:
to sin 0 = 3-74 (7.28)
Thus the sine of the beam width is inversely proportional to the normalised height
or the normalised width. This equation can be used as a design aid to obtain the
required aperture dimensions. The equivalent equation for the —10 dB half beam-
width is ka sin 6 =6-4.
192 Rectangular and elliptical corrugated horns

Crosspolar results have not been computed using eqn. 7.27 because it is known
[2] that the theoretical predictions considerably underestimate the level of cross-
polarisation present in a practical horn. This is believed to be due to the presence
of the EH modes. The accurate prediction of the crosspolar characteristics awaits
a more sophisticated theory. Experimental models have given relatively high levels
of crosspolarisation when compared to equivalent size circular horns, though
some of this high level was due to higher order mode excitation. Note that the
rectangular horn does not have a crosspolar characteristic with a bandpass type of
frequency characteristic. This is in contrast to the circular corrugated horn where
the crosspolar level theoretically goes to zero at one combination of radius,
frequency and slot depth. In the rectangular horn case the crosspolar level decreases
monotonically with frequency.

12

Fig. 7.6 Radiation patterns for HEX1 mode in rectangular corrugated waveguide
H plane and E plane (Ignoring cos d factor)
E plane in smooth-wall rectangular waveguide

The above discussion has been confined to open-ended rectangular corrugated


waveguides. The effect of the flare which is present on a horn is similar to the
circular case (see Section 4.7 and 5.2). As in the circular case, after the spherical
phase error has exceeded half a wavelength, the copolar patterns tend towards a
constant beamwidth and for large flare angles the beamwidth is determined prin-
cipally by the flare angle and not the aperture size.
Reference has already been made to the need for particular care in designing
the throat section of a rectangular corrugated horn. The principle of design is
similar to the circular case, namely, the first slot should be half a guidewavelength
deep. The fact that the HE12 mode has a TE12/TM12 type field at cut-off and is not
Rectangular and elliptical corrugated horns 193

the dominant mode in the rectangular waveguide poses additional problems. The
half guide wavelength slot should be chosen at the highest operating frequency so
that it is never deeper than Xg/2. The slot depths can then be tapered from Xg/2
to Xg/4 over a few wavelengths. The throat size should be as large as feasible
because the modes propagating in the deep slots are very dispersive (in contrast to
the circular horn). This means that the VSWR characteristics are also highly frequency
dependent, although the large throat size means that the absolute level can be kept
low. The maximum size of the throat is determined by the need to avoid higher
order mode excitation. Unfortunately, the waveguide fields of the next HE/EH
modes above the HE12 mode have similar patterns to the HE12 mode. This means
that higher order modes can be easily excited. Add to this difficulty the approximate
nature of the theory and the difficulties of manufacturing rectangular corrugated
horns and it is easy to see why they have not attracted much interest in the past.
The only reason to choose a rectangular structure rather than a circular structure
is the ability to generate elliptical radiation patterns. For all other criteria the
circular horn is superior.

7.3 Elliptical corrugated horns


Elliptical corrugated waveguides have recieved relatively little attention to date.
This is partly due to the difficulties of analysis and partly due to the problems
of manufacturing an elliptical corrugated horn. However, the need to produce
elliptically contoured beams with high polarisation purity is increasing the interest
in these horns. Because this type of horn is still being studied we will confine
ourselves to stating the fundamentals of the analysis of elliptical corrugated
waveguides.
The first work on elliptical corrugated horns appears to have been done by
Jeuken et al. [78, 83] who used a simple analytical model and also built a horn
with good circular polarisation characteristics. The horn was made by machining.
The crosspolar performance of a horn made by a series of elliptical discs was
reported by Guy and Ashton [67]. They found that the peak crosspolarisation
when the electric field was parallel to the minor axis of the ellipse was below
— 35dB over a 35% band but when the electric field was parallel to the major
axis of the ellipse this band was reduced to only a few percent. A detailed analysis
of the elliptical corrugated waveguide has been given by Al-Hariri [4] who studied
the low attenuation properties of the waveguide.
A sketch of the elliptical corrugated waveguide is shown in Fig. 7.7(a). The
elliptic cylinder co-ordinate system (£, ??, z) is shown in Fig. 7.7(b), where the
contour surfaces of constant £ are confocal elliptic cylinders, and those of constant
r) are confocal hyperbolic cylinders. The constant h represents half the distance
between the foci. The confocal cylinder £ = £0 coincides with the boundary of the
waveguide, and the z-axis coincides with the longitudinal axis of the elliptic wave-
guide. The eccentricity e of the cross-section is given by 1/cosh £0- The major and
minor axes are 2a = 2h cosh if0 and 2b = 2h sinh %0, respectively.
194 Rectangular and elliptical corrugated horns

The wave equation in the elliptic cylinder system has the form:
32A 32A
~^2~ T T 2 ^ ( c o s h 2£ - cos 2T?) A = 0
+ (7.29)

where 4# = K2h2 and AT2 = co2

T)=180'

£=2
270°

Fig. 7.7 Elliptical corrugated waveguide


a Waveguide
b Elliptical co-ordinate system consisting of confocal ellipses and hyperbolas. P has
co-ordinates £ — 2, 77 = TT/3
c Two ellipses with the same eccentricity, showing varying slot depth with 17

In order to obtain solutions of eqn. 7.29 we set:


A(£,JU) = /?(£)0(r?) (7.30)
Substituting into eqn. 7.29 and applying the usual separation of variable procedure,
two ordinary differential equations are obtained:

^ } - + (C - 2q cos = 0 (7.31)
Rectangular and elliptical corrugated horns 195
and

^ P ~(C-2q cosh 2|>R($) = 0 (7.32)

where Cis the separation constant.


Eqn. 7.31 is the angular Mathieu differential equation; eqn. 7.32 which follows
from eqn. 7.31 by the transformation r\ = ± j£, is the modified Mathieu differential
equation.
For a physically admissible single-valued electromagnetic field, A(£, r?) must be
a periodic function of rj, of period n or 2n. The separation constant C, which in this
case must be a function of q, is an infinite set of characteristic values for every q.
When q is real the characteristic values are real; when q is negative imaginary, as
it is in one of the cases which is considered, the characteristic values can be either
real or complex.
Corresponding to q = 0 there are two independent periodic solutions, namely
sin (vq) and cos (vrj) with the separation constant C = v2, where v is an integer. It
can be shown that when q differs from zero, a characteristic value C determines
one and only one periodic solution which is either even or odd in 17. The charac-
teristic value C, giving rise to even and odd solutions are denoted by av(q) and bv(q),
respectively. The subscript v identifies those sets of characteristic values which
approach v2 as q -> 0.
For an arbitrary positive real q, the periodic solutions of Mathieu eqn. 7.31 are:

= cev(r}, q) even av(q)


Hv) (7.33)
= sev(n,q) odd bv(q)

where cev(r\, q) and sev(r], q) are, respectively, the even and odd angular Mathieu
functions. The corresponding solutions for the modified Mathieu eqn. 7.32 are

(7.34)
= L3Sev(l q) + L4Geyv& q) odd bv(q)

where Cev(%, q) and Sev{%, q) are radial even and odd Mathieu functions of the
first kind, Feyv(%, q) and Geyv(%, q) are radial even and odd Mathieu functions of
the second kind. L\, L2, L3 and Z 4 are constants.
The fields in the inner region of the corrugated waveguide for modes whose
axial magnetic field component is represented by the even Mathieu functions are
196 Rectangular and elliptical corrugated horns

then given by:

Z BvSev(qu%)sev(qur\)
v=o

V- 1

. &

k
3 -jj^i YrAvCev{ql^)cev{quri)

(7.35)

k
A v

- j -jj^Y. AvCe'v(qu%)cev(gl9ri)

The factor exp j (cot ~ j(3z) is assumed throughout, and


/ = /*(cosh 2 £-cos 2 T?) 1/2
These equations are derived under the assumption that there are a large number
of corrugations per wavelength and that the higher order space harmonics can be
ignored.
The fields in the corrugations will be assumed to be those of a cut-off TM
type mode. One of the complications of the analysis of elliptical corrugated wave-
guides is that the ellipse which forms the outer boundary of the waveguide has a
different eccentricity from the ellipse which forms the inner boundary. This can
be seen from Fig. 1.1(6). If the eccentricities of the two ellipses were identical
then the slot depth would vary around the perimeter. Thus a separate set of
parameters q2, £2 and eigennumber r are needed in the outer region.
Rectangular and elliptical corrugated horns 197

The field components in the outer region are then given by:

(7.36)
IK

where 4q{ = h2(k2 - j32), 4q2 =h2k\k2 = co2e0Mo.


In the above equation

j r r T ( 7 3 7 )

The characteristic equation is obtained in the usual manner by matching the


fields at the boundary £ = £1. The equation must be derived assuming an infinite
set of eigennumbers v and r. This leads to considerable complexity in the analysis
and is due to the presence of the angular Mathieu functions which are not only
functions of 17, the angular co-ordinate, but also of the electrical properties of
the medium in which they apply. The impedance of the slot region 2EZI2H11 is a
function of the aximuthal position (77) around the waveguide. This is in contrast
to the circular waveguide where the impedance is independent of the azimuthal
co-ordinate.
After manipulation the characteristic equation for even HE modes with only
the first term in the v summation becomes:

+ J32 !Sex(qu$l) C l U l
*^ l\ a1>rSr>1 = 0 (7.38)
where
r27T
se
0v>n = Jo v(quV)se^(q

A similar equation can be obtained for odd type Mathieu functions which then
198 Rectangular and elliptical corrugated horns

leads to odd type HEn modes. If the elliptical waveguide is made circular
(eccentricity = 0), then eqn. 7.38 becomes the familiar characteristic equation for
circular corrugated waveguides.
The computation of the characteristic equation is time consuming because the
Mathieu functions are difficult to calculate. The angular Mathieu functions may be
expanded in terms of an infinite series of trigonometric functions and the modified
Mathieu functions may be expanded in terms of an infinite series of products of
Bessel functions [178]. Computation of the Mathieu functions involves finding
the root of an infinite continued-fraction transcendental equation and then using
a recurrence relation.
Typical dispersion characteristics are shown in Fig. 7.8 for even and odd HE n
modes. The two modes are degenerate at the frequency where the slot depth is
approximately a quarter of a wavelength deep.

Fig. 7.8 Dispersion curves for HEn mode, ex —0-7,e2 = 0-56


odd mode
even mode

The radiation characteristics of elliptical corrugated waveguides can be derived


by Fourier transforming the aperture fields to give the far-fields. Jansen and Jeuken
[78] have done this under the assumption that the impedance of the slots are
Zn = 0 and Z = °°, and are independent of frequency. They found that the far-
field is circularly polarised everywhere, if the even and odd HEn modes in the
aperture had a phase difference of 90°. They also found that the even mode and
the corresponding odd mode give rise to the same radiation pattern.
Appendix

Cylindrical corrugated waveguide


analysis including space harmonics

Introduction

In this appendix the exact mathematical analysis including space harmonics is


presented for cylindrical corrugated waveguides.
The propagation equation, giving a solution for angular frequency OJ in terms of
the phase change coefficient |3, is derived by the usual technique for waveguides
with two or more regions. That is, the waveguide is divided into the inner region
I ( r O x ) and the outer region ll(r<ri) (slot region). The field expansions which
satisfy Maxwell's equations in the two regions are matched at the common
boundary r = rx and the field coefficients eliminated so as to leave a propagation
equation determinant. This transcendental determinant is equated to zero and the
correct |3:co relationship computed by an iterative process.
The field coefficients and hence electric and magnetic field components are
obtained by back substitution using the propagation determinant.
The geometry to be analysed is shown in Fig. 3.2. The waveguide is assumed to
be infinite in length. In this appendix we adopt a notation widely used in waveguide
analysis and call the azimuthal order n, whereas elsewhere in this book m has been
used to follow convention in papers dealing specifically with horns.
The slots will support a TMn0 mode (Ez, Er, H^ ) and an infinite set of TM nm
and TE n m standing waves which satisfy the boundary conditions at z = ± (b/2) and
r - rQ. In practice the infinite set is truncated at a finite value L so that
E2 E2m
Er= X Erm
m=0

(0
Hz Hzm
L
Hr= Hr
„,?, m
200 Cylindrical corrugated waveguide analysis including space harmonics

The simple surface impedance theory assumes that only the TMn0 mode is present
in the slot, i.e. L — 0 in this analysis. This mode has no z dependence in the slot.
In periodic structures, Floquet's therein states that it is sufficient to consider the
fields in just one slot or unit cell. The fields in any other unit cell, distance Qp in
the z direction (Q interger) from the basic cell are e^oQp times those in the basic
cell where j30p is the phase change per unit cell.
In the inner region, r<rit the boundary conditions to be satisfied are that the
tangential electric field and the normal magnetic flux density vanish at r = rx for
b/2 < \z\ < p / 2 . All fields over the plane z = p/2 will be e^° p times the corre-
sponding fields at z = — (p/2). Each fundamental mode in the inner region has
associated with it a double infinite set of space harmonics. As in the case of the
outer region the series is restricted to a finite value, ±K, so that
E = E (2)
' X '»
and similarly for all other field components.
The Ez and Hz fields are derived from the wave equation and the other field
components from:
1 d2Ez JCOJUO bHz
E +

*r bz2 T^ dr

r
" K2 3^3z K2r

_
K2 brbz r 3^
All fields are assumed to have a time dependence «

Field components in corrugated waveguide

Region I (inner) r<rx

(4b)
Cylindrical corrugated waveguide analysis including space harmonics 201

N nJn(KNr)
(KNr) ""
(4e)

(4/)

Region II (outer) r>rx


(5a)
cosr?' m z] si (5b)

(5c)
v1 m

COS77 m Z

(5d)

^ sinr? m z
J

amn ,, 2
v1 m'/
202 Cylindrical corrugated waveguide analysis including space harmonics
In the above equations the symbols have the following meanings:
2rtN
ftv = £o + —
/" T/~ \ 2 —2 ni 1
=:
\KjqY) Y HN^"

C X0

_ 2m7T

nmr = (2m~l)~ r

(TmY)2 =r-2-(rlmr)2

K
n — Jn\imr) 1
n\\mr0) J
n\\mr0) I
n\\mr)

*n = 4(r m r) ^ ( r m r 0 ) - ^ ( r r o r 0 ) F n (r m r)
s% = «/n(r m r)r n (r m r 0 )-/ n (r m r 0 )^(r m r)
^n = 4(r m r)F n (r m r 0 )-4(r m r 0 )F n (r m r)
i^^,' R&sts&iTlnf) replaces (r m r).
In the above and in eqns. 4 and 5, when KNY, VmY or T^r are imaginary, Jn(x) is
replaced by In (x) and F n (x) by Kn (x).
AN, BN coefficients of E and H mode components in inner region
cm, c'm coefficients of E mode components in outer region
dm, d'm coefficients of H mode components in outer region
y0 admittance of free space

Propagation equation

The fields in regions I and II are matched at r = rt by equating the z and ty


components of the total fields at the interface r = r1. The Ez and E^, field com-
ponents in the slot —(b/2) <z < (b/2) are equated to the Ez and E^, fields over
the length -~(b/2)<Cz<(b/2) in the inner region and the orthogonality of the
e~j%2 functions used to give the coefficients of any space harmonic in the inner
region in terms of all the coefficients in the slot region.
Similarly, the Hz and H^ field components are equated over the length —
(b/2) < z < (b/2) and the orthogonality of the sin (r]mz) and cos (v}mz) functions
Cylindrical corrugated waveguide analysis including space harmonics 203

used to give the coefficients of any mode in the slot region in terms of all the
coefficients in the inner region.
Hence:

I"'* EW*N'te =\m E?eifiN' (6)


J-p/2 J-b/2
rP/2 rb/2
£^eJflw'dz = ., EtfeVN'dz (7)
J-P/2 J-O/2
b/2 UCOS^2 rb/2 COS t]'mZ
H = Hl AZ
b/2 * • ^ U/2 (8)
6/2 on
sinn m z sin7?mz
b/2 COS17mZ rb/2 COS r ? m 2
//|J dz = | Hi dz (9)

The orthogonality integrals are on the left-hand side while the integrals on the right-
hand side are of the form:

-b/2

= \ J_ J 2 [cos (r?m - j3iv)z + cos (r?m + ^ ) z ] dz

or

2
\
Similarly:

cos(
f"W2). „., (U)
— i/3 = I sin T?W
J-b/2
sm Vf-'/v«-'/ •*-/ / 1 o~\
204 Cylindrical corrugated waveguide analysis including space harmonics

U=

(13)
Then substituting the field components of eqns. 4 and 5 into eqns. 6 to 9 gives:

Ez components:

ANpJn(KNri) = X [CmSDnh+c'mSDnI2] (14)


m=0

components:

\\mr\)

f o r 7 V = 0 , ± 1 , ± 2, . . . ± A ' .

Real and imaginary i/xj, components:

-r \c

= I iBi^—nJniKjr^+Aj-p-f^Kjr^h (16)

form=0, 1,2.. ./,.

form = 1,2,.. .Z.


Cylindrical corrugated waveguide analysis including space harmonics 205

Real and imaginary Hz components:


D
nI == -Kf (18)

(19)
=- K

both for m = 1,2,. . . 1 .


Eqn. 16, left-hand side, should be multiplied by two for m = 0. Substituting
eqns. 16 to 19 into eqns. 14 and 15 gives two sets of simultaneous equations
for the As and B$:

+ VNIBj\ = 0
(20)

I + YmBj] = 0
/ =- K
for TV = 0, ± 1, ± 2, ± A", where

(21)
m =o

I (22)
m =0

PNI (23)

T) Y\
YNI = SIN~
is the kronecker delta, = 1 for / = N; = 0 for / =£ iV.

Z
m=0

+ 2 cos (Pji/2) (frv i/2) cos ( ^ 3/2)

(25)
6
I u i s l _ L | UIS z"£-
soo — - soo
UIS (j.. ms
at)
UTS
(93) Uc, A
*\ soo U SOO £ -
ucli = iu/ - " I T \ I -7
N* 7 _i_ soo
soo
_ ™IN
so/uoLUjey aoeds Buipnpuf s/sA/eue apmBdABAA paieBnnoo /eoupu//AQ QQZ
Cylindrical corrugated waveguide analysis including space harmonics 207

SD' irXm-
cos

2Jn(KIr1)T1

cos ! cos (28)


T
em = 1 for m = 0; = 2 for m ^= 0.
In the above equations, the integrals/i — 74 have been written for convenience as:

C0S
2 / \ 0 /~iVm

/3jv =

/ 4 j v = ft cos

where
(-If (-1)"

The determinantal equation to be solved is

U V
H = = 0 (29)
X Y
if is &4K+2 square determinant and U, V, X, Fare ^4-1 square sub-determinants.
Then
208 Cylindrical corrugated waveguide analysis including space harmonics

Um = HK+1 +N,K+I+I
=
VNI HK+ l+N,3K+2+ I

YNI = H3K+2+ N,K+1+I

YNI — H3K + 2+ N,3K+2+ I

Computing procedure
For a fixed $rx and trial values of oorxjc the determinant H can be computed. An
iterative process is used to predict a more accurate value of corjc. This process
being repeated until the desired accuracy is achieved.

Normalisation of equations
For a large number of space harmonics (K large) the factor (K^i)2 =r\ —
2nl(ri/p))2 becomes large and negative. Then

becomes extremely large. This causes problems in the determinant evaluation.


To avoid this difficulty the A and B coefficients can be written as
AN =

In the final eqns. 20 to 29 all the terms of a row are divided by Jn(Kjri).

Special case when L =0, no slot modes


In this case the transverse fields of eqns. 5 become
EZQ =

Hz =0
A (30)
=

—S^

Then equating these field components to the field components of region I, eqns.
4, gives the simplified characteristic equation:
niV 1 h K
T* ___ \

Sn H P N=-

(31)
Cylindrical corrugated waveguide analysis including space harmonics 209

Calculation of the electric and magnetic field components

The matrix equation is

<32>
' t" y"] [11 " °
To solve this equation we must normalise against one of the coefficients in order
to provide an equation of the form

U V
(33)
X Y NO

The procedure for solution is best explained with an example. The matrix equation
for the K = 1 case will be

Co-i ^oo u0+1 ^00

tf+1-1 tf+io V+i-


= 0 (34)
X-1 _ i X- io x.l+ i Y-x-

*Vl ^00 IV i
A
+ l -1 ^+10 x+1 + i Y
+ l .

The procedure is then:


(i) use either Ao ox Bo as reference.
(ii) place column K 4- 1 (for Ao) or 3K + 2 (for /?0) at the end of the matrix by
moving each column on the right of K + 1 (or 3K + 2) one place to the left,
(iii) Normalise coefficients and ignore the last row, giving (for i40)

F_!_! ^-10 " - 1 + 1 ^-10

U0 + i ^0-1 Foo ^0 + l £/oo 4 + 1 /^

Kl-1 n,o n1+ 1 F+10 = 0 (35)


X_! + 1 y-io , AT_10 fioMo
x0+1 ~y
^00
y0+. 5+1/4

for convenience the main matrix has been written in the form of an 'augmented
matrix', i.e. the right-hand side of eqn. 33 has been included in the matrix on the
left-hand side,
(iv) This matrix equation is solved for AN$ and i ^ s .
210 Cylindrical corrugated waveguide analysis including space harmonics

(v) The AN$ and BNs are substituted into eqns. 18 and 19 to obtain dm anddm,
then into eqns. 16 and 17 to obtain cm and cm.

sin (36)
**• n I—K

d 7r(m - i)D' Im Bj (37)


'm = ~5' I cos

(V r

Prediction of radiation characteristics

The far-field radiation patterns can be computed in closed form using the Fourier
Transform method. The radiated fields Epd and Ep^ are given in terms of the
aperture fields Er E^ by
J
e c
P {^r cos (0 - * ) + ^
J
sin (0 - *)}
0

(40)

j e
sin (0 - * ) + ^ * cos

(41)
Cylindrical corrugated waveguide analysis including space harmonics 211

Substituting the expressions for the space harmonic field components of the
corrugated waveguide given in eqn. 4 in the above equations leads to the required
far-field radiation formulae. The resulting equations can be evaluated in closed
form by making use of the following integral expressions:

2nf °?ns n<t>Jn(x) (42)

Jo
— I sin<

Jn(ri sin B)KNrlJn{KNrl)-Jn{rl sin B)J'n(K^r^r^nJ

KM — I I

&N l l sin i

Jn(Fi ski e^NrxJniKNr^-JniFi sin e)fn(KNrl)rl mid


2

- I - I sin 2 (

Eqns. 43 and 44 are the Lommel integrals.


The resulting closed-form expressions for eqns. 40 and 41 are given for theMh
space harmonic, after some manipulation, by
-i—R
-f+ 1 c
E
[A
e-w ^ie CQSK0 I -

(45)
- jj —
c R r -
f+1rie cos (9 sin n<t>

(46)
where
J'n(ri %in B)KNrlJn(KNri) — Jn(ji sin B)fn{KNri)Tl sin B

(47)
212 Cylindrical corrugated waveguide analysis including space harmonics

The total radiated field is


+N
Ee(0,4>) = I Epee~^Nzcosncj> (48)
-N

+ JV

>) = X E^t'^N^inn^ (49)


-N
Bibliography

1 ADATIA, N.A., RUDGE, A.W. and PARINI, C.G.: 'Mathematical modelling of the
radiation fields from microwave primary-feed antennas', Proc 7th European Microwave
conf., Copenhagen, 1977, pp. 329
2 ADATIA, N.A., WATSON, B., DEWEY, R. and RUDGE, A.W.: 'Studies of feeds and
reflectors for high gain spacecraft antennas', ERA Report No. RFTC 220676, (ESA
Contract 2742/76/NL/AK), Feb. 1978
3 Al-HARIRI, A.M.B., CLARRICOATS, P.J.B. and OLVER, A.D.: 'Low attenuation
properties of corrugated rectangular waveguide', Electron. Lett., 191 A, 10, p. 25
4 AL-HARIRI, A.M.B.: 'Low attenuation microwave waveguides', Ph.D. thesis, London
University, Oct 1974
5 ASHTON, R.W.: 'Improvements in or relating to corrugated horns (dual depth horn)'
UK Patent 1498 905, January 1978
6 AU, H.K.: 'Hybrid modes in conical corrugated horns with narrow flare angle and
arbitrary length', Electron. Lett., 1970, 6, p. 769
7 AUBRY, C. and BITTER, D.: 'Radiation patterns of a corrugated conical horn in terms
of Laguerre-Gaussian function', Electron. Lett., 1975, 11, p. 155-6
8 BACH ANDERSON, J.: 'Propagation in the general, rectangular waveguide', Proc. 5th
European Microwave Conf., Hamburg, Sept 1975, pp. 595-599
9 BAHRET, W.F. and PETERS, L.: 'Small aperture - small flare angle corrugated horns',
IEEE Trans., 1968, AP-16, p. 494
10 BALDWIN, R. and McINNES, P.A.: 'Attenuation in corrugated rectangular waveguide',
Electron. Lett., 1971, 7, pp. 770-772
11 BALDWIN, R. and McINNES, P.A.: 'Corrugated rectangular horns for use as microwave
feeds', Proc. IEE, 1975, 122, pp. 465-469
12 BALDWIN, R. and McINNES, P.A.: 'Radiation from rectangular horns with modified
boundaries', Proc. European Microwave Conference, 1971, p B7/1
13 BALDWIN, R. and Mclnnes, P.A.: 'Surface wave radiation from a corrugated horn',
Electron. Lett., 1970, 6, pp. 259-260
14 BALDWIN, R. and McINNES, P.A.: 'A rectangular corrugated feed horn' IEEE Trans
1975, AP-23, pp. 814-817
15 BATHKER, D.A.: 'A stepped mode transducer using homogeneous waveguides', Trans
IEEE, 1967, MTT-15, pp. 128-130
16 BIELLI, P., DE PADOVA, S. and PAGANA, E.: 'Analysis and synthesis of conical
corrugated horns', Proc. 6th European Microwave Conf., Sept 1967, Rome, pp. 153
17 BITTER, D. and AUBRY,1 C : 'Sum and difference radiation patterns of a corrugated
conical horn by means of Laguerre-Gaussian functions', Proc. 5th European Microwave
Conference, 1975, pp. 677-681
18 BRUNSTEIN, S.A.: 'A new wideband feedhorn with equal E- and H-plane beam widths
214 Bibliography
and suppressed sidelobes', Jet Propulsion Laboratory. Deep Space Network Space Pro-
grams Summary 37-58, July 1969, pp. 61-64
19 BRYANT, G.H.: 'Monopulse multimode feed for military terminals', IEE Conf. on Earth
Station Tech. (pub 72), Oct 1970, pp. 245-249.
20 BRYANT, G.H.: 'Propagation in corrugated waveguides', Proc. IEE, 1969, 116, pp.
203-213
21 BUCHMAYER, S.K.: 'Corrugations lock horns with poor beamshapes', Microwaves,
1973, 12, pp. 44-49
22 CALDECOTT, R., MENTZER, C.A. and PETERS, L.: 'The corrugated horn as an antenna
range standard', IEEE Trans., 1973, AP-21, p. 562
23 CHONG, S.L.: 'Corrugated waveguides for low attenuation transmission', Ph.D. thesis,
London University, Now. 1974
24 CHU, T.S. and LEGG, W.E.: 'Gain of corrugated conical horns', 7>ans. IEEE, 1982,
AP-30, pp. 698-703
25 CLARRICOATS, PJ.B. and ELLIOT, R.D.: 'Multimode corrugated waveguide feed for
monopulse radar', i>oc. IEE, 1981, 128 Pt. H,pp. 102-110
26 CLARRICOATS, PJ.B. and OLVER, A.D.: 'Improvements in the theory and design of
corrugated waveguide feeds for low cross polarisation antenna applications', Proc. URSI
Int. Symp. on EM Theory, Stanford, June 1977, pp. 233
27 CLARRICOATS, PJ.B. and POULTON, G.T.: 'High efficiency microwave reflector
antennas: A review', Proc. IEEE, 1977, 65, pp. 1470-1504
28 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Analysis of spherical hybrid modes in a
corrugated conical horn', Electron. Lett., 1969, 5, pp. 189-190
29 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Propagation and radiation behaviour of
corrugated feeds, Pt I Corrugated waveguide feeds', Proc. IEE, 1971, 118, p. 167
30 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Propagation and radiation behaviour of cor-
rugated feeds, Pt II Corrugated conical horn feeds', Proc. IEE, 1971 118, p. 187
31 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Radiation from wide flare angle scalar horns',
Electron. Lett., 1969, 5, p. 376
32 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Radiation pattern of a lens corrected conical
scalar horn', Electron. Lett., 1968, 4, pp. 592-593
33 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Scalar feeds for earth station antennas', Proc.
Conf. on Earth Station Technology, London, Oct 1970, pp. 240-244
34 CLARRICOATS, PJ.B. and SAHA, P.K.: 'Theoretical analysis of cylindrical hybrid
modes in a corrugated horn', Electron. Lett., 1969,5, pp. 187-189
35 CLARRICOATS, PJ.B. and SENG, L.M.: 'Influence of horn length on the radiation
pattern of oblique flare angle corrugated horns', Electron. Lett., 1973, 9, p. 15
36 CLARRICOATS, PJ.B. and SENG, L.M.: 'Propagation and radiation characteristics of
corrugated horns', Electron. Lett., 1973,9, pp. 7-9
37 CLARRICOATS, PJ.B., MAHMOUD, S.F. and OLVER, A.D.: 'Cross-polar behaviour of
wide-angle corrugated horns', IEEE Symp. on Ant. & Prop., Los Angeles, June 1981,
pp. 65-68
38 CLARRICOATS, PJ.B., MAHMOUD, S.F. and OLVER, A.D.: 'Low cross-polar radiation
from wide flare angle conical corrugated horns', Proc. 11th European Microwave Conf.,
Amsterdam, Sept. 1981, pp. 735-739
39 CLARRICOATS, PJ.B., OLVER, A.D. and CHONG, S.L.: 'Attenuation in corrugated
circular waveguides, Pt. I Theory', Proc. IEE, 1975,122, p. 1173
40 CLARRICOATS, PJ.B., OLVER, A.D. and CHONG, S.L.: 'Attenuation in corrugated
circular waveguides, Pt. II Experiment', Proc. IEE, 1975, 122, p. 1180
41 CLARRICOATS, PJ.B., OLVER, A.D. and PARINI, C.G.: 'Optimum design of corrugated
feeds for low crosspolarisation radiation', Proc. 6th European Microwave Conf., Rome
1976,pp.148
42 CLARRICOATS, PJ.B., OLVER, A.D. and PARINI, C.G.: 'Radiation method for the
measurement of mode conversion levels', Electron. Lett., 1974,10, p. 252
Bibliography 215

43 CLARRICOATS, P.J.B., OLVER, A.D. and SAHA, P.K.: 'Near field radiation charac-
teristics of corrugated horns', Electron. Lett., 1971, 7, pp. 446-448
44 CLARRICOATS, P.J.B., OLVER, A.D., PARINI, C.G. and POULTON, G.T.: 'Corrugated
waveguide feeders for microwave antennas', Proc. 5th European Microwave Conference,
Hamburg, Sept. 1975. p. 240
45 CLARRICOATS, P.J.B.: 'Hybrid mode feeds for microwave reflector antennas" in Modern
Topics in Electromagnetics and Antennas (Peter Peregrinus, 1979)
46 CLARRICOATS, P.J.B.: 'Similarities in the electromagnetic behaviour of optical wave-
guides and corrugated feeds', Electron. Lett., 1970, 6, 6 March,
47 CLARRICOATS, P.J.B.: 'Some recent advances in microwave reflector antennas', Proc.
IEE, 1979, 126, pp. 9-25
48 COLEMAN, H.P.: 'Paraboloidal reflector off-set fed with a corrugated conical horn',
Trans. IEEE, 1975, AP-23, pp. 817-819
49 COOPER, P.N.: 'Complex propagation coefficients and the step discontinuity in
corrugated cylindrical waveguide', Electron. Lett., 1971, 7, pp. 135-6
50 DAVIS, D.: 'Corrugations improve monopulse feed horns', Microwaves, 1972, pp. 5 8 -
63
51 DOANE, J.L.: 'Mode converters for generating the HE n mode from TE01 in circular
waveguide', Int. J. Electronics, 1982
52 DRABOWITCH, S.: 'A new tracking mode coupler using a corrugated feed for satellite
communication earth station antenna', Proc. 6th European Microwave Conf., Rome,
1976, p. 287
53 DRAGONE, C: 'An improved antenna for microwave radio systems consisting of two
cylindrical reflectors and a corrugated horn', Bell Syst. Tech. J., 1974, 53, pp. 1351 —
1377
54 DRAGONE, C: 'Characteristics of a broadband microwave corrugated feed: A comparison
between theory and experiment', Bell Syst. Tech. /., 1977, 56, pp. 869-888
55 DRAGONE, C: 'High frequency behaviour of waveguides with finite surface impedances',
Bell Syst. Tech. J., 1981,60, pp. 80-116
56 DRAGONE, C: 'Reflection, transmission and mode conversion in a corrugated feed',
Bell. Syst. Tech. J., 1977, 56, pp. 835-867
57 DYBDAL, R.B., PEAK, W. and PETERS, L.: 'Propagation in corrugated waveguides',
Proc. IEE, 1970, 117, pp. 931-932
58 DYBDAL, R.B., PETERS, L. and PEAK, W.: 'Rectangular waveguides with impedance
walls', Trans. IEEE, 1971, MTT-19, p. 2
59 ELLIOT, R.D. and CLARRICOATS, P.J.B.: 'Corrugated waveguide monopulse feed',
Electron. Lett, 1980, 16, pp. 324
60 ELLIOTT, R.S.: 'On the theory of corrugated plane surfaces', IRE Trans., 1974, AP-2,
pp. 71-81
61 ESTIN, A.J., STUBENRAUCH, C.F., REPJAR, A.G. and NEWELL, A.C.: 'Optimized
wavelength-sized scalar horns as antenna radiation standards', Trans. IEE, March 1982,
IM-31,pp. 53
62 FRANK, Z.: 'Very wide band corrugated horns', Electron. Lett., 1975, 11, 20 pp. 131-
133
63 GHOSH, S.. ADATIA, N. and WATSON, B.K.: 'Hybrid mode feed for multiband appli-
cations having a dual-depth corrugation boundary', Electron. Lett. 1982, 18, pp. 860-862
64 GHOSH, S.: 'A corrugated waveguide feed for discrete multiband applications having dual
depth corrugations', IEEE Int. Symp. on Antennas & Propagation, Quebec, Canada, June
1980, pp. 217-220
65 GRUNNER, R.W.: 'A 4- and 6-GHz prime focus, CP feed with circular pattern symmetry',
IEEE Ant. & Prop. Symp. Digest, June 1974, pp. 72-74
66 GUISSARD, A.C. and KANDIL, A.A.: 'A new solution for the fields in corrugated square
waveguides', Int. Conf. on Ant. & Prop., Nov. 1978, IEE Conf. Pub. 169, pp. 235-239
216 Bibliography

67 GUY, R.F.E. and ASHTON, R.W.: 'Cross-polar performance of an elliptical corrugated


horn antenna', Marrow. Lett., 1979, 15, p. 400
68 HOCKHAM, G.A. and ELLIOT, R.D.: 'Optimum design of choked waveguide feeds', Int.
Conf. on Antennas & Propagation, London, Nov. 1978, IEE Conf. Pub. 169, p. 369
69 HOCKHAM, G.A. and OLVER, A.D.: 'Cross-polarized performance of small corrugated
feeds', Proc. IEEE Ant. & Prop. Symp. (Washington), 1978, p. 431
70 HOCKHAM, G.A. and SENG, L.M.: 'Radiation from a parallel plate waveguide with a
finite corrugated flange', Electron. Lett., 1975. 11, p. 292
71 HOCKHAM, G.A.: 'Investigation of a 90 degree corrugated horn', Electron. Lett., 1976,
12, p. 199
72 HOGG, D.C. et al.\ 'An antenna for dual-wavelength radiometry at 21 and 32 GHz', Trans.
IEE, 1979, AP-27, pp. 764-771
73 ISSAAC, P.R.: 'Comments on rectangular waveguides with impedance walls', Trans. IEEE,
1974,MTT-22, p. 972
74 JAMES, G.L. and THOMAS, MacA.B.: 'TEH to HE11 corrugated cylindrical waveguide
mode converters using ring-loaded slots', Trans. IEEE, 1982, MTT-30, pp. 278
75 JAMES, G.L.: 'Analysis and design of TEH to HE11 corrugated cylindrical waveguide
mode converters using ring-loaded slots', Trans. IEEE, 1981, MTT-29, pp. 1059-1066
76 JAMES, G.L.: 'Surface reactance of corrugated planes', Electron. Lett., 1979, 15, p. 751
77 JAMES, G.L.: 'TEH to HE11 mode converters for small angle corrugated horns', Trans.
IEEE, 1982, AP-30,1057-1062
78 JANSEN, J.K.M. and JEUKEN, M.E.J.: 'Circularly polarised horn antenna with an
asymmetrical pattern', Proc. Colloquium on Microwave Communications, Budapest, 1974,
pp. 179-188
79 JANSEN, J.K.M. and JEUKEN, M.E.J.: 'Surface waves in the corrugated conical horn',
Electron. Lett., 1972, 8, p. 342
80 JANSEN, J.M.K., JEUKEN M.E.J. and LAMBRECHTSE, C.W.: 'The scalar feed', Archiv
fur Elektronik & Uber, 1972, 26, pp. 22-30
81 JANSSEN, M.A. and BEDNARCYK S.M.: 'Pattern measurements of a low-sidelobe horn
antenna', Trans. IEEE, 1979, AP-27, pp. 551-555
82 JEUKEN, M.E.J. and LAMBRECHTSE' C.W.: 'Small corrugated conical horn antenna
with wide flare angle', Electron. Lett., 1969, 5, pp. 489-490
83 JEUKEN, M.E.J. and THURLINGS, L.F.G.: 'The corrugated elliptical horn antenna',
IEEE Int. Symp. on Ant. & Prop. Illinois, 1975
84 JEUKEN, M.E.J. and VOKURKA, V.J.: 'Multi-frequency band corrugated conical horn
antennas', Proc. European Microwave Conf., Brussels, Sept. 1973
85 JEUKEN, M.E.J.: 'Experimental radiation patterns of the corrugated conical horn antenna
with small flare angle', Electron. Lett., 1969, 5, pp. 484
86 JEUKEN, M.E.J.: 'Frequency independent and symmetric properties of corrugated conical
horn antennas with small flare angles', Ph.D. thesis, Eindhoven Univ. of Technology,
Netherlands, 1970
87 KAMAL, A.K., GUPTA, S.C. and NAIR, R.A.: 'Radiation characteristics of a dielectric
sphere loaded corrugated E-plane sectoral horn', Trans. IEEE, 1978, AP-26, pp. 625-626
88 KAY, A.F.: 'A wide flare angle horn. A novel feed for low noise broadband and high
aperture efficiency antennas', US Air Force Cambridge Research Labs Report 62-757,
Oct 1962
89 KAY, A.F.: 'The scalar feed', US Air Force Cambridge Research Labs Report 62-347,
AD60169, March 1964
90 KNOP, CM. and WIESSENFARTH, H.J.: 'On the radiation from an open ended corru-
gated pipe carrying the HE11 mode', Trans. IEEE, 1972, AP-20, p. 644
91 KOCH, G.F.: 'Coaxial feeds for high aperture efficiency and low spill-over of paraboloidal
reflector antennas', Trans. IEEE, 1973, AP-21, pp. 164-169
Bibliography 217

92 LAWRIE, R.E. and PETERS, L.: 'Modifications of horn antennas for low sidelobe levels',
Trans. IEEE, 1966, AP-14, p. 605
93 LOEFER, G.R., NEWTON, J.M., SCHUCHARDT, J.M. and DEES, J.W.: 'Computer
analysis speeds corrugated horn design', Microwaves, 199r6, pp. 53-65
94 LOVE, A.W. (Ed): 'Electromagnetic horn antennas' (IEEE Press, N.Y., 1976)
95 MAHMOUD, S.F. and CLARRICOATS, P.J.B.: 'Radiation from wide flare-angle corru-
gated conical horns', Proc. IEE, 1982,129, Pt. H. pp. 221-228
96 MANWARREN, T. and FARRAR, A.: 'Pattern shaping with hybrid mode corrugated
horns', Trans. IEEE, 1974, AP-22, pp. 484-487
97 MARINCIC, A. and MILOVANOVIC, B.: 'Frequency dependence of corrugated horn
mode order', Proc. 7th European Microwave Conf., Copenhagen, 1977, p. 306
98 MENTZER, C.A. and PETERS, L.: 'Pattern analysis of corrugated horn antennas', Trans.
IEEE, 1976, AP-24, pp. 304-309
99 MENTZER, C.A.: 'A corrugated horn antenna using V shape corrugations', Trans. IEEE,
1975, AP-23, p. 93
100 MENTZER, L. and PETERS, L.: 'Properties of cut-off corrugated surfaces for corrugated
horn design', Trans. IEEE, 197r4, AP-22, pp. 191-196
101 MINNETT, H.C. and THOMAS, MacA.B.: 'A method of synthesising radiation patterns
with axial symmetry', Trans. IEEE, 1966, AP-14, pp. 654-656
102 MINNETT, J.C. and THOMAS, MacA.B.: 'Propagation and radiation behaviour of corru-
gated feeds', Proc. IEE, 1972, 119, p. 1280
103 MORRIS, G.: 'A broadband constant beamwidth corrugated rectangular horn', Trans.
IEEE, 1982, AP-30, pp. 966-974
104 MOTO MISUZAW and FUMIO TAKEDA: 'Radiation characteristics of a corrugated
conical horn', Elec. & Comm. in Japan, 1973,56B
105 McINNES, P.A. and BOOKER, D.D.: 'Computer predicted performance of corrugated
conical feeds using experimental primary radiation patterns', Electron. Lett., 1970, 6,
pp. 18-20
106 McINNES, P.A.: 'Corrugated conical horns with very wide flare angles', Proc. European
Microwave Conf 1971, p. 87
107 NARASIMHAN, M.S. and GOVINDARAJAN, M.S.: 'Eigenvalues of spherical hybrid
modes in corrugated conical horns', Trans. IEEE, 1974, AP-22, pp. 122-123
108 NARASIMHAN, M.S. and RAO, B.V.: 'Diffraction by wide flare angle corrugated conical
horn', Electron Lett., 1970, 6, pp. 469-470
109 NARASIMHAN, M.S. and RAO, B.V.: 'Hybrid modes in corrugated conical horns',
Electron. Lett., 1970, 6, pp. 32-34
110 NARASIMHAN, M.S. and RAO, B.V.: 'Modes in a conical horn: A new approach', Proc.
IEE, 1971, 118, pp. 187-192
111 NARASIMHAN, M.S. and RAO, K.S.: 'GTD analysis of the near-field patterns of conical
and corrugated horns', Trans. IEEE, 1979, AP-27, pp. 705-708
112 NARASIMHAN, M.S. and RAO, V.V.: 'Comments on rectangular waveguides with
impedance walls', Trans. IEEE, 1974, MTT-22, p. 973
113 NARASIMHAN, M.S. and RAO, V.V.: 'Radiation characteristics of corrugated E plane
sectoral horns', Trans. IEEE, 1973, AP-21, pp. 320-327
114 NARASIMHAN, M.S. and RAO, V.V.: 'Radiation from wide flare angle corrugated E
plane sectoral horns', Trans. IEEE, 1974, AP-22, pp. 603-608
115 NARASIMHAN, M.S. and SHESHADARI, M.S.: 'Propagation and radiation characteristics
of dielectric loaded corrugated dual frequency circular waveguide horn feeds', Trans. IEEE,
1979, AP-27, p. 858
116 NARASIMHAN, M.S. and SHESHADRI, M.S.: 'GTD analysis of the radiation patterns of
wide flare corrugated conical and E-plane sectoral horns', Trans. IEEE, 1979, AP-27, pp.
276-279
218 Bibliography

111 NARASIMHAN, M.S.: 'Corrugated conical horns as a space feed for phased array illumi-
nation', Tram. IEEE, 1974, AP-22, p. 720-722
118 NARASIMHAN, M.S.: 'Corrugated conical horns with arbitrary corrugation depth', Radio
& Electrical Engineer, 1973, 43, pp. 188-192
119 NARASIMHAN, M.S.: 'Eigenvalues of spherical hybrid modes in corrugated conical
horns',Proc. IEE, 1973, 120, pp. 965-967
120 NARASIMHAN, M.S.: 'Radiation from conical horns with large flare angles', Trans. IEEE,
1971,AP-19, p. 678
121 NARASIMHAN, M.S.: 'Radiation from conical horns', Int. J. Electronics, 1969, 27, p. 377
122 OHTERA, I. and UJILLE, H.: 'Nomographs for phase centers of conical corrugated and
TEH mode horns', Trans. IEEE, 1974, AP-22, pp. 858-859
123 OLVER, A.D. and DUBROVKA, F.F.: 'Propagation characteristics of open cylindrical
corrugated waveguides', Proc. 12th European Microwave Conf., Helsinki, 1982, p. 306
124 OLVER, A.D., CLARRICOATS, P.J.B. and YANG, K.Z.: 'Predicted performance of dual
depth corrugated horns', IEEE Int. Symp. on Antennas & Propagation, Houston, Texas,
May 1983
125 OLVER, A.D., CLARRICOATS, P.J.B., HOCKHAM, G.A. and PARINI C.G.: 'Design of
corrugated feeds for low cross polarisation', Int. Conf. on Ant. & Prop., London, Nov.
1978, IEE Conf. Pub. 169, p.355
126 OLVER A.D.: 'Design of conical corrugated horns', Jet Propulsion Laboratory TDA
Progress Report 42-65,1981, p. 86-93
127 PARINI, C.G. and OLVER, A.D.: 'Accurate prediction of the cross-polar performance of
narrow flare angle corrugated horns', Int. URSI Symp. on EM Theory, Munich, Aug. 1980
128 PARINI, C.G., CLARRICOATS, P.J.B. and OLVER, A.D.: 'Cross polar radiation from
open ended corrugated waveguides', Electron. Lett. 1975, 11, p. 567
129 PARINI, C.G.: 'Transmission and radiation characteristics of corrugated waveguides',
Ph.D. thesis, London University, Oct. 1976.
130 POTTER, P.D.: 'Efficient antenna systems: A new computer program for the design and
analysis of high performance conical feedhorns', Jet Prop. Lab Tech. Report 32—1526,
Feb. 1973
131 PROFERA, C.E.: 'Improvement of TE11 mode coaxial waveguide and horn radiation
patterns by incorporation of a radial aperture reactance', Trans. IEEE, 1976, AP-24,
pp. 203-206
132 RUDGE, A.W., MILNE, K., OLVER, A.D. and KNIGHT, P. (eds): 'The Handbook of
Antenna Design', Chapter 4 (Peter Peregrinus, 1983)
133 RULF, B. and HURD, R.A.: 'Radiation from an open-waveguide with reactive walls',
Trans. IEEE, 1978, AP-26, pp. 668-673
134 RUMSEY, V.H.: 'Horn antennas with uniform power patterns around their axis', Trans.
IEEE, 1966, AP-14, pp. 656-658
135 SAKR, L. and SARREMEJEAN, A.: 'The performance of square and conical corrugated
horns as monopulse feeds in the millimetric band', Proc. 9th European Microwave Conf.,
Brighton, 1979, p. 346
136 SCHEFFER, H.: 'Improvements in the development of coaxial feeds for paraboloidal
reflector antennas', Proc. 5th European Microwave Conf., Hamburg, 1975, pp. 235-239
137 SHAFAI, L. and HANSEN, J.: 'Matrix formulation of corrugated feeds by using impedance
boundary conditions', Electron. Lett., 1977,13, p. 310
138 SIMONS, A.J. and KAY, A.F.: 'The scalar feed - A high performance feed for large para-
boloid reflectors', IEE Conf. Pub. no21, 1966, pp. 213-217
139 SPIKE, W.: 'Realisation of a hybrid mode secondary focus feed for the 100 m radiotele-
scope at Effelsburg', Proc. 1971 European Microwave Conf. pB6/l
140 TAKEDA, F. and HASHIMOTO, T.: 'Broadbanding of corrugated conical horns by means
of the ring-loaded corrugated waveguide structure', Trans. IEEE, 1978, AP-26, pp. 367-
372
Bibliography 219

141 TAKEICHI, Y., HASHIMOTO, T. and TAKEDA, F.: 'The ring-loaded corrugated
waveguide', Trans. IEEE, 1971, MTT-19, pp. 947-950
142 TERZUOLI, A.J.: 'VSWR properties of E-plane dihedral corrugated horns', Trans. IEEE,
1978, AP-26, pp. 239-243
143 TERZUOZLI, A.J., RICHMOND, J.H. and PETERS, L.: 'Shielding effectiveness of corru-
gations in corrugated horns', Trans. IEEE, 1978, AP-26, p. 654
144 THOMAS, MacA.B. and GREENE, J.: 'A curved-aperture corrugated horn having very
low cross-polar performance', Trans. IEEE, 1982, AP-30, pp. 1068-1072
145 THOMAS, MacA.B.: 'Bandwidth properties of corrugated conical horns', Electron. Lett.,
1969, 5, pp. 561-563
146 THOMAS, MacA.B. and GREENE, K.J.: 'Minimizing the cross-polar energy from corru-
gated horns', IREE Conv., 1979, pp. 13-16
147 THOMAS, MacA.B. and MINNETT, H.C.: 'Modes of propagation in cylindrical waveguides
with anisotropic walls', Proc. IEE, 1978, 125, pp. 929-932
148 THOMAS, MacA.B. and MINNETT, H.C.: 'Propagation in cylindrical waveguides with
anisotropic walls', Radiophysics Publication RPP 1346, CSIRO, Sydney, Jan. 1977
149 THOMAS, MacA.B.: 'Design of corrugated conical horns', Trans. IEEE, 1978, AP-26, pp.
367-372
150 THOMAS, MacA.B.: 'Mode conversion using circumferentially corrugated cylindrical
waveguide', Electron. Lett., 1972, 8, p. 394
151 THOMAS, MacA.B.: 'Prime focus focus one and two hybrid mode feeds', Electron. Lett.,
1970,6, pp. 460-461
152 THOMAS, MacA.B.: 'Theoretical performance of prime focus paraboloids using cylindrical
hybrid mode feeds', iVoc. IEE, 1971,118, pp. 1539-1549
153 THOMAS, R.F. and BATHKER, D.A.: 'A dual hybrid mode feedhorn for DSN antenna
performance enhancement', Jet Prop. Lab. Deep Space Network Progress Report 42-22,
1974, pp. 101-108
154 VIGGH, M.: 'Study of design procedures and limitations for monopulse scalar feeds',
Rome Air Development Centre Report No. RADC-Tr-69-303, ASTIA Document 862515,
Nov. 1969
155 VU, T.B. and HIEN, N.V.: 'A new type of high performance monopulse feed', Trans.
IEEE, 1973, AP-21, pp. 855-857
156 VU, T.B. and VU, Q.M.: 'Corrugated horns - recent developments', Proc. 1971 European
Microwave Conf. pB6/3
157 VU, T.B. and VU, Q.M.: 'Optimum feed for large radiotelescopes: Experimental results',
Electron. Lett., 1970, 6, pp. 159-160
158 VU, T.B.: 'Corrugated horns as high performance monopulse feed', Int. J. Electronics,
1973, 32, pp. 433-444
159 VU, T.B.: 'On the bandwidth characteristics of corrugated feed horns', Int. J. Electronics,
1970,29, p. 449
160 VU, T.B.: 'Optimisation of performance of corrugated feed for paraboloidal antenna',
Int. J. Electronics, 1971, 30, p. 189
161 WATSON, B.K., DANG, N.D. and GHOSH, S.: 'A mode extraction network for RF
sensing in satellite reflector antennas', Int. Conf. on Ant. & Prop. IEE Conf. Publ. 195,
April 1981, pp. 323-327
162 WATSON, B.K., DANG, R., RUDGE, A.W. and OLVER, A.D.: 'Compact low cross-polar
corrugated feed for ECS', IEEE Int. Symp. on Ant. & Prop., Quebec, June 1980
163 WILLIAMS, W.F. and WITHINGTON, J.R.: 'A common aperture S- and X-band feed for
the deep space network', Antenna Applications Symp. University of Illinois, Illinois, Sept
1979
164 WOHLLEBEN, R., MATTES, H. and LOCHNER, O.: 'Simple small primary feed for large
opening angles and high aperture efficiency', Electron. Lett., 1972, 8, pp. 474-476
220 Bibliography

165 WOHLLEBEN, R., WIELHEBINSKI, R. and MATTES, H.: 'Feeds for the 100 m Effels-
berg telescope', Proc. European Microwave Conf., 1971, pp. B5/5
166 YOSHIRO, T., TSUTOMV, M. and FUMIO, T.: 'The ring loaded corrugated waveguide',
Trans. IEEE, 1971, MTT-19, pp. 947-950
167 ZACHARIAH, E.J., VASUDEVAN, K. and NAIR, K.G.: 'Metal flanges with more para-
meters for beam shaping', Trans. IEEE, 1979, AP-27, pp. 708-711
168 FASOLD, D. and GALL, H.: 'Theory and computer program for the radiation pattern
prediction of conical corrugated horns', Nachrichtentechnische Zeitschrift, 1978, 31,
pp. 294
169 FASOLD, D. and PECHER, H.: 'Gain of rectangular horns', Microwave Journal, March
1979, pp. 76-79
170 FASOLD, D. and PECHER, H.: 'Berechnung und Auslegung von Rechteckrillenhorn-
strahlern', Mikrowellen Magazin, 1981, pp. 39
171 KUHN, E. and HOMBACH, V.: 'Computer aided analysis of corrugated horns with axial
or ring-loaded radial slots', 3rd Int. Conf. on Ant. and Prop., IEE Conf. Pub. 219, April
1983, pp. 879-884
172 JAMES, G.L.: 'Propagation and radiation properties of corrugated cylindrical coaxial
waveguides', Trans. IEEE, 1983, AP-31, pp. 477-483
173 SHAFAI, L., KISHK, A.A., ITTIPIBOON, A. and BRIDGRES, E.: 'Performance of cor-
rugated feeds with simple corrugation shapes', IEEE Int. Symp. on Ant. and Prop.,
Houston, Texas, May 1983, pp. 586-588
174 OLVER, A.D., YANG, K.Z. and CLARRICOATS, P.J.B.: 'Dual depth corrugated horn
design', Proc. 13th European Microwave Conf., Nuremburg, Sept 1983, pp. 879-884
175 SRIDHAR, N. and SRIVASTAVA, G.P.: 'Large bandwidth realisation in dual mode
corrugated guide feeds', Proc. 13th European Microwave Conf., Nuremburg, Sept. 1983,
pp.891
176 WANG, H.L. and ZHANG, R.Z.: 'Complex eigenvalues of corrugated waveguide', Proc.
13th European Microwave Conf., Nuremburg, Sept. 1983, pp. 885-890
177 MAHMOUD, S.F.: 'Mode conversion in profiled corrugated conical horns', Proc. IEE,
PtH,Oct 1983,130

GENERAL REFERENCES

178 ABRAMOWITZ, M. and STEGUN, LA.: 'Handbook of Mathematical Functions' (Dover


Publications, New York, 1965)
179 BARTLETT, H.E. and MOSELEY, R.E.: 'Dielguides - high efficiency low noise antenna
feeds', Microwave /., 1966, 9, pp. 53-58
180 BARTLETT, H.E., MOSELEY, R.E. and PIETSCH: United States Patent Nos. 3414903
(Dec. 68), 3430244 (Feb. 1969) and 3611391 (Oct. 1971)
181 BEM, D.J.: 'Electric field distribution in the focal region of an offset paraboloid', Proc.
IEE, 1969, 116, pp. 579-584
182 CLARRICOATS, P.J.B. and SALEMA, C.E.R.C: 'Antennas employing conical-dielectric
horns: Part I - Propagation and Radiation characteristics of dielectric cones, Part II - The
Cassegrain antenna',Proc. IEE, 1973,121, pp. 741-756
183 HARRINGTON, R.F.: 'Time harmonic electromagnetic field' (McGraw Hill, New York,
1961)
184 IEEE Standard Test Procedures for Antennas, IEEE Standard 149-1979, IEEE, New
York, 1979
185 LUDWIG, A.C.: 'The definition of cross-polarisation', IEEE Trans., 1973, AP-21, pp.
116-119
186 MINNETT, H.C. and THOMAS, MacA.B.: 'Fields in the image space of symmetrical
focussing reflectors', Proc. IEE, 1968, 115, p. 1419
Bibliography 221

187 PHILLIPS, C.J.E. and CLARRICOATS, P.J.B.: 'Optimum design of a Gregorian corrected
spherical reflector antenna', Proc. IEE, 1970,117, p. 718
188 POTTER, P.D.: 'A new horn with suppressed sidelobes and equal beamwidths', Microwave
J. 1963,6, pp. 71-76
189 RUDGE, A.W. and ADATIA, N.A.: 'A new class of primary-feed antennas for use with
off-set parabolic reflector antennas', Electron. Lett., 1975, 11, pp. 597-599
190 RUDGE, A.W. and ADATIA, N.A.: 'Matched-feeds for offset parabolic reflector antennas',
Proc. European Microwave Conf. 1976, p. 143
191 RUDGE, A.W., MILNE, K., OLVER, A.D. and KNIGHT, P.K. (eds): 'The Handbook of
Antenna Design', Vol. I (Peter Peregrinus, 1982)
192 THOMAS, MacA.B., MINNETT, H.C. and VU, T.B.: 'Fields in the focal region of a
spherical reflector', IEEE Trans., 1969, AP-17, pp. 229-231
222 Program

Program 1
1000 REM***Program to compute the propagation
1010 REM and attenuation characteristics of
1020 REM corrugated waveguide using surface
1030 REM impedance formulation of propagation
1040 REM coefficient and pertubation formula
1050 REM for attenuation.
1060 DIM C(20), F(20)
1070 REM
1080 PI=3-14159
1090 REM Input Data
1100 INPUT N,RA,R1,D,TT,RH
1110 INPUT FR,C(1)
1120 IF F R = - 1 GOTO 1090
1130 R=1/(RA/R1 + 1)
1140 WW=FR*2*PI*Rl/0-3
1150 IF C(l)=—1 THENC(l)=SQR(WW*2-5-7)-0 01
1160 REM
1170 REM Print Titles
1180 PRINT:PRINT
1190 PRINT SPC(16); "Corrugated Waveguide"
1200 PRINT:PRINT" Azimuthal No. =";N
1210 PRINT" R1/R0 =";R
1220 PRINT" Rl =";R1 "metres"
1230 PRINT" Slot depth =";RA;"metres"
1240 PRINT" Slot Width =";D;"metres"
1250 PRINT" Ridge Width =";TT; "metres"
1260 PRINT" Conductivity =";RH
1270 PRINT
1280 CO=0: T=0: M=0: V=l: CH=1
1290 AC=0-001
1300 GOTO 1310
1310 REM
1320 REM Function Evaluation
1330 BA=C(M+1)
1340 K2=WW*WW-BA*BA
1350 SS=SGN(K2)
1360 IF SS=1 AND ABS (K2) XH)1 THEN K=SQR (K2)
1370 IF SS=-1 AND ABS (K2) X)-01 THEN K=SQR (-K2)
1380 IF SS=0 OR ABS (K2) <=0-01 GOTO 2040
1390 X=WW/R: MM=1: GOSUB 2220
1400 Y1=J1:Y2=J2:Y3=J3:Y4=J4
1410 X=K: MM=SS: GOSUB 2220
1420 Z1=J1:Z2=J2:Z3=J3:Z4=J4
Program 223

1430 X=WW: MM=1: GOSUB 2220


1440 SN=J2*Y3-Y1*J4
1450 SD=J1*Y3-Y1*J3
1460 S=WW*SN/SD
1470 FT=K*Z2/Z1
1480 M=M+1
1490 F(M)=K*K2*SN*Z1*Z2/WW~K*K*Z2*Z2*SD
1500 F(M)= F(M)4-N*N*BA*BA*Z1 *Z1 *SD/(WW*WW)
1520 IF V=l GOTO 1550
1530 IF M>=2 GOTO 1660
1540 GOTO 1310
1550 REM
1560 REM Track to find root
1570 BA=BA+0-02*CH
1580 IFBA>=WWTHENBA=BA-001: C H = - 1
1590 IFM=1 GOTO 1340
1600 SS=SGN(F(M): ST=SGN(F(M~-1))
1610 IF SS=ST THEN F(M-1)=F(M): M=M~1: GOTO 1340
1620 C(l)=BA-0-005*CH
1630 C(2)=BA~0-01*CH
1640 V=0: T=T+1: M = 0
1650 GOTO 1310
1660 REM
1670 REM Lagrangian inverse interpolation
1680 H=C(M): IF M>= 20 GOTO 2040
1690 FOR 1=1 T O M - 1
1700 C(M)=(C(I)*F(M)-H*F(I))/(F(M)-F(I))
1710 H=C(M):NEXT
1720 IF H<0 THEN M = M - 1 : GOTO 2040
1730 C(M+1)=H
1740 IF ABS (C(M-1)-C(M))/ABS(C(M))>AC GOTO 1310
1750 CO=0: BA=H
1760 REM
1770 REM Attenuation
1780 RS=SQR(0-04*FR/RH)*PI
1790 Pi =2/(PI*PI!i:SD*SD)-(S+l)*(S+l)/2
1800 Pl=Pl-(WW*WW-l-N*N)/2
1810 A1=2*PI*RS*Z1*Z1*R1*R1/(WW*WW*PI*PI*144OO)
1820 P1=A1*P1
1830 P2 = A1*2*D*R/(PI*PI*R1*SD*SD)
1840 DE=-BA*N/(WW*FT)
1850 P3 = N*BA*DE/WW+FT
1860 P3 = DE*DE+P3*P3/K2
1870 P3-A1*WW*WW*P3*TT/(2*R1)
224 Program
1880 H=(D+TT)*BA/(2*R1*SIN(BA*D/(2*R1)))
1890 H=H*H
1900 P1 = H*P1:P2=H*P2
1910 PO=BA*(l+DE*DE)*(FT+K2/2-N*N/2+FT*FT/2)/WW
1920 PO=PO+N*DE*(1+BA*BA/(WW*WW))
1930 PO=PO*PI*WW*WW*Z1 *Z1 *R1 *R1 */(K2*K2* 120*PI)
1940 AL=4343*(P1+P2+P3)/(PO*(D+TT))
1950 REM
1960 REM Print
1970 PRINT" Frequency =";FR;"GHz"
1980 PRINT" k*rl =";WW
1990 PRINT" Beta*rl =";BA
2000 PRINT" Hybrid factor =";DE
2010 PRINT" Attenuation =";AL;"dB/m"
2020 PRINT
2030 GOTO 1110
2040 REM
2050 REM Modification
2060 CO=CO+1
2070 IF CO>2 GOTO 2110
2080 PRINT" MODIFICATION"
2090 BA=BA-0-002
2100 GOTO 1340
2110 REM
2120 REM Error
2130 PRINT" ERROR"
2140 GOTO 1090
2150 END
2160 REM Subroutine BESSEL(X,N,M,J)
2170 REM Computes J (X), dJ(X)/dX, Y(X),dY(X)/dX,
2180 REM Order N, when M = + l . IF M = - l the
2190 REM corresponding modified functions I & K
2200 REM are computed. Values are stored in Jl,J2
2210 REM J3,J4 in the above order.
2220 J3=0: J5=0: J6=0: J 2 = l : J7=—1
2230 FORI=2*INT(0-8*X+4) TO1STEP-1
2240 J1=2*I*J2/X~J3*MM: J3=J2: J2=J1
2250 IF ( - I ) I O I GOTO 2270
2260 J7=-J7*MM: J5=J7*J3/I+J5: J6=J6+J3: GOTO 2280
2270 IFMM=-1 THENJ6=J6+J3
2280 NEXTI
2290 J6=J6+J6+J1: J5=J5*J7: J7 = l/J6
2300 IFMM=-1THENJ7 = EXP(X)*J7
2310 J8=-0-7267604*MM+3-2732395
Program 225

2320 J6=LOG(X)-0-1159315: J6=0-25*Jl*J6+J5*MM


2330 J5=J6*J8*MM
2340 J1=J2*J7: J2=J3*J7: J3=J5*J7
2350 IFX*MM>=-1 GOTO 2440
2360 J5=0-0091894
2370 DATA-0-668098, 0-2184518,-04262633, 0-5575368
2380 DATA-0-5247277, 0-379241 ,-0-229985, 0-1344596
2390 DATA-0-0913909, 0-0881113,-0-1566642,1-2531414
2400 RESTORE 23 70: FOR I = 1 TO 12
2410 READJ6: J5=J5/X+J6
2420 NEXTI
2430 J3 = EXP(~0-5*LOG(X)-X)*J5
2440 J4=J3*J2-0-25*J8/X: J4=-J4*MM/J1: J2=-MM*J2
2450 IFN=OGOTO2510
2460 FORI = 1TON
2470 J5=J1:J8=J3
2480 J1=((I-1)*J5/X-J2)*MM: J3=(I-1)*J8/X-J4
2490 J2= J5-PJ1/X: J4=J8*MM-I*J3/X
2500 NEXTI
2510 RETURN
Program 2
1000 REM***Program to compute co-polar and
1010 REM cross-polar patterns from open-
1020 REM ended corrugated waveguide.
1030 REM Valid for krl large.
1040 PRINT
1050 PRINT" Radiation Pattern of Corrugated Waveguide"
1060 PRINT
1070 INPUT" Inner radius in Wavelengths =";R1
1080 KR=6-2832*R1
1090 INPUT"Slot depth in Wavelengths =";A
1100 PRINT
1110 KA=6-2832*A
1120 Y=1/TAN(KA)
1130 PRINT" Slot admittance =";Y
1140 U0=2-405
1150 U=U0*(l-Y/(2*KR))
1160 U2=U0*U0
1170 G = 1-U2*(Y/(2*KR)-Y*Y*U2*(4+U2)/(8*KR*KR))
1180 PRINT" Hybrid factor =";G: PRINT
1190 INPUT" Increment angle, Final angle =";DT,FT
1200 X=U:GOSUB 1400
1210 U0=J0:Ul=Jl
226 Program

1220 PR1NT:PRINT" ANGLE CO-POLAR X-POLAR"


1230 PRINT SPC(15);"(dB)";SPC(ll);"(dB)"
1240 FORT=0TOFTSTEPDT
1250 V=KR*SIN(T/57-295): X=V
1260 IFT=0THENZ=0: J0 = l: Jl=0: GOTO 1280
1270 GOSUB1400
1280 N2=-A*A/(V*V-U*U)
1290 NO=(V*UO*J1-U*JO*U1)*N2
1300 IF T=0 GOTO 1340
1310 N2=(V*(-UO+2*U1/U)*J1-U*(-JO+2*J1/V)*U1)*N2
1320 Z=(G-1)/(G+1)*N2
1330 XP=8-686*LOG(ABS(Z))
1340 CP=8-686*LOG(ABS(NO+Z))
1350 IFT=0THENCO=CP:XP=~100+CO
1360 CP=CP-CO: XP=XP-CO
1370 PRINT T,CP,XP
1380 NEXT
1390 GOTO 1040
1400 REM Subroutine to compute Bessel functions
1410 REMJ0=JO(X)andJl=Jl(X)
1420 REM
1430 J2=O;J3=O;J1 = 1
1440 FOR I=2*INT(0-8*X+4) TO 1 STEP-1
1450 JO = 2*I*J1/X~J2: J2=J1: J1=JO
1460 IF (-1)~I=1 THEN J3=J3+J2
1470 NEXTI
1480 J3=J3+J3+J0: J4 = l/J3
1490 J0=J0*J4: J1=J2*J4
1500 RETURN
Index

Airy, George 11 Compact Antenna Range 172


Anechoic chambers 171-175 Compact horns 49, 97-98,127-128,
Alignment with laser 174 155-157
Assessment of performance 175-177 Computer programs 222-226
Funnel 173 Conical c. waveguides 58-96,158-159
Instrumentation 175 (see also Wide flare angle horns)
Rectangular 173 Bandwidth 68
Aperture diameter (choice of) 142 Computer prediction 158-159
Attenuation in c. waveguides 2, 20, 40-44, Frequency dependence of crosspolar
143 radiation 68
Attenuation coefficient 41-44 Gain 68-69
Balanced-hybrid condition 43-44 Higher order mode generation 69-86
Power flow 41 Hybrid factor 61-62
Power loss 42 Modal characteristics 58-65
Poynting vector 41 Modal equation 61-63
Modal fields 60-61, 64-65
Balanced hybrid condition 16, 27-30, 62, 90, Propagation 58-65
105, 106, 157-158 Radiation by Kirchoff-Huygen method
Balanced hybrid frequency 115-116,119, 86-89
122,167 Radiation by Laguerre-Gaussian method
Beam waist of Gaussian beam 95-96 94-96
Beam waveguide feed system 97, 102 Radiation by SWEX method 65-69
Beamwidth design curves 107 Radiation by waveguide approximation
Bessel function computation 159-160 90-94
Bibliography 213-221 Copolar pattern measurement 177-180
Blockage loss 139 Copolar pattern with simple formula 167
Broadband horns 98, 153-154 Copolar radiation characteristics 100-114
Beamwidth design curves 107-109
Cassegrain antenna feeds 5, 20 Definition 101
Cassegrain reflector antennas 1, 97, 138, 141 Frequency behaviour 107
Casting of horns 165 Gaussian pattern 101-103
Characteristic equations for propagation Higher order modes 111-114
Conical c. horns 61—62 Near-field patterns 109-111
Cylindrical c. waveguides 25, 27-29, Normalised patterns 105
202-208 Optimum edge taper 102
Space harmonic 202—208 Corrugated horn design 141-156
Surface impedance 25, 27-29 Corrugations
Choked feeds 4 Geometry (choice of) 142
Choked flange 145 Influence on crosspolarisation 114—122
228 Index
Orientation 121-122 Phase 8,138
Shape 168-169 Sidelobe levels 5
Width 120 Spillover 5, 8,138,140
Crosspolar pattern measurement 177-180 Subreflector shape 7
Crosspolarisation Surface error 8
Bandwidth 115,121-122 Wide flare angle horns 138,140
Conical c. horns 64-65, 68-69, 78, 94 Electrical design parameters 99
Cylindrical c. waveguides 45-50 Electroforming 164-165
Cylindrical waveguide approximation Elliptical c. horns 2,10,193-198
90-94 Attenuation 193
Definition 101 Characteristic equation 197
Experimental results 120 Crosspolar peformance 193
Near-field patterns 122-123 Field components 195-196
Parametric study 117-122 History 193
Radiation characteristics 114-122 Mathieu equations 194-195
Sources of crosspolar power 108 Mathieu functions 195-198
Space harmonic model results 114—122 Propagation characteristics 198
Surface impedance model 114,120 Radiation characteristics 198
Cylindrical c. waveguides 20-57,143, Wave equation 194
160-161
Computer prediction 157-158
Far-field distance 109,172
Far-field test range 171
Design 97-162
Fields in a cylindrical waveguide 36-39
Design guidelines 142-143
Dielectric cone feed 1, 5 Five horn cluster 150
Dielectric waveguide feeds Flare angle (choice of) 142-143
Hybrid mode feed 9 Flare section 73-86, 99,122-128
Rectangular cross-section 10 Floquet's theorem 200
Dielguide 1 Focal fields of a reflector 10-19
Directivity Efficiency 14
Antenna 135 Four horn cluster 150
Circular aperture 135 Fourier Transform method 8, 21, 37,
Corrugated horns 135-137 43-55,109,143, 160, 190-191,
Maximum 136 210-212
Peak 135
Dual depth horns 2,122,154-155 G/T parameter 140-141
Dual hybrid mode horn 153 GTD22
Dual-shaped reflector 16, 138 Gain
Antenna 5,135
Earth-station antennas 5,16 Corrugated horns 135-137
Edge taper 102,138 Measurement 180
Efficiency Reflector antenna 137
Blockage 8 Standard (use as) 97,136
Crosspolarisation 8,138 Gaussian beam 94-96,102-103
Earth stations 5 Gaussian pattern 101
Focal fields 13 Gregorian subreflector 17, 97
Higher order mode improvement 55
Illumination 8,137-138,140 Higher order modes
Maximum 16-17 Attenuation 44
Microwave-relays 5 Conical horn radiation 68-69
Noise-temperature 5 Copolar patterns 111-114
Open-ended c. waveguides 138 Crosspolar patterns 114, 122
Optimum edge taper 102 Generation 69-86
Index 229

Generation with constant flare angle Microwave-relay applications 5


73-78,127 Modal expansion method 50
Generation with variable flare angle Mode conversion 114, 73-86,146
78-86,128,157 Multifrequency horns 97-98, 154-155
Improved efficiency with 54,138, Multimode horns and waveguides 13,16,
151-153 18,54-57,98,150-153
Matched feeds 55, 153 Efficiency improvement 54
Multimode waveguides and horns 54-57, Field equations 55-57
150-153 Tracking capability 54
Propagation characteristics 23-25,
62-63 NASA Deep Space Network antennas 154
Throat region 73,135,146 Narrow flare angle horns 49, 94, 97-98,
History of c. horns 1-3 141-143
Hybrid factor 25-32,61,116 (see also Cylindrical c. waveguides)
Hybrid-mode feeds 5-19 (see also Radiation from c. waveguides)
Balanced-hybrid condition 17 Attenuation 143
Bandwidth 9 Copolar patterns 103 — 114
Corrugated feeds 5, 9 Crosspolar patterns 114-120
Dielectric waveguides 9 Design procedure 142-143
Dielectric-cone feed 5 Numerical prediction 160 — 162
Impedance of boundary 7 Near-field patterns 89, 109-111,122-123
Near-field test range 172
INTELSAT-A earth stations 97,102 Noise temperature 5,140-141
Ideal feed 7, 137 Numerical prediction 157-162
Illumination loss 5 Bessel functions 159-160
Illumination of a reflector 5-10 Computer programs 159, 162
Conical c. waveguides 158 — 159
Junctions in c. waveguides 50-53, 128-135 Corrugated circular waveguide 157-158
Propagation characteristics 157 — 160
Kirchhoff-Huygen method 22, 43, 86-89, 96 Radiation characteristics 160-162
Comparison with SWEX method 87-89
OTS satellite 143
Laguerre-Gaussian method 58, 94-96
Legendre function approximation 62 Paraboloidal reflector, offset 16,55
Low noise systems 138 Paraboloidal reflector, prime focus 151
Ludwig erosspolarisation definition 46,101 Parkes radiotelescope 55
Pattern Comparison method 176 — 177
Machining of horns 163-164 Phase centre 109, 124-126,146
Manufacture of c. horns 163-169 Near-field 124,126
Casting 165 Profile horn 125
Electroforming 164—165 Wide flare horn 125
Machining 163-164 Phase pattern measurement 180-
Shape of corrugations 168-169 Potter horn 4, 54
Spark erosion 165 Prime focus feeds 20
Thermal contraction 166 Profile horns, 20, 49, 78-86, 97-98,
Tolerances 163,167-168 125-128,155-157
Matched feeds 2, 55, 111, 153 Mode conversion along 78-86
Measurement of horns Phase centre 125, 157
Copolar patterns 177-180 Propagation in conical c. waveguides 58-65
Crosspolar patterns 177-180 Modal equations 61—63
Gain 180 Modal fields 60-61, 64-65
Phase patterns 180 Propagation in cylindrical c. waveguide
Measurements 169-180 22-34,157-160,202-208
230 Index

Balanced-hybrid conditions 27-30 Satellite communications 2


Cut-off condition 25-26 Satellite-broadcast antennas 10
Fast-wave to slow-wave transition 30 Satellite-communication earth station 5
High-frequency cut-off 30-32 Scalar feeds 97
Numerical prediction 157—160 Separation of variables assumption 62
Short-circuit condition 30 Shaped beam antennas 10
Space harmonic analysis 202-208 Shaped pattern horns 151-153
Space harmonics (effect of) 32-34 Sidelobe levels 5,141
Surface impedance approach 24 Slots
Pyramidal horn 174 Geometry (choice of) 142
Influence on crosspolarisation 114-122
Quiet zone 174 Orientation 121-122
Shape 168-169
Radar absorbing material 174 Wide for millimeter wavelengths 120
Radar antennas 10,17 Small aperture horns 143-145
Ultra-low sidelobe 17 Space attenuation loss 151
Radiation from c. waveguides 44-50, Space harmonic analysis 199-212
160-162,210-212 Calculation of field components 209-210
Flange (effect of) 44 Computing procedure 208
Fourier Transform method 44-50, Field components 200-202
210-212 Propagation equation 202-208
GTD44 Radiation characteristics 210-212
Kirchhoff-Huygen method 44 Space-harmonics 20, 32-34, 114,199-212
Numerical prediction 160-162 Field in cylindrical waveguide 39
Radiation formulae 47 Spacecraft antennas 17, 57, 97
Radio astronomy antennas 1,16, 97,138 Spark erosion technique 165
Rayleigh, Lord 14 Spherical reflector 17-19
Rectangular c. horns 2, 10,181-193 Spherical wave expansion method (SWEX)
Attenuation 182 58-86,107,109,135
Balanced hybrid condition 185,187, 188, Spillover 138
191 Spillover loss 5
Beam width 191 Spillover noise 140-141
Crosspolarisation 192 Standard gain corrugated horn 136
Cut-off frequencies 188 Subreflector shape 7
E plane sidelobes 181 Surface error loss 139
Field components 184 Surface impedance representation 20, 24
Four-wall horn 181-182,188-190
High frequency cut-off 188 Test ranges 170-177
History 2,181-182 Anechoic chamber design 172-175
Horn Hare (effect of) 192 Assessment of performance 175-177
Magnetic wall 185 Types 171-172
Manufacture 181 Testing of corrugated horns 169-180
Mode nomenclature 182 Thermal contraction 166
Propagation characteristics 187-188, 190 Throat region of horn 735 99,128-135,143,
Radiation characteristics 190-193 146-148,154
Square horn 190-191 Design procedure 128-135
Surface impedance 185 Frequency band 133
Throat region 181, 182,187,192 Mode conversion 133
Two-walled horn 181-188 Mode generation 73
VSWR characteristics 193 Non-linear profile 132
Reflectivity level 172, 176 Reflection coefficient 129
Ring-loaded slots 133-135 Slot geometry 131-133
Index 231

Tracking modes 135 Wide flare angle horns


Tolerances 163,167-168 (see also Conical c. waveguides)
Tracking horns 150-151 Copolar patterns 103-111
Corrugation geometry 146
, t , , n, Crosspolar patterns 114,120-121
VSWR method of test range assessment 176 p^. . 146
Mode conversion 77
Waveguide approximation for horns 58, Numerical prediction 158-159,161-162
90 -94,161 Phase centre 125,146
Spherical phase factor 92 Throat region 146-148
Corrugated horns for microwave antennas

P.J.B. Clarricoats and A.D. Olver

About the book


Oyer the last twenty years Corrugated horns have become widely used as
feeds for microwave reflector antennas because of their high efficiency, good
pattern symmetry and low cross-polarisation. They are increasingly used in
antennas for satellite communications, radar, microwave point to point
communications and radio astronomy.

The authors draw on more than fifteen years experience to write the first
book to appear devoted to the theory and design of corrugated horns and
scalar feeds.

The book explains why hybrid mode feeds are ideal feeds for reflectors. The
propagation and radiation behaviour of cylindrical corrugated wave­guides,
narrow flare angle conical corrugated horns and wide flare angle corrugated
horns are described. Factors relevant to the practical design of corrugated
horns are discussed. Other topics treated include the measure­ment of
corrugated horns.

About the authors


Professor Peter Clarricoats is the Head of the Department of Electrical and
Electronic Engineering at Queen Mary College in the University of London. He
is a past Chairman of the Electronics Division of the IEE and has also chaired
the European Microwave Conference and the IEE International Conference on
Antennas and Propagation. He has been an Editor of Electronics Letters since
its foundation in 1965 and is one of the Editors of the Electromagnetic Wave
Series in which this volume appears.

Dr David Olver is Reader in Electrical and Electronic Engineering at Queen


Mary College. He is Chairman of the IEE Professional Group on Antennas and
Propagation and is an Editor of IEE Proceedings ‘Microwaves Optics and
Antennas’. He was one of the Editors of ‘The Handbook of Antenna Design’.

Peter Clarricoats and David Olver share responsibility for a Group which has
been active in advanced antenna design for more than 15 years. During this
time they have made notable contributions to the design of corrugated horns
and other hybrid mode feeds.

Peter Peregrinus Ltd.,


Southgate House, Stevenage, Herts SG1 1HQ, England
ISBN 0 86341 003 0
Printed in the United Kingdom

You might also like